Physical Chemistry for the Life Sciences 2e [PDF] | Online Book Share (2024)

618 Pages • 315,887 Words • PDF • 33.3 MB

+ Life+ Sciences+ Chemistry+ Physical

Uploaded at 2021-07-06 07:15

Report DMCA

PREVIEW PDF

This page intentionally left blank

Physical Chemistry for the Life Sciences

Library of Congress Number: 2010940703 © 2006, 2011 by P.W. Atkins and J. de Paula All rights reserved. Printed in Italy by L.E.G.O. S.p.A First printing Published in the United States and Canada by W. H. Freeman and Company 41 Madison Avenue New York, NY 10010 www.whfreeman.com ISBN-13: 978-1-4292-3114-5 ISBN-10: 1-4292-3114-9 Published in the rest of the world by Oxford University Press Great Clarendon Street Oxford OX2 6DP United Kingdom www.oup.com ISBN: 978-0-19-956428-6

Physical Chemistry for the Life Sciences Second edition

Peter Atkins Professor of Chemistry, Oxford University

Julio de Paula Professor of Chemistry, Lewis & Clark College

W. H. Freeman and Company New York

This page intentionally left blank

Contents in brief Prolog Fundamentals

xxi 1

PART 1 Biochemical thermodynamics

21

1 2 3 4 5

The First Law The Second Law Phase equilibria Chemical equilibrium Thermodynamics of ion and electron transport

23 69 94 135 181

PART 2 The kinetics of life processes

217

6 The rates of reactions 7 Accounting for the rate laws 8 Complex biochemical processes

219 243 273

PART 3 Biomolecular structure 9 Microscopic systems and quantization 10 The chemical bond 11 Macromolecules and self-assembly

PART 4 Biochemical spectroscopy 12 Optical spectroscopy and photobiology 13 Magnetic resonance

311 313 364 407

461 463 514

Resource section 1 Atlas of structures 2 Units 3 Data

546 558 560

Answers to odd-numbered exercises Index of Tables Index

573 577 579

This page intentionally left blank

Full contents Prolog

xxi

The structure of physical chemistry

xxi

(a)

The organization of science

xxi

(b)

The organization of our presentation

xxii

Applications of physical chemistry to biology and medicine (a)

Techniques for the study of biological systems

xxii

1.4

The measurement of heat

32

(a) Heat capacity

33

(b) The molecular interpretation of heat capacity

34

Internal energy and enthalpy 1.5

The internal energy (a) Changes in internal energy

xxii

34 35 35

(b) Protein folding

xxiii

Example 1.1 Calculating the change in internal energy

(c)

Rational drug design

xxv

(b) The internal energy as a state function

37

(d)

Biological energy conversion

xxv

(c) The First Law of thermodynamics

38

1.6 The enthalpy

36

38

Fundamentals

1

(a) The definition of enthalpy

39

F.1 Atoms, ions, and molecules

1

(b) Changes in enthalpy

39

(c) The temperature dependence of the enthalpy

41

(a)

Bonding and nonbonding interactions

1

(b)

Structural and functional units

2

(c)

Levels of structure

3

F.2 Bulk matter (a)

States of matter

4 5

(c)

8

F.3 Energy (a)

Varieties of energy

(b)

The Boltzmann distribution

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

(a) Bomb calorimeters

4

(b) Physical state Equations of state

In the laboratory 1.1 Calorimetry

Example 1.2 Calibrating a calorimeter and measuring the energy content of a nutrient (b) Isobaric calorimeters (c) Differential scanning calorimeters

10 11 13

17 17 18 18 19

Physical and chemical change 1.7

21

1 The First Law

23

The conservation of energy

23

1.1

Systems and surroundings

24

1.2

Work and heat

25

44

46 46 46

Thermochemical properties of fuels

The combination of reaction enthalpies

Example 1.4 Using Hess’s law Standard enthalpies of formation

Example 1.5 Using standard enthalpies of formation

49

51 52

55 57

58 58

59

25

1.12

Enthalpies of formation and computational chemistry

61

(b) The molecular interpretation of work and heat

26

1.13

The variation of reaction enthalpy with temperature

62

(c) The molecular interpretation of temperature

26

(a) Exothermic and endothermic processes

Case study 1.1 Energy conversion in organisms 1.3

44

47

Case study 1.2 Biological fuels

1.11

43

(b) Enthalpies of vaporization, fusion, and sublimation

Example 1.3 Using mean bond enthalpies

1.10

42

(a) Phase transitions 1.8 Bond enthalpy

1.9

PART 1 Biochemical thermodynamics

Enthalpy changes accompanying physical processes

42

27

The measurement of work

29

(a) Sign conventions

29

(b) Expansion work

30

(c) Maximum work

31

Example 1.6 Using Kirchhoff’s law

63

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

64 65 65 65 68

viii

FULL CONTENTS

2 The Second Law

69

Entropy

70

2.1

The direction of spontaneous change

70

2.2

Entropy and the Second Law

71

(a) The definition of entropy

71

2.3

2.5

112

(c) The chemical potential of a solute

114

Example 3.2 Determining whether a natural water can support aquatic life

116

Case study 3.2 Gas solubility and breathing

117

73

(c) The entropy change accompanying a phase transition

75

(d) Real solutions: activities

118

(d) Entropy changes in the surroundings

77

Case study 3.3 The Donnan equilibrium

119

Absolute entropies and the Third Law of thermodynamics

77

Example 3.3 Analyzing a Donnan equilibrium

121

78

The molecular interpretation of the Second and Third Laws

80

(a) The Boltzmann formula

80

(b) The relation between thermodynamic and statistical entropy

81

(c) The residual entropy

82

Entropy changes accompanying chemical reactions

82

(a) Standard reaction entropies

82

(b) The spontaneity of chemical reactions

83

The Gibbs energy 2.6

112

(b) The chemical potential of a solvent

(b) The entropy change accompanying heating

In the laboratory 2.1 The measurement of entropies 2.4

(a) The chemical potential of a gas

84

Focusing on the system

84

(a) The definition of the Gibbs energy

84

(b) Spontaneity and the Gibbs energy

Case study 2.1 Life and the Second Law

85

85

(e) The thermodynamics of dissolving

Colligative properties 3.9

The modification of boiling and freezing points

3.10 Osmosis

121

122 123 125

In the laboratory 3.1 Osmometry

127

Example 3.4 Determining the molar mass of an enzyme from measurements of the osmotic pressure

127

Checklist of key concepts Checklist of key equations Further information 3.1 The phase rule Further information 3.2 Measures of concentration

128 129 129 130

Example 3.5 Relating mole fraction and molality

131 132 132 134

2.7

The hydrophobic interaction

86

2.8

Work and the Gibbs energy change

88

Discussion questions Exercises Projects

Example 2.1 Estimating a change in Gibbs energy for a metabolic process

89

4 Chemical equilibrium

135

Case study 2.2 The action of adenosine triphosphate

90

Thermodynamic background

135

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

90 91 91 91 92

3 Phase equilibria

94

The thermodynamics of transition

94

4.1

The reaction Gibbs energy

135

4.2

The variation of ΔrG with composition

137

(a) The reaction quotient

137

Example 4.1 Formulating a reaction quotient (b) Biological standard states

Example 4.2 Converting between thermodynamic and biological standard states 4.3

142

94

(a) The significance of the equilibrium constant

3.2

The variation of Gibbs energy with pressure

95

(b) The composition at equilibrium

3.3

The variation of Gibbs energy with temperature

98

(b) The location of phase boundaries

101

(c) Characteristic points

103

(d) The phase diagram of water

105

Phase transitions in biopolymers and aggregates 3.5

The stability of nucleic acids and proteins

Example 3.1 Predicting the melting temperature of DNA 3.6

99 100

Phase transitions of biological membranes

Case study 3.1 The use of phase diagrams in the study of proteins The thermodynamic description of mixtures

106 106

140 140

The condition of stability

(a) Phase boundaries

139

Reactions at equilibrium

3.1

3.4 Phase diagrams

138

143

Example 4.3 Calculating an equilibrium composition

143

(c) The molecular origin of chemical equilibrium

144

Case study 4.1 Binding of oxygen to myoglobin and hemoglobin

144

4.4

The standard reaction Gibbs energy

Example 4.4 Calculating the standard reaction Gibbs energy of an enzyme-catalyzed reaction

146

146

107

(a) Standard Gibbs energies of formation

147

108

(b) Stability and instability

149

The response of equilibria to the conditions 109

4.5

The presence of a catalyst

4.6

The effect of temperature

149 150 150

110

3.7

The chemical potential

110

3.8

Ideal and ideal–dilute solutions

111

Coupled reactions in bioenergetics

151

Case study 4.2 ATP and the biosynthesis of proteins

152

FULL CONTENTS

Case study 4.3 The oxidation of glucose

153

Proton transfer equilibria

156

ix

Example 5.4 Converting a standard potential to a biological standard value

200

4.7 Brønsted–Lowry theory

156

In the laboratory 5.1 Ion-selective electrodes

201

4.8

Protonation and deprotonation

157

Applications of standard potentials

202

(a) The strengths of acids and bases

158

(b) The pH of a solution of a weak acid

161

Example 4.5 Estimating the pH of a solution of a weak acid (c) The pH of a solution of a weak base

161 163

Example 4.6 Estimating the pH of a solution of a weak base

163

(d) The extent of protonation and deprotonation

163

(e) The pH of solutions of salts

164

4.9 Polyprotic acids

165

Case study 4.4 The fractional composition of a solution of lysine

166

(a) The fractional composition of amino acid solutions (b) The pH of solutions of amphiprotic anions 4.11 Buffer solutions

170

172

Case study 4.5 Buffer action in blood

173

Checklist of key concepts Checklist of key equations Further information 4.1 The contribution of autoprotolysis to pH Further information 4.2 The pH of an amphiprotic salt solution Discussion questions Exercises Projects

174 175 175 176 177 177 180

5 Thermodynamics of ion and electron transport

181

Transport of ions across biological membranes

181

5.1

5.2

205

(c) Calculation of the standard reaction entropy and enthalpy

206

The electrochemical series

207

Electron transfer in bioenergetics

207

5.9

5.10

The respiratory chain

207

(a) Electron transfer reactions

208

(b) Oxidative phosphorylation

208

5.11 Plant photosynthesis

Checklist of key concepts Checklist of key equations Discussion questions Exercises Project

209

211 212 212 212 215

PART 2 The kinetics of life processes

217

6 The rates of reactions

219

Reaction rates

219

In the laboratory 6.1 Experimental techniques

219

(a) The determination of concentration

219

(b) Monitoring the time dependence

220

6.1

The definition of reaction rate

221

(a) Activity coefficients

182

6.2

Rate laws and rate constants

223

(b) Debye–Hückel theory

184

Passive and active transport of ions across biological membranes

Ion channels and ion pumps

Redox reactions 5.4 Half-reactions

224

The determination of the rate law

225

(a) Isolation and pseudo-order reactions

225

(b) The method of initial rates

226

187 188

188

Example 6.1 Using the method of initial rates 6.5

Integrated rate laws

227 228

189

(a) Zeroth-order reactions

228

190

(b) First-order reactions

228

(c) Second-order reactions

231

190

Example 5.3 Writing the reaction quotient for a half-reaction

191

Reactions in electrochemical cells

192

(a) Galvanic and electrolytic cells

192

(b) Varieties of electrodes

194

(c) Electrochemical cell notation

194

The Nernst equation

195

5.7 Standard potentials

6.3 Reaction order 6.4

186

Example 5.2 Expressing a reaction in terms of half-reactions

5.6

205

181

Case study 5.1 Action potentials

5.5

204

Ions in solution

Example 5.1 Estimating a membrane potential 5.3

203

Example 5.6 Calculating a standard potential from two other standard potentials

169

Example 4.8 Assessing buffer action

202

(a) Calculation of the equilibrium constant

(b) Calculation of standard potentials

169 169

The determination of thermodynamic functions

Example 5.5 Calculating the equilibrium constant of a biological electron transfer reaction

164

Example 4.7 Calculating the concentration of carbonate ion in carbonic acid

4.10 Amphiprotic systems

5.8

197

(a) Thermodynamic standard potentials

198

(b) Variation of potential with pH

198

(c) The biological standard potential

200

Case study 6.1 Pharmaco*kinetics

234

The temperature dependence of reaction rates

235

6.6

The Arrhenius equation

Example 6.2 Determining the Arrhenius parameters 6.7

Preliminary interpretation of the Arrhenius parameters

Checklist of key concepts Checklist of key equations Discussion questions Exercises Project

235

236 237

239 239 239 240 242

x

FULL CONTENTS

7 Accounting for the rate laws Reaction mechanisms 7.1

243 243

Example 8.3 The isoelectric point of a protein 8.8

293

Transport across ion channels and ion pumps

294

(a) The potassium channel

294

The approach to equilibrium

243

(a) The relation between equilibrium constants and rate constants

243

(b) The time-dependence of the approach to equilibrium

245

8.9

The rates of electron transfer processes

245

8.10

The theory of electron transfer processes

298

8.11

Experimental tests of the theory

299

8.12

The Marcus cross-relation

300

In the laboratory 7.1 Relaxation techniques in biochemistry 7.2 Elementary reactions

247

7.3 Consecutive reactions

249

(b) The proton pump

Electron transfer in biological systems

295

296 296

(a) The variation of concentration with time

249

Example 8.4 Using the Marcus cross-relation

302

(b) The rate-determining step

251

Checklist of key concepts Checklist of key equations Further information 8.1 Fick’s laws of diffusion

303 303 304

1. Fick’s first law of diffusion

304

Example 7.1 Identifying a rate-determining step

252

(c) The steady-state approximation

252

(d) Pre-equilibria

253

Case study 7.1 Mechanisms of protein folding and unfolding

254

7.4 Diffusion control

256

7.5

258

Kinetic and thermodynamic control

Reaction dynamics

304

Discussion questions Exercises Projects

305 305 308

259

7.6 Collision theory

259

7.7

Transition state theory

261

(a) Formulation of the theory

261

(b) Thermodynamic parameterization

262

In the laboratory 7.2 Time-resolved spectroscopy for kinetics

2. Fick’s second law

263

PART 3 Biomolecular structure

311

9 Microscopic systems and quantization

313

Principles of quantum theory

313

The emergence of the quantum theory

314

264

(a) Atomic and molecular spectra

314

Example 7.2 Analyzing the kinetic salt effect

266

(b) Wave–particle duality

315

Checklist of key concepts Checklist of key equations Further information 7.1 Collisions in the gas phase

267 267 267

7.8

The kinetic salt effect

(a) The kinetic model of gases

267

(b) The Maxwell distribution of speeds

268

(c) Molecular collisions

9.1

Example 9.1 Estimating the de Broglie wavelength of electrons

316

In the laboratory 9.1 Electron microscopy

317

9.2

269

The Schrödinger equation

318

(a) The formulation of the equation

319

(b) The interpretation of the wavefunction

320

Discussion questions Exercises Projects

270 270 272

Example 9.2 Interpreting a wavefunction

Example 9.3 Using the uncertainty principle

322

8 Complex biochemical processes

273

Applications of quantum theory

323

Enzymes 8.1

The Michaelis–Menten mechanism of enzyme catalysis

Example 8.1 Analyzing a Lineweaver–Burk plot 8.2

8.3

273 276 277

(a) Sequential reactions

277

(b) Ping-pong reactions

278

The catalytic efficiency of enzymes

279

(a) Motion in one dimension

Case study 9.1 The electronic structure of b-carotene (b) Tunneling

In the laboratory 9.2 Scanning probe microscopy (c) Motion in two dimensions 9.5 Rotation

280

Example 8.2 Distinguishing between types of inhibition

282

Case study 8.1 The molecular basis of catalysis by hydrolytic enzymes

284

Transport across biological membranes

285

The uncertainty principle

9.4 Translation

274

The analysis of complex mechanisms

8.4 Enzyme inhibition

9.3

(a) A particle on a ring

Case study 9.2 The electronic structure of phenylalanine (b) A particle on a sphere 9.6 Vibration

320 321

324 324

327 328

329 330 331 332

334 334 335

8.5

Molecular motion in liquids

285

Case study 9.3 The vibration of the N–H bond of the peptide link

336

8.6

Molecular motion across membranes

288

Hydrogenic atoms

337

8.7

The mobility of ions

290

9.7

291

9.8 Atomic orbitals

In the laboratory 8.1 Electrophoresis

The permitted energy levels of hydrogenic atoms

338 339

FULL CONTENTS

(a) Shells and subshells

340

(b) The shapes of s orbitals

341

(a) Ethene

387

(c) The shapes of p orbitals

344

(b) Benzene

390

(d) The shapes of d orbitals

345

The structures of many-electron atoms 9.9

346

The orbital approximation and the Pauli exclusion principle

346

9.10

Penetration and shielding

348

9.11

The building-up principle

349

(a) Neutral atoms

349

9.12

(a) Crystal field theory

Example 10.3 Low- and high-spin complexes of Fe(II) in hemoglobin

391 392 392

394

351

(b) Ligand-field theory: s bonding

394

(c) Ligand-field theory: p bonding

396

(a) Atomic and ionic radii

352

(b) Ionization energy

353

(c) Electron affinity

355

357 358 358 359 359 360 360 363

10 The chemical bond

364

Valence bond theory

365

10.1 Diatomic molecules

365

(a) Formulation of the VB wavefunction

365

(b) The energy of interaction

366

(c) s and p bonds

366

10.2 Polyatomic molecules

367

(a) Promotion

368

(b) Hybridization

368

Example 10.1 Bonding in the peptide group

371

(c) Resonance

372

(d) The language of valence bonding

372

Molecular orbital theory

373

Linear combinations of atomic orbitals

373

(a) Bonding orbitals

373

(b) Antibonding orbitals

374

(c) Inversion symmetry

375

hom*onuclear diatomic molecules (a) Criteria for the formation of molecular orbitals

Example 10.2 Assessing the contribution of d orbitals

375 376

Case study 10.4 Ligand-field theory and the binding of O2 to hemoglobin

397

Computational biochemistry

398

10.9 Computational techniques

398

(a) Semi-empirical methods

399

(b) Density functional theory

399

(c) Ab initio methods

400

10.10 Graphical output

400

10.11 The prediction of molecular properties

400

(a) Electrochemical properties

401

(b) Spectroscopic properties

401

(c) Chemical reactivity

402

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

402 403 403 404 406

11 Macromolecules and self-assembly

407

Determination of size and shape

407

11.1 Ultracentrifugation

407

(a) The sedimentation rate

408

(b) Sedimentation equilibrium

409

Example 11.1 The molar mass of a protein from ultracentrifugation experiments

409

11.2 Mass spectrometry

410

11.3

Laser light scattering

412

(a) Rayleigh scattering

412

Example 11.2 Determining the molar mass and size of a protein by laser light scattering (b) Dynamic light scattering

413 414

11.4 X-ray crystallography

414

(a) Diffraction

415

378

(b) Crystal systems

415

(b) The hydrogen molecule

379

(c) Crystal planes

416

(c) Many-electron hom*onuclear diatomic molecules

380

(d) Bond order

382

Case study 10.1 The biochemical reactivity of O2 and N2 Heteronuclear diatomic molecules

382 384

(a) Polarity and electronegativity

384

(b) Molecular orbitals in heteronuclear species

385

Case study 10.2 The biochemistry of NO 10.6

10.8 d-Metal complexes

352

Checklist of key concepts Checklist of key equations Further information 9.1 A justification of the Schrödinger equation Further information 9.2 The separation of variables procedure Further information 9.3 The Pauli principle Discussion questions Exercises Projects

10.5

Case study 10.3 The unique role of carbon in biochemistry

387

(b) Cations and anions

356

10.4

10.7 Hückel theory

Three important atomic properties

Case study 9.4 The biological role of Zn2+

10.3

xi

The structures of polyatomic molecules

Example 11.3 Using the Miller indices (d) Bragg’s law

418 419

Example 11.4 Using Bragg’s law

419

(e) Fourier synthesis

420

386

Example 11.5 Calculating an electron density by Fourier synthesis

420

387

In the laboratory 11.1 The crystallization of biopolymers

421

xii

FULL CONTENTS

In the laboratory 11.2 Data acquisition in X-ray crystallography

422

(b) Stimulated and spontaneous transitions

470

(c) Populations and intensities

471

(d) Linewidths

472

Case study 11.1 The structure of DNA from X-ray diffraction studies

423

The control of shape

424

In the laboratory 12.2 Biosensor analysis

473

Vibrational spectra

474

11.5

Interactions between partial charges

425

11.6

Electric dipole moments

426

Example 11.6 Calculating the dipole moment of the peptide group

428

11.7

Interactions between dipoles

429

11.8

Induced dipole moments

431

(a) Dipole–induced-dipole interactions

432

(b) Dispersion interactions

432

11.9 Hydrogen bonding

433

11.10 The total interaction

435

12.3

Example 12.2 The effect of isotopic substitution on the vibrational frequency of O2 12.4 Vibrational transitions (a) Infrared transitions

Example 12.3 Identifying species that contribute to climate change 12.5

Case study 11.2 Molecular recognition in biology and pharmacology

437

Levels of structure

438

11.11 Minimal order: gases and liquids

438

11.12 Random coils

440

The vibrations of diatomic molecules

474

475 476 476

476

(b) Raman transitions

478

The vibrations of polyatomic molecules

478

(a) Normal modes

479

(b) Infrared transitions

480

(c) Raman transitions

481

Case study 12.1 Vibrational spectroscopy of proteins

482

(a) Measures of size

440

In the laboratory 12.3 Vibrational microscopy

483

(b) Conformational entropy

441

Ultraviolet and visible spectra

485

11.13 Proteins

442

12.6

(a) The secondary structure of a protein

442

12.7 Chromophores

487

(b) Higher-order structures of proteins

445

12.8

488

11.14 Nucleic acids

446

11.15 Polysaccharides

448

11.16 Micelles and biological membranes

449

(a) Micelles

449

(b) Bilayers, vesicles, and membranes

450

(c) Interactions between proteins and biological membranes 450 11.17 Computer-aided simulation

Optical activity and circular dichroism

Radiative and non-radiative decay 12.9 Fluorescence

451

(b) Molecular dynamics and Monte Carlo simulations

451

490 490

12.10 Phosphorescence

491

In the laboratory 12.4 Fluorescence microscopy

492

In the laboratory 12.5 Single-molecule spectroscopy

493

Photobiology

494

12.11 The kinetics of decay of excited states

494

453

12.12 Fluorescence quenching (a) The experimental analysis

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

486

451

(a) Molecular mechanics calculations (c) Quantitative structure–activity relationships

The Franck–Condon principle

455 456 457 457 460

Example 12.4 Determining the quenching rate constant (b) Mechanisms of quenching

497 497

498 499

12.13 Fluorescence resonance energy transfer

500

Case study 12.2 Vision

501

Case study 12.3 Photosynthesis

503

PART 4 Biochemical spectroscopy

461

Case study 12.4 Damage of DNA by ultraviolet radiation

504

12 Optical spectroscopy and photobiology

463

Case study 12.5 Photodynamic therapy

505

General features of spectroscopy

463

In the laboratory 12.1 Experimental techniques

464

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

507 508 508 508 511

467

13 Magnetic resonance

514

(b) The determination of concentration

468

Principles of magnetic resonance

514

The intensities of transitions: theoretical aspects

469

13.1

Electrons and nuclei in magnetic fields

515

(a) The transition dipole moment

469

13.2

The intensities of NMR and EPR transitions

517

12.1

The intensities of spectroscopic transitions: empirical aspects

466

(a) The Beer–Lambert law

466

Example 12.1 The molar absorption coefficient of tryptophan 12.2

FULL CONTENTS

The information in NMR spectra

xiii

519

The information in EPR spectra

537

The chemical shift

519

13.10 The g-value

538

(a) The d scale

520

13.11 Hyperfine structure

539

(b) Contributions to the shift

521

The fine structure

522

Example 13.3 Predicting the hyperfine structure of an EPR spectrum

540

(a) The appearance of fine structure

522

In the laboratory 13.3 Spin probes

540

Example 13.1 Accounting for the fine structure in a spectrum

524

Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

541 542 543 543 545

Resource section 1 Atlas of structures 2 Units 3 Data

546 558 560

Answers to odd-numbered exercises Index of Tables Index

573 577 579

13.3

13.4

13.5

(b) The origin of fine structure

525

Conformational conversion and chemical exchange

527

Example 13.2 Interpreting line broadening

527

Pulse techniques in NMR

528

13.6

Time- and frequency-domain signals

13.7 Spin relaxation

In the laboratory 13.1 Magnetic resonance imaging

528 530

531

13.8 Proton decoupling

533

13.9

533

The nuclear Overhauser effect

In the laboratory 13.2 Two-dimensional NMR

535

Case study 13.1 The COSY spectrum of isoleucine

536

Preface The second edition of this text—like the first edition—seeks to present all the material required for a course in physical chemistry for students of the life sciences, including biology and biochemistry. To that end we have provided the foundations and biological applications of thermodynamics, kinetics, quantum theory, and molecular spectroscopy. The text is characterized by a variety of pedagogical devices, most of them directed toward helping with the mathematics that must remain an intrinsic part of physical chemistry. One such new device is the Mathematical toolkit, a boxed section that—as we explain in more detail in the ‘About the book’ section below— reviews concepts of mathematics just where they are needed in the text. Another device that we continue to invoke is A note on good practice. We consider that physical chemistry is kept as simple as possible when people use terms accurately and consistently. Our Notes emphasize how a particular term should and should not be used (by and large, according to IUPAC conventions). Finally, new to this edition, each chapter ends with a Checklist of key concepts and a Checklist of key equations, which together summarize the material just presented. The latter is annotated in many places with short comments on the applicability of each equation. Elements of biology and biochemistry continue to be incorporated in the text’s narrative in a number of ways. First, each numbered section begins with a statement that places the concepts of physical chemistry about to be explored in the context of their importance to biology. Second, the narrative itself shows students how physical chemistry gives quantitative insight into biology and biochemistry. To achieve this goal, we make generous use of A brief illustration sections (by which we mean quick numerical exercises) and Worked examples, which feature more complex calculations than do the illustrations. Third, a unique feature of the text is the use of Case studies to develop more fully the application of physical chemistry to a specific biological or biomedical problem, such as the action of ATP, pharmaco*kinetics, the unique role of carbon in biochemistry, and the biochemistry of nitric oxide. Finally, the new In the laboratory sections highlight selected experimental techniques in modern biochemistry and biomedicine, such as differential scanning calorimetry, gel electrophoresis, electron microscopy, and magnetic resonance imaging. All the illustrations (nearly 500 of them) have been redrawn and are now in full color. Another innovation in this edition is the Atlas of structures, in the Resource section at the end of the book. Many biochemically important structures are referred to a number of times in the text, and we judged it appropriate and convenient to collect them all in one place. The Resource section also includes data used in a variety of places in the text.

PREFACE

xv

A text cannot be written by authors in a vacuum. To merge the languages of physical chemistry and biochemistry we relied on a great deal of extraordinarily useful and insightful advice from a wide range of people. We would particularly like to acknowledge the following people, who reviewed draft chapters of the text: Professor Björn Åkerman, Chalmers University of Technology

Professor Tim Keiderling, University of Illinois at Chicago

Dr Perdita Barran, University of Edinburgh

Dr Paul King, Birkbeck College

Professor Bo Carlsson, University of Kalmar

Professor Krzysztof Kuczera, University of Kansas

Dr Monique Cosman, California State University, East Bay

Professor H.E. Lundager Madsen, University of Copenhagen

Dr Erin E. Dahlke, Loras College

Dr Jeffrey Mack, California State University, Sacramento

Prof Roger DeKock, Calvin College

Dr Jeffry Madura, Duquesne University

Professor Steve Desjardins, Washington and Lee University

Dr John Marvin, Brescia University

Dr Bridgette Duncombe, University of Edinburgh

Dr Stephen Mezyk, California State University, Long Beach

Dr Niels Engholm Henriksen, Technical University of Denmark

Dr Yorgo Modis, Yale University

Professor Andrew Fisher, University of California, Davis

Dr Brent Ridley, Biola University

Dr Peter Gardner, Royal Holloway University of London

Dr Martha Sarasua, University of West Florida

Dr Anton Guliaev, San Francisco State University Dr Magnus Gustafsson, University of Gothenburg Dr Hal Harris, University of Missouri- St. Louis Dr Lars Hemmingsen, Copenhagen University Dr Hans A. Heus, Radboud University Nijmegen Dr Martina Huber, Leiden University Dr Eihab Jaber, Worcester State College Dr Ryan R. Julian, University of California, Riverside

Dr Lee Reilly, University of Warwick Dr Jens Risbo, University of Copenhagen Prof Steve Scheiner, Utah State University Dr Andrew Shaw, University of Exeter Dr Suzana K. Straus, University of British Columbia Dr Cindy Tidwell, University of Montevallo Professor Geoff Thornton, University College London Dr Andreas Toupadakis, University of California, Davis Dr Jeffrey Watson, Gonzaga University Dr Andrew Wilson, University of Leeds

We have been particularly well served by our publishers, and would wish to acknowledge our gratitude to our editors Jonathan Crowe of Oxford University Press and Jessica Fiorillo of W.H. Freeman and Company, who helped us achieve our goal. We also thank Valerie Walters for proofreading the text so carefully and Charles Trapp and Marshall Cady for compiling the solutions manual and making very helpful comments in the course of its development. PWA, Oxford

JdeP, Portland

About the book Numerous features in this text are designed to help you learn physical chemistry and its applications to biology, biochemistry, and medicine. One of the problems that makes the subject so daunting is the sheer amount of information. To help with that problem, we have introduced several devices for organizing the material in your mind: see Organizing the information. We appreciate that mathematics is often troublesome, and therefore have included several devices for helping you with this enormously important aspect of physical chemistry: see Mathematics support. Problem solving, especially, ‘where do I start?’, is often a problem, and we have done our best to help you find your way over the first hurdle: see Problem solving. Finally, the web is an extraordinary resource, but you need to know where to go for a particular piece of information; we have tried to point you in the right direction: see Using the Web. The following paragraphs explain the features in more detail.

Organizing the information Equation and concept tags The most significant equations and concepts—

and which we urge you to make a particular effort to remember—are flagged with an annotation, as shown here.

Checklist of key concepts Here we collect together the major concepts that

we have introduced in the chapter. You might like to check off the box that precedes each entry when you feel that you are confident about the topic.

Checklist of key equations This is a collection of the most important equations introduced in the chapter.

Case studies We incorporate general concepts of biology and biochemistry

throughout the text, but in some cases it is useful to focus on a specific problem in some detail. A Case study contains some background information about a biological process, such as the action of adenosine triphosphate or the metabolism of drugs, and may be followed by a series of calculations that give quantitative insight into the phenomena.

ABOUT THE BOOK

In the laboratory Here we describe some of the modern techniques of biology,

biochemistry, and medicine. In many cases, you will use these techniques in laboratory courses, so we focus not on the operation of instruments but on the physical principles that make the instruments perform a specific task.

Notes good practice Science is a precise activity, and using its language

accurately can help you to understand the concepts. We have used this feature to help you to use the language and procedures of science in conformity to international practice and to avoid common mistakes.

Justifications On first reading you might need the ‘bottom line’ rather than a detailed development of a mathematical expression. However, once you have collected your thoughts, you might want to go back to see how a particular expression was obtained. The Justifications let you adjust the level of detail that you require to your current needs. However, don’t forget that the development of results is an essential part of physical chemistry, and should not be ignored. Further information In some cases, we have judged that a derivation is too long, too detailed, or too different in level for it to be included in the text. In these cases, you will find the derivation at the end of the chapter.

Mathematics support A brief comment A topic often needs to draw on a mathematical procedure or

a concept of physics; a brief comment is a quick reminder of the procedure or concept.

Mathematical toolkit It is often the case that you need a more full-bodied account of a mathematical concept, either because it is important to understand the procedure more fully or because you need to use a series of tools to develop an equation. The Mathematical toolkit sections are located in the chapters, primarily where they are first needed.

Problem solving Brief illustrations A Brief illustration (don’t confuse this with a diagram!) is a short example of how to use an equation that has just been introduced in the text. In particular, we show how to use data and how to manipulate units correctly.

Examples An Example is a much more structured form of Brief illustration, often involving a more elaborate procedure. Every Example has a Strategy section to suggest how you might set up the problem (you might prefer another way: setting up problems is a highly personal business). Then we provide the workedout Answer.

xvii

xviii

ABOUT THE BOOK

Self-tests Every Example has a Self-test, with the answer provided, so that you

can check whether you have understood the procedure. There are also freestanding Self-tests where we thought it a good idea to provide a question for you to check your understanding. Think of Self-tests as in-chapter Exercises designed to help you to monitor your progress.

Discussion questions The end-of-chapter material starts with a short set of

questions that are intended to encourage you to think about the material you have encountered and to view it in a broader context than is obtained by solving numerical problems.

Exercises The real core of testing your progress is the collection of end-of-

chapter Exercises. We have provided a wide variety at a range of levels.

Projects Longer and more involved exercises are presented as Projects at the end of each chapter. In many cases, the projects encourage you to make connections between concepts discussed in more than one chapter, either by performing calculations or by pointing you to the original literature.

Media and supplements W. H. Freeman has developed an extensive package of electronic resources and printed supplements to accompany the second edition of Physical Chemistry for the Life Sciences. The Book Companion Website The Book Companion Website provides teaching and learning resources to augment the printed book. It is free of charge, and contains additional material for download, much of which can be incorporated into a virtual learning environment. The Book Companion Website can be accessed by visiting www.whfreeman.com/pchemls2e/ Note that instructor resources are available only to registered adopters of the textbook. To register simply visit www.whfreeman.com/pchemls2e/ and follow the appropriate links. You will be given the opportunity to select your own username and password, which will be activated once your adoption has been verified. For Students

Living Graphs A living graph can be used to explore how a property changes as

a variety of parameters are changed. To encourage the use of this resources (and the more extensive Explorations in Physical Chemistry 2.0; below), we have included a suggested interactivity to many of the illustrations in the text, iconed in the book.

ABOUT THE BOOK

Animated Molecules A visual representation of each molecule found through-

out the text is also available on the Companion Website, courtesy of ChemSpider, the popular online search engine that aggregates chemical structures and their associated information from all over the web into a single searchable repository. You’ll also find 2D and 3D representations, as well as information on each structures’ inherent properties, identifiers, and references. For more information on ChemSpider, visit www.chemspider.com. For Instructors

Textbook Images Almost all of the figures, tables, and images from the text are

available for download in both .JPEG and PowerPoint® format. These can be use for lectures without charge, but not for commercial purposes without specific permission. Other supplements Explorations in Physical Chemistry 2.0

Valerie Walters, Julio de Paula, and Peter Atkins www.whfreeman.com/explorations ISBN: 0-7167-8586-2 Explorations in Physical Chemistry 2.0 consists of interactive Mathcad® worksheets, interactive Excel® workbooks, and stimulating exercises, designed to motivate students to simulate physical, chemical, and biochemical phenomena with their personal computers. Students can manipulate over 75 graphics, alter simulation parameters, and solve equations, to gain deeper insight into physical chemistry. It covers: • Thermodynamics, including applications to biological processes. • Quantum chemistry, including interactive three-dimensional renderings of atomic and molecular orbitals. • Atomic and molecular spectroscopy, including tutorials on Fouriertransform techniques in modern spectroscopy. • Properties of materials, including metals, polymers, and biological macromolecules. • Chemical kinetics and dynamics, including enzyme catalysis, oscillating reactions, and polymerization reactions. Explorations of Physical Chemistry 2.0 is available exclusively online. Physical Chemistry for the Life Sciences Coursesmart eBook

www.coursesmart.com An electronic version of the book is available for purchase from CourseSmart. CourseSmart eBooks are an economically alternative to printed textbooks (40% less) that are convenient, easy to use, and better for the environment. Each CourseSmart eBook reproduces the printed book exactly, page-for-page, and includes all the same text and images. CourseSmart eBooks can be purchased as either an online eBook, which is viewable from any Internet-connected computer with a standard Web browser, or as a downloadable eBook, which can be installed on any one computer and then viewed without an Internet connection. For more information, visit www.coursesmart.com

xix

xx

ABOUT THE BOOK

Solutions Manual for Physical Chemistry for the Life Sciences, Second Edition

Charles Trapp, University of Louisville, and Marshall Cady, Indiana University Southeast. ISBN: 1-4292-3125-4 The Solutions Manual contains complete solutions to the end-of-chapter exercises, discussion questions, and projects from each chapter in the textbook. These worked-out-solutions will guide you through each step and help you refi ne your problem-solving skills.

Prolog Chemistry is the science of matter and the changes it can undergo. Physical chemistry is the branch of chemistry that establishes and develops the principles of the subject in terms of the underlying concepts of physics and the language of mathematics. Its concepts are used to explain and interpret observations on the physical and chemical properties of matter. This text develops the principles of physical chemistry and their applications to the study of the life sciences, particularly biochemistry and medicine. The resulting combination of the concepts of physics, chemistry, and biology into an intricate mosaic leads to a unique and exciting understanding of the processes responsible for life.

The structure of physical chemistry Like all scientists, physical chemists build descriptions of nature on a foundation of careful and systematic inquiry. (a) The organization of science The observations that physical chemistry organizes and explains are summarized by scientific laws. A law is a summary of experience. Thus, we encounter the laws of thermodynamics, which are summaries of observations on the transformations of energy. Laws are often expressed mathematically, as in the perfect gas law (or ideal gas law; see Section F.2), pV = nRT. This law is an approximate description of the physical properties of gases (with p the pressure, V the volume, n the amount, R a universal constant, and T the temperature). We also encounter the laws of quantum mechanics, which summarize observations on the behavior of individual particles, such as molecules, atoms, and subatomic particles. The first step in accounting for a law is to propose a hypothesis, which is essentially a guess at an explanation of the law in terms of more fundamental concepts. Dalton’s atomic hypothesis, which was proposed to account for the laws of chemical composition and changes accompanying reactions, is an example. When a hypothesis has become established, perhaps as a result of the success of further experiments it has inspired or by a more elaborate formulation (often in terms of mathematics) that puts it into the context of broader aspects of science, it is promoted to the status of a theory. Among the theories we encounter are the theories of chemical equilibrium, atomic structure, and the rates of reactions. A characteristic of physical chemistry, like other branches of science, is that to develop theories, it adopts models of the system it is seeking to describe. A model is a simplified version of the system that focuses on the essentials of the problem. Once a successful model has been constructed and tested against known observations and any experiments the model inspires, it can be made more sophisticated and incorporate some of the complications that the original model ignored.

xxii

PROLOG

Thus, models provide the initial framework for discussions, and reality is progressively captured rather like a building is completed, decorated, and furnished. One example is the nuclear model of an atom, and in particular a hydrogen atom, which is used as a basis for the discussion of the structures of all atoms. In the initial model, the interactions between electrons are ignored; to elaborate the model, repulsions between the electrons are taken into account progressively more accurately. (b) The organization of our presentation The text begins with an investigation of thermodynamics, the study of the transformations of energy, and the relations between the bulk properties of matter. Thermodynamics is summarized by a number of laws that allow us to account for the natural direction of physical and chemical change. Its principal relevance to biology is its application to the study of the deployment of energy by organisms. We then turn to chemical kinetics, the study of the rates of chemical reactions. We shall establish how the rates of reactions can be determined and how experimental data give insight into the molecular processes by which chemical reactions occur. To understand the molecular mechanism of change, we also explore how molecules move, either in free flight in gases or by diffusion through liquids. Chemical kinetics is a crucial aspect of the study of organisms because the array of reactions that contribute to life form an intricate network of processes occurring at different rates under the control of enzymes. Next, we develop the principles of quantum theory and use them to describe the structures of atoms and molecules, including the macromolecules found in biological cells. Quantum theory is important to the life sciences because the structures of its complex molecules and the migration of electrons cannot be understood except in its terms. We extend these theories of structure to solids, principally because that most revealing of all structural techniques, X-ray diffraction, depends on the availability and features of crystalline samples. Finally, we explore the information about biological structure and function that can be obtained from spectroscopy, the study of interactions between molecules and electromagnetic radiation. The spectroscopic techniques available for the investigation of structure, which includes shape, size, and the distribution of electrons in ground and excited states, make use of most of the electromagnetic spectrum. We conclude with an account of perhaps the most important of all spectroscopies, nuclear magnetic resonance (NMR).

Applications of physical chemistry to biology and medicine Here we discuss some of the important problems in biology and medicine being tackled with the tools of physical chemistry. We shall see that physical chemists contribute importantly not only to fundamental questions, such as the unravelling of intricate relationships between the structure of a biological molecule and its function, but also to the application of biochemistry to new technologies. (a) Techniques for the study of biological systems Many of the techniques now employed by biochemists were first conceived by physicists and then developed by physical chemists for studies of small molecules

PROLOG

and chemical reactions before they were applied to the investigation of complex biological systems. Here we mention a few examples of physical techniques that are used routinely for the analysis of the structure and function of biological molecules. X-ray diffraction and nuclear magnetic resonance (NMR) spectroscopy are two very important tools commonly used for the determination of the threedimensional arrangement of atoms in biological assemblies. An example of the power of the X-ray diffraction technique is the recent determination of the three-dimensional structure of the ribosome, a complex of protein and ribonucleic acid with a molar mass exceeding 2 × 106 g mol−1 that is responsible for the synthesis of proteins from individual amino acids in the cell. This work led to the 2009 Nobel Prize in Chemistry, awarded to Venkatraman Ramakrishnan, Thomas Steitz, and Ada Yonath. Nuclear magnetic resonance spectroscopy has also advanced steadily through the years and now entire organisms may be studied through magnetic resonance imaging (MRI), a technique used widely in the diagnosis of disease. Throughout the text we shall describe many tools for the structural characterization of biological molecules. Advances in biotechnology are also linked strongly to the development of physical techniques. The ongoing effort to characterize the entire genetic material, or genome, of organisms as simple as bacteria and as complex as hom*o sapiens will lead to important new insights into the molecular mechanisms of disease, primarily through the discovery of previously unknown proteins encoded by the deoxyribonucleic acid (DNA) in genes. However, decoding genomic DNA will not always lead to accurate predictions of the amino acids present in biologically active proteins. Many proteins undergo chemical modification, such as cleavage into smaller proteins, after being synthesized in the ribosome. Moreover, it is known that one piece of DNA may encode more than one active protein. It follows that it is also important to describe the proteome, the full complement of functional proteins of an organism, by characterizing the proteins directly after they have been synthesized and processed in the cell. The procedures of genomics and proteomics, the analysis of the genome and proteome, of complex organisms are time-consuming because of the very large number of molecules that must be characterized. For example, the human genome contains about 20 000 to 25 000 protein-encoding genes and the number of active proteins is likely to be much larger. Success in the characterization of the genome and proteome of any organism will depend on the deployment of very rapid techniques for the determination of the order in which molecular building blocks are linked covalently in DNA and proteins. An important tool is gel electrophoresis, in which molecules are separated on a gel slab in the presence of an applied electrical field. It is believed that mass spectrometry, a technique for the accurate determination of molecular masses, will be of great significance in proteomic analysis. We discuss the principles and applications of gel electrophoresis and mass spectrometry in Chapters 8 and 11, respectively. (b) Protein folding Proteins consist of flexible chains of amino acids. However, for a protein to function correctly, it must have a well-defined conformation. Although the amino acid sequence of a protein contains the necessary information to create the active conformation of the protein from a newly synthesized chain, the prediction of the conformation from the sequence, the so-called protein folding problem,

xxiii

xxiv

PROLOG

is extraordinarily difficult and is still the focus of much research. Solving the problem of how a protein finds its functional conformation will also help us to understand why some proteins fold improperly under certain circ*mstances. Misfolded proteins are thought to be involved in a number of diseases, such as cystic fibrosis, Alzheimer’s disease, and ‘mad cow’ disease (variant Creutzfeldt– Jakob disease, v-CJD). To appreciate the complexity of the mechanism of protein folding, consider a small protein consisting of a single chain of 100 amino acids in a well-defined sequence. Statistical arguments lead to the conclusion that the polymer can exist in about 1049 distinct conformations, with the correct conformation corresponding to a minimum in the energy of interaction between different parts of the chain and the energy of interaction between the chain and surrounding solvent molecules. In the absence of a mechanism that streamlines the search for the interactions in a properly folded chain, the correct conformation can be attained only by sampling every one of the possibilities. If we allow each conformation to be sampled for 10−20 s, a duration far shorter than that observed for the completion of even the fastest of chemical reactions, it could take more than 1021 years, which is much longer than the age of the Universe, for the proper fold to be found. However, it is known that proteins can fold into functional conformations in less than 1 s. The preceding arguments form the basis for Levinthal’s paradox and lead to a view of protein folding as a complex problem in thermodynamics and chemical kinetics: how does a protein minimize the energies of all possible molecular interactions with itself and its environment in such a relatively short period of time? It is no surprise that physical chemists are important contributors to the solution of the protein-folding problem. We discuss the details of protein folding in Chapters 8 and 11. For now, it is sufficient to outline the ways in which the tools of physical chemistry can be applied to the problem. Computational techniques that employ both classical and quantum theories of matter provide important insights into molecular interactions and can lead to reasonable predictions of the functional conformation of a protein. For example, in a molecular mechanics simulation, mathematical expressions from classical physics are used to determine the structure corresponding to the minimum in the energy of molecular interactions within the chain at the absolute zero of temperature. Such calculations are usually followed by molecular dynamics simulations, in which the molecule is set in motion by heating it to a specified temperature. The possible trajectories of all atoms under the influence of intermolecular interactions are then calculated by consideration of Newton’s equations of motion. These trajectories correspond to the conformations that the molecule can sample at the temperature of the simulation. Calculations based on quantum theory are more difficult and time-consuming, but theoretical chemists are making progress toward merging classical and quantum views of protein folding. As is usually the case in physical chemistry, theoretical studies inform experimental studies and vice versa. Many of the sophisticated experimental techniques in chemical kinetics to be discussed in Chapter 6 continue to yield details of the mechanism of protein folding. For example, the available data indicate that, in a number of proteins, a significant portion of the folding process occurs in less than 1 ms (10−3 s). Among the fastest events is the formation of helical and sheetlike structures from a fully unfolded chain. Slower events include the formation of contacts between helical segments in a large protein.

PROLOG

(c) Rational drug design The search for molecules with unique biological activity represents a significant portion of the overall effort expended by pharmaceutical and academic laboratories to synthesize new drugs for the treatment of disease. One approach consists of extracting naturally occurring compounds from a large number of organisms and testing their medicinal properties. For example, the drug pacl*taxel (sold under the tradename Taxol), a compound found in the bark of the Pacific yew tree, has been found to be effective in the treatment of ovarian cancer. An alternative approach to the discovery of drugs is rational drug design, which begins with the identification of molecular characteristics of a disease-causing agent—a microbe, a virus, or a tumor—and proceeds with the synthesis and testing of new compounds to react specifically with it. Scores of scientists are involved in rational drug design, as the successful identification of a powerful drug requires the combined efforts of microbiologists, biochemists, computational chemists, synthetic chemists, pharmacologists, and physicians. Many of the targets of rational drug design are enzymes, proteins, or nucleic acids that act as biological catalysts. The ideal target is either an enzyme of the host organism that is working abnormally as a result of the disease or an enzyme unique to the disease-causing agent and foreign to the host organism. Because enzyme-catalyzed reactions are prone to inhibition by molecules that interfere with the formation of product, the usual strategy is to design drugs that are specific inhibitors of specific target enzymes. For example, an important part of the treatment of acquired immune deficiency syndrome (AIDS) involves the steady administration of a specially designed protease inhibitor. The drug inhibits an enzyme that is key to the formation of the protein envelope surrounding the genetic material of the human immunodeficiency virus (HIV). Without a properly formed envelope, HIV cannot replicate in the host organism. The concepts of physical chemistry play important roles in rational drug design. First, the techniques for structure determination described throughout the text are essential for the identification of structural features of drug candidates that will interact specifically with a chosen molecular target. Second, the principles of chemical kinetics discussed in Chapters 6 and 7 govern several key phenomena that must be optimized, such as the efficiency of enzyme inhibition and the rates of drug uptake by, distribution in, and release from the host organism. Finally, and perhaps most importantly, the computational techniques discussed in Chapters 10 and 11 are used extensively in the prediction of the structure and reactivity of drug molecules. In rational drug design, computational chemists are often asked to predict the structural features that lead to an efficient drug by considering the nature of a receptor site in the target. Then synthetic chemists make the proposed molecules, which are in turn tested by biochemists and pharmacologists for efficiency. The process is often iterative, with experimental results feeding back into additional calculations, which in turn generate new proposals for efficient drugs, and so on. Computational chemists continue to work very closely with experimental chemists to develop better theoretical tools with improved predictive power. (d) Biological energy conversion The unraveling of the mechanisms by which energy flows through biological cells has occupied the minds of biologists, chemists, and physicists for many decades. As a result, we now have a very good molecular picture of the physical

xxv

xxvi

PROLOG

and chemical events of such complex processes as oxygenic photosynthesis and carbohydrate metabolism: oxygenic photosynthesis C6H12O6(s) + 6 O2(g) 6 CO2(g) + 6 H2O(l) ffffg bcccc carbohydrate metabolism

where C6H12O6 denotes the carbohydrate glucose. In general terms, oxygenic photosynthesis uses solar energy to transfer electrons from water to carbon dioxide. In the process, high-energy molecules (carbohydrates, such as glucose) are synthesized in the cell. Animals feed on the carbohydrates derived from photosynthesis. During carbohydrate metabolism, the O2 released by photosynthesis as a waste product is used to oxidize carbohydrates to CO2. This oxidation drives biological processes, such as biosynthesis, muscle contraction, cell division, and nerve conduction. Hence, the sustenance of much of life on Earth depends on a tightly regulated carbon–oxygen cycle that is driven by solar energy. We shall encounter photosynthesis and carbohydrate metabolism throughout the text. As we shall see in Chapter 12, the harvesting of solar energy during photosynthesis occurs very rapidly and efficiently. Within about 100–200 ps (1 ps = 10−12 s) of the initial light absorption event, more than 90 per cent of the energy is trapped within the cell and is available to drive the electron transfer reactions that lead to the formation of carbohydrates and O2. Sophisticated spectroscopic techniques pioneered by physical chemists for the study of chemical reactions are being used to track the fast events that follow the absorption of solar energy. The electron transfer processes of photosynthesis and carbohydrate metabolism drive the flow of protons across the membranes of specialized cellular compartments. The chemiosmotic theory, discussed in Chapter 5, describes how the energy stored in a proton gradient across a membrane can be used to synthesize adenosine triphosphate (ATP), a mobile energy carrier. Intimate knowledge of thermodynamics and chemical kinetics is required to understand the details of the theory and the experiments that eventually verified it. The structures of nearly all the proteins associated with photosynthesis and carbohydrate metabolism have been characterized by X-ray diffraction or NMR techniques. Together, the structural data and the mechanistic models afford a nearly complete description of the relations between structure and function in biological energy conversion systems. This knowledge is now being used to design and synthesize molecular assemblies that can mimic oxygenic photosynthesis. The goal is to construct devices that trap solar energy in products of light-induced electron transfer reactions. One example is light-induced water splitting: H2O(l) fg 12 O2(g) + H2(g) light

The hydrogen gas produced in this manner can be used as a fuel in a variety of other devices. The preceding is an example of how a careful study of the physical chemistry of biological systems can yield not only surprising insights but also new technologies.

Fundamentals

We begin by reviewing material fundamental to the whole of physical chemistry and its application to biology, but which should be familiar from introductory courses. Matter and energy are the principal focus of our discussion.

F.1 Atoms, ions, and

molecules F.2 Bulk matter

1 4

F.3 Energy

10

Checklist of key concepts

17

Checklist of key equations

18

Discussion questions

18

Exercises

18

Projects

19

F.1 Atoms, ions, and molecules Atoms, ions, and molecules are the currency of discourse in the whole of chemistry and of biochemistry in particular. These concepts will be familiar from introductory chemistry and need little review here. However, it is important to keep in mind the following points. Atoms are characterized by their atomic number, Z, the number of protons in the nucleus. According to the nuclear model of an atom, a nucleus of charge Ze and containing most of the mass of the atom is surrounded by Z electrons, each of charge −e. Isotopes are atoms of the same atomic number but different mass number (or nucleon number), A, the total number of protons and neutrons in the nucleus. The loss of electrons results in cations (such as Na+ and Ca2+) and the gain of electrons results in anions (such as Cl− and O2−). When atoms are arranged in the order of increasing atomic number their properties show periodicities that are summarized by the periodic table with its familiar groups and periods (see inside the back cover). (a) Bonding and nonbonding interactions

There are three types of interaction that result in atoms bonding together into more elaborate structures. Ionic bonds arise from the electrostatic attraction between cations and anions and give rise to typically hard, brittle arrays known as ‘ionic solids’. Covalent bonds are due to the sharing of electrons and are responsible for the existence of discrete molecules, such as H2O and elaborate proteins. Metallic bonds arise when atoms are able to pool one or more of their electrons into a common sea and give rise to metals with their characteristic lustre and electrical conductivity. Covalent bonding is of the greatest importance in biology as it is responsible for the stabilities of the frameworks of organic molecules, such as DNA and proteins. However, there are interactions between regions of molecules that although much weaker than covalent bonding play a very important role in determining their shapes, and in biology molecular shape is closely allied with function. One such interaction is the hydrogen bond, A–H···B, where A and B are one of the atoms N, O, or F. Although only about 10 per cent as strong as a covalent bond, hydrogen bonding plays a major role in determining the shape of a biological macromolecule. Moreover, because it is quite weak, it permits the changes of

2

FUNDAMENTALS

shape that allow an enzyme or nucleic acid to function. Weaker still are nonbonding interactions, commonly called van der Waals interactions, which are attractions between groups of atoms in different regions of a macromolecule or between different molecules. These forces also contribute to the shapes of molecules and the interactions between them, as we shall see. The connectivity of a molecule, the pattern of covalent bonds it forms, is commonly represented by a Lewis structure, in which bonds are shown by lines, with two lines for double bonds (two shared electron pairs) and three lines for triple bonds (three shared pairs). Lone pairs, electron pairs not involved directly in bonding are also shown in Lewis structures, such as that for water (1) and acetic acid (2). Structural formulas of organic molecules are essentially Lewis structures without the explicit display of lone pairs. The rules for writing Lewis structures (such as the ‘octet rule’ relating to the number of electrons around each atom) should be familiar from introductory chemistry courses. A crucially important aspect of a double bond between two atoms, such as that in ethene (3) and on a more extensive scale in the visual pigment retinal (4), is that it confers torsional rigidity (resistance to twisting) in the region of the bond. Lewis structures of all but the simplest molecules do not show the shape of the molecule. A collection of rules known as valence-shell electron repulsion theory (VSEPR theory), in which regions of electron density (attached atoms and lone pairs) are supposed to adopt positions that minimize their repulsions, is often a helpful guide to the local shape at an atom, such as the tetrahedral arrangement of single bonds around a carbon atom. This theory should also be familiar from introductory chemistry courses. (b) Structural and functional units

Biochemistry effectively elaborates the concept of atoms by recognizing that characteristic groups of molecules can be regarded as building blocks from which the elaborate structures characteristic of organisms are constructed. These building blocks include the amino acids from which proteins are built as polypeptides, the bases that decorate the DNA double helix and constitute the genetic code, and carbohydrate molecules, such as glucose, that link together to form polysaccharides. It will already be familiar from introductory courses that proteins, which are either structural or biochemically active molecules, are polypeptides formed from different a-amino acids of general form NH2CHRCOOH (5) strung together by the peptide link, –CONH– (6). Each monomer unit in the chain is referred to as a peptide residue. About 20 amino acids occur naturally and differ in the nature of the group R. These fundamental building blocks are illustrated in the Atlas of structures, Section A, in the Resource section at the end of the text. Nucleic acids, which primarily store and transmit genetic information, are polynucleotides in which base–sugar–phosphate units are connected by phosphodiester bonds built from phosphate–ester links like that shown in (7). In DNA the sugar is b-d-2-deoxyribose (as shown in 8) and the bases are adenine (A), cytosine (C), guanine (G), and thymine (T); see the Atlas of structures, Section B. In RNA the sugar is b-d-ribose and uracil (U) replaces thymine. Polysaccharides are polymers of simple carbohydrates, such as glucose (9), linked together by C–O–C groups. They perform a variety of structural and functional roles in the cell, including energy storage and the mediation of interactions between cells (including those involved in immunological response). See the Atlas of structures, Section S.

F.1 ATOMS, IONS, AND MOLECULES

Third among the major structural units are the lipids, which are long-chain hydrocarbons, typically in the range C14–C24, with a variety of polar head groups at one end of the chain, such as –CH2CH2N(CH3)3+ and –COOH. The basic structural element of a cell membrane is a phospholipid, in which one or more hydrocarbon chains are attached to a phosphate group (see the Atlas of structures, Section L). Phospholipids form a membrane by stacking together to form a lipid bilayer, about 5 nm across (Fig. F.1), leaving the polar groups exposed to the aqueous environment on either side of the membrane. (c) Levels of structure

The concept of the ‘structure’ of a biological macromolecule takes on different meanings for the different levels at which we think about the spatial arrangement of the polypeptide chain:

3

Fig. F.1 The long hydrocarbon chains of a phospholipid can stack together to form a bilayer structure with the polar groups (represented by the spheres) exposed to the aqueous environment.

• The primary structure of a macromolecule is the sequence in which the units are linked in the polymer (Fig. F.2a). • The secondary structure of a macromolecule is the (often local) spatial arrangement of the chain. Examples of secondary structure motifs are random coils and ordered structures, such as helices and sheets, held together primarily by hydrogen bonds (Fig. F.2b). The secondary structure of DNA arises primarily from the winding of two polynucleotide chains around each other to form a double helix (Fig. F.3) held

The structural hierarchy of a biological macromolecule, in this case a protein, and a simplified representation in terms of cylinders. (a) The primary structure, the sequence of amino acid residues; (b) the local secondary structure (in this case a helix); (c) the tertiary structure: several helical segments connected by short random coils pack together; (d) the quaternary structure: several subunits with specific structures pack together. Fig. F.2

4

FUNDAMENTALS

together by hydrogen bonds involving A–T and C–G base pairs that lie parallel to each other and perpendicular to the major axis of the helix. • The tertiary structure is the overall three-dimensional structure of a macromolecule. The hypothetical protein shown in Fig. F.2c has helical regions connected by short random-coil sections. The helices interact to form a compact tertiary structure. • The quaternary structure of a macromolecule is the manner in which large molecules are formed by the aggregation of others. The DNA double helix, in which two polynucleotide chains are linked together by hydrogen bonds between adenine (A) and thymine (T), and between cytosine (C) and guanine (G). Fig. F.3

Figure F.2d shows how several molecular subunits, each with a specific tertiary structure, aggregate together. F.2 Bulk matter Atoms, ions, and molecules cohere to form bulk matter. The broadest classification of the resulting materials is as gas, liquid, or solid. The term ‘state’ has many different meanings in chemistry, and it is important to keep them all in mind. Here we review the terms ‘state of matter’ and ‘physical state’. (a) States of matter

At a ‘macroscopic’ (observational) level, we distinguish the three states of matter by noting the behavior of a substance enclosed in a rigid container: A gas is a fluid form of matter that fills the container it occupies. A liquid is a fluid form of matter that possesses a well-defined surface and (in a gravitational field) fills the lower part of the container it occupies. A solid retains its shape regardless of the shape of the container it occupies. One of the roles of physical chemistry is to establish the link between the properties of bulk matter and the behavior of the particles of which it is composed. As we work through this text, we shall gradually establish and elaborate the following models for the states of matter at a ‘microscopic’ (atomic) level:

Mathematical toolkit F.1

Quantities and units

The result of a measurement is a physical quantity that is reported as a numerical multiple of a unit: physical quantity = numerical value × unit It follows that units are treated like algebraic quantities and may be multiplied, divided, and canceled. Thus, the expression (physical quantity)/unit is the numerical value (a dimensionless quantity) of the measurement in the specified units. For instance, the mass m of an object could be reported as m = 2.5 kg or m/kg = 2.5. See Resource section 2 for a list of units. Units may be modified by a prefix that denotes a factor of a power of 10. Among the most common prefixes

are those listed in Table 3 of Resource section 2. Examples of the use of these prefixes are: 1 nm = 10−9 m 1 ps = 10−12 s 1 mmol = 10−6 mol Powers of units apply to the prefix as well as the unit they modify. For example, 1 cm3 = 1 (cm)3 and (10−2 m)3 = 10−6 m3. But note that 1 cm3 does not mean 1 c(m3). When carrying out numerical calculations, it is usually safest to write out the numerical value of an observable as powers of 10.

F.2 BULK MATTER

5

A gas is composed of widely separated particles in continuous rapid, disordered motion. A particle travels several (often many) diameters before colliding with another particle. For most of the time the particles are so far apart that they interact with each other only very weakly. A liquid consists of particles that are in contact but are able to move past one another in a restricted manner. The particles are in a continuous state of motion but travel only a fraction of a diameter before bumping into a neighbor. The overriding image is one of movement but with molecules jostling one another. A solid consists of particles that are in contact and unable to move past one another. Although the particles oscillate around an average location, they are essentially trapped in their initial positions and typically lie in ordered arrays. The main difference between the three states of matter is the freedom of the particles to move past one another. If the average separation of the particles is large, there is hardly any restriction on their motion, and the substance is a gas. If the particles interact so strongly with one another that they are locked together rigidly, then the substance is a solid. If the particles have an intermediate mobility between these extremes, then the substance is a liquid. We can understand the melting of a solid and the vaporization of a liquid in terms of the progressive increase in the liberty of the particles as a sample is heated and the particles become able to move more freely. (b) Physical state

By physical state (or just ‘state’) is meant a specific condition of a sample of matter that is described in terms of its physical form (gas, liquid, or solid) and the volume, pressure, temperature, and amount of substance present. (The precise meanings of these terms are described below.) So, 1 kg of hydrogen gas in a container of volume 10 dm3 at a specified pressure and temperature is in a particular state. The same mass of gas in a container of volume 5 dm3 is in a different state. Two samples of a given substance are in the same state if they are the same state of matter (that is, are both present as gas, liquid, or solid) and if they have the same mass, volume, pressure, and temperature. To report the physical state of a sample we need to specify a number of properties in terms of their appropriate units. The manipulation of units, which almost always will be from the International System of units (SI, from the French Système International d’Unités) described in the Resource section, is explained in Mathematical toolkit F.1. These properties and their units include the following: • Mass, m, is a measure of the quantity of matter a sample contains. Unit: 1 kg. Thus, 2 kg of lead contains twice as much matter as 1 kg of lead and indeed twice as much matter as 1 kg of anything. For typical laboratory-sized samples it is usually more convenient to use a smaller unit and to express mass in grams (g), where 1 kg = 103 g. • Volume, V, is a measure of the space a sample occupies. Unit: 1 m3. For volume we write V = 100 cm3 if the sample occupies 100 cm3 of space. Units used to express volume include cubic meters (m3), cubic decimeters (dm3), liters (L), and milliliters (mL). The liter is not an SI unit, but is exactly equal to 1 dm3. • Amount of substance, n, is a measure of the number of specified entities a sample contains. Unit: 1 mol.

A note on good practice

Physical quantities are denoted by italic, and sometimes Greek, letters (as in m for mass or r for mass density). Units are denoted by Roman letters (as in m for meter).

6

FUNDAMENTALS

A note on good practice

The unit mole should never be used without specifying the entities. Thus we speak of 1 mol H if we mean 1 mol of hydrogen atoms, and 1 mol H2 if we mean 1 mol of H2 molecules (the latter corresponds to 2 mol H).

The amount is expressed in moles (mol), where 1 mole is defined as the same number of specified entities as there are atoms in exactly 12 g of carbon-12. In practice, the amount of substance is related to the number of entities, N, by n = N/NA, where NA is Avogadro’s constant (NA = 6.022 × 1023 mol−1). Note that NA is a constant with units, not a pure number. To convert from an amount to an actual number, N, of entities we write N = nNA

Pressure units and conversion factors*

pascal, Pa

1 Pa = 1 N m−2

bar

1 bar = 105 Pa

atmosphere, atm

1 atm = 101.325 kPa = 1.013 25 bar

torr, Torr†

760 Torr = 1 atm 1 Torr = 133.32 Pa

*Values in bold are exact. †The name of the unit is torr; its symbol is Torr.

A note on good practice

The names of units derived from names of people are lowercase (as in newton and pascal), but their symbols are uppercase (as in N and Pa).

A brief comment

We shall see later (in Section F.3b) that temperature determines how molecules populate the energy levels available to them. Related to this interpretation is the fact that for molecules in a gas, the temperature determines their mean or average speed (caverage ∝ (T/M)1/2).

(F.1)

To express a known mass of matter as an amount we use the molar mass, M, of the entities: n=

Table F.1

Relation between amount and number

m M

Relation between mass and amount

(F.2)

The molar mass, M, is the mass of a sample of an element or compound divided by the amount of atoms, molecules, or formula units it contains: M=

m n

Definition of molar mass

(F.3)

The atomic weight of an element is the numerical value of the molar mass of the atoms it contains, the molecular weight of a molecular compound is the numerical value of the molar mass of its molecules, and the formula weight of an ionic compound is the molar mass of a specified formula unit of the compound. In each case ‘numerical value’ means M/(g mol−1). • Pressure, p, is the force a sample is subjected to divided by the area to which that force is applied. Unit: 1 Pa. Because force (see later) is measured in newtons (1 N = 1 kg m s−2), pressure is reported in newtons per square meter, or pascals (1 Pa = 1 N m−2). The atmosphere (atm) is commonly used as a unit of pressure, but is not an SI unit. To convert between atmospheres and pascals use 1 atm = 101.325 kPa exactly. See Table F.1. If an object is immersed in a gas, it experiences a pressure over its entire surface because molecules collide with it from all directions and exert a force during every collision. We are incessantly battered by molecules of gas in the atmosphere and experience this battering as ‘atmospheric pressure’. The pressure is greatest at sea level because the density of air, and hence the number of colliding molecules, is greatest there. The pressure of the atmosphere at sea level is about 100 kPa. When a gas is confined to a cylinder fitted with a movable piston, the position of the piston adjusts until the pressure of the gas inside the cylinder is equal to that exerted by the atmosphere. When the pressures on either side of the piston are the same, we say that the two regions on either side are in mechanical equilibrium (Fig. F.4). • Temperature, T, is the property of an object that determines in which direction energy will flow when it is in contact with another object: energy flows from higher temperature to lower temperature. Unit: 1 K. When the two bodies have the same temperature, there is no net flow of energy between them. In that case we say that the bodies are in thermal equilibrium (Fig. F.5). The symbol T is used to denote the thermodynamic temperature, which is an absolute scale with T = 0 as the lowest point. Temperatures above

F.2 BULK MATTER

A system is in mechanical equilibrium with its surroundings if it is separated from them by a movable wall and the external pressure is equal to the pressure of the gas in the system.

Fig. F.4

Fig. F.5 The temperatures of two objects act as a signpost showing the direction in which energy will flow as heat through a thermally conducting wall: (a) heat always flows from high temperature to low temperature. (b) When the two objects have the same temperature, although there is still energy transfer in both directions, there is no net flow of energy.

T = 0 are then most commonly expressed by using the Kelvin scale, in which the gradations of temperature are called kelvin (K). The Kelvin scale is defined by setting the triple point of water (the temperature at which ice, liquid water, and water vapour are in mutual equilibrium) at exactly 273.16 K. The freezing point of water (the melting point of ice) at 1 atm is then found experimentally to lie 0.01 K below the triple point, so the freezing point of water is approximately 273.15 K. The Kelvin scale is unsuitable for everyday measurements of temperature, and it is common to use the Celsius scale, which is defined in terms of the Kelvin scale as q/°C = T/K − 273.15

Relation between Kelvin and Celsius scales

(F.4)

(The 273.15 is exact in this definition.) Thus, the freezing point of water is 0°C and its boiling point (at 1 atm) is found to be 100°C. Note that in this text T invariably denotes the thermodynamic (absolute) temperature and that temperatures on the Celsius scale are denoted q (theta).

Self-test F.1

7

Use eqn F.4 to express body temperature, 37°C, in kelvins. Answer: 310 K

Temperature is an example of an intensive property, a property that is independent of the size of the sample. A property that does depend on the size (‘extent’) of the sample is called an extensive property. More formally, if we think of a sample as being divided into portions (‘subsystems’), then the value of an extensive property is the sum of the contribution from each of the subsystems. For instance, the mass of a 10 mg sample of a protein is the sum of the masses of the 10 portions, each of 1 mg, into which it can be imagined as being divided. The value of an intensive property is the same for each of the subsystems and of the overall system itself. For instance, the temperature of a uniform 100 cm3 flask of water is the same as that of each of the 10 regions, each of volume 10 cm3, into which it can be regarded as being divided. Mass, volume, and amount of substance are all

A note on good practice

We refer to absolute zero as T = 0, not T = 0 K. There are other ‘absolute’ scales of temperature, all of which set their lowest value at zero. Insofar as it is possible, all expressions in science should be independent of the units being employed, and in this case the lowest attainable temperature is T = 0 regardless of which absolute scale we are using. On the other hand, we write q = 0°C not q = 0 because the Celsius scale has an arbitrarily defined zero point.

8

FUNDAMENTALS

extensive properties. Temperature and pressure are intensive properties. Molar mass is intensive because the size-dependence of m and n cancel in the ratio m/n. All molar properties, Xm = X/n, where X is an extensive property, are intensive for the same reason. Mass density, r = m/V, is also intensive. (c) Equations of state

Although the state of any sample of substance can be specified by giving the values of its volume, the pressure, the temperature, and the amount of substance, a remarkable experimental fact is that these four quantities are not independent of one another. For instance, we cannot arbitrarily choose to have a sample of 5.5 mmol H2O in a volume of 100 cm3 at 100 kPa and 500 K: it is found experimentally that that state simply does not exist. If we select the amount, the volume, and the temperature, then we find that we have to accept a particular pressure (in this case, close to 230 kPa). The same is true of all substances, but the pressure in general will be different for each one. This experimental generalization is summarized by saying the substance obeys an equation of state, an equation of the form p = f(n,V,T)

A general equation of state

(F.5)

This expression tells us that the pressure is some function of amount, volume, and temperature, and that if we know those three variables, then the pressure can have only one value. The equations of state of most substances are not known, so in general we cannot write down an explicit expression for the pressure in terms of the other variables. However, certain equations of state are known. In particular, the equation of state of a low-pressure gas is known and proves to be very simple and very useful: p=

Table F.2 The gas constant in various units

R = 8.314 47

J K−1 mol−1

8.314 47

kPa dm3 K−1 mol−1

8.205 74 × 10−2

atm dm3 K−1 mol−1

62.364

Torr dm3 K−1 mol−1

1.987 21

cal K−1 mol−1

nRT V

Perfect gas equation of state

(F.6)

where R is the gas constant R = 8.314 J K−1 mol−1 (for values of R in other and sometimes more convenient units see Table F.2). Although the properties of gases might seem to be of little direct relevance to biochemistry, this equation is used to describe the behavior of gases taking part in a variety of biologically important processes (such as respiration), the properties of the gaseous environment we inhabit (the atmosphere), and as a starting point for the discussion of the properties of species in aqueous environments (such as the cell). The perfect gas equation of state—more briefly, the ‘perfect gas law’—is socalled because it is an idealization of the equations of state that gases actually obey. Specifically, it is found that all gases obey the equation ever more closely as the pressure is reduced toward zero. That is, eqn F.6 is an example of a limiting law, a law that becomes increasingly valid as the pressure is reduced and is obeyed exactly at the limit of zero pressure. A hypothetical substance that obeys eqn F.6 at all pressures is called a perfect gas.1 From what has just been said, an actual gas, which is termed a real gas, behaves more and more like a perfect gas as its pressure is reduced toward zero. In practice, normal atmospheric pressure at sea level (p ≈ 100 kPa) is already low enough for most real gases to behave almost perfectly and, unless stated 1

The term ‘ideal gas’ is also widely used.

F.2 BULK MATTER

otherwise, we shall always assume in this text that the gases we encounter behave like a perfect gas. The reason why a real gas behaves differently from a perfect gas can be traced to the attractions and repulsions that exist between actual molecules and that are absent in a perfect gas (Chapter 11).

A brief illustration

Consider the calculation of the pressure in kilopascals exerted by 1.25 g of nitrogen gas in a flask of volume 250 mL (0.250 dm3) at 20°C. The amount of N2 molecules (of molar mass M = 28.02 g mol −1) present is n=

m 1.25 g 1.25 = = mol M 28.02 g mol −1 28.02

The temperature of the sample is T/K = 20 + 273.15. Therefore, from p = nRT/V, n

T = 293 K

1 4 2 4 3

1 4 4 4 2 4 4 4 3

1 4 4 2 4 4 3

(1.25/28.02) mol × (8.3145 kPa dm K mol ) × (20 + 273.15 K) 0.250 dm3 3

−1

−1

A note on good practice

1 2 3

p=

R

V = 250 mL

= 435 kPa where we have used more convenient units for the constant R. Note how all units (except kPa in this instance) cancel like ordinary numbers (see Mathematical toolkit F.1). Self-test F.2 Calculate the pressure exerted by 1.22 g of carbon dioxide confined to a flask of volume 500 mL at 37°C.

Answer: 143 kPa

The molar volume, Vm, is the volume a substance (not just a gas) occupies per mole of molecules. It is calculated by dividing the volume of the sample by the amount of molecules it contains: Vm =

V n

Definition of molar volume

(F.7)

The perfect gas law can be used to calculate the molar volume of a perfect gas at any temperature and pressure. When we combine eqns F.6 and F.7, we get Vm =

V nRT RT = = n np p

Molar volume of a perfect gas

(F.8)

This expression lets us calculate the molar volume of any gas (provided it is behaving perfectly) from its pressure and its temperature. It also shows that, for a given temperature and pressure, provided they are behaving perfectly, all gases have the same molar volume. Chemists have found it convenient to report much of their data at a particular set of standard conditions, as summarized in Table F.3. The ‘standard state’ of a substance (at a specified temperature, not necessarily 298 K) is discussed further in Section 1.7. The condition SATP for the discussion of gases is now favored over

It is best to postpone the actual numerical calculation to the last possible stage and carry it out in a single step. This procedure avoids rounding errors.

9

10

FUNDAMENTALS

Table F.3

A summary of standard conditions

Name

Conditions

Comment

Standard pressure, p3

p3 = 1 bar

1 bar is exact

Standard ambient temperature and pressure (SATP)

25°C (more precisely, 298.15 K) and 1 bar

At SATP, Vm = 24.79 dm3 mol−1 for a perfect gas

Standard temperature and pressure (STP)

0°C and 1 atm

At STP, Vm = 22.41 dm3 mol−1 for a perfect gas

Standard state

Pure substance at 1 bar

Temperature to be specified. See Section 1.7.

the earlier STP on account on the shift of emphasis from 1 atm to 1 bar in the specification of standard states. A mixture of perfect gases, such as to a good approximation the atmosphere, behaves like a single perfect gas. According to Dalton’s law, the total pressure of such a mixture is the sum of the partial pressures of the constituents, the pressure to which each gas would give rise if it occupied the container alone: p = pA + pB + · · ·

Dalton’s law

(F.9)

Each partial pressure, pJ, can be calculated from the perfect gas law in the form pJ = nJRT/V. F.3 Energy A property that will continue to occur in just about every chapter of the following text is ‘energy’. Indeed, we begin the text with a discussion of the deployment of energy in living organisms. Energy, E, is the capacity to do work. Work is the process of moving against an opposing force. A fully wound spring can do more work than a half-wound spring (that is, it can raise a weight through a greater height or move a greater weight through a given height). A hot object has the potential for doing more work than the same object when it is cool and therefore has a higher energy. In his formulation of classical mechanics Isaac Newton focused on the role of force, F, an agent that changes the state of motion of a body. His mechanics was built on three laws, the second of which relates the acceleration, a, the rate of change of velocity, of a body of mass m to the strength of the force it experiences: F = ma

Newton’s second law

(F.10)

A brief illustration

A stationary ball of mass 150 g is hit by a bat, and in 0.20 s reaches a speed of 80 km h−1 (8.0 × 104 m/3600 s = 22 m s−1) before being slowed down by air resistance. The initial acceleration of the ball is (22 m s−1)/(0.20 s) = 110 m s−2. The force exerted by the bat on the ball is therefore F = (0.150 kg) × (110 m s−2) = 16.5 kg m s−2 = 16.5 N We have expressed the result in newtons, with 1 N = 1 kg m s−2.

F.3 ENERGY

Force, like acceleration, is actually a ‘vector’ quantity, a quantity with direction as well as magnitude, but in most instances in this text we need consider only its magnitude. The magnitude of the work done in moving against a constant opposing force, w, is the product of the distance moved, d, and the strength of the force: w = Fd

Definition of work

(F.11)

A brief illustration

A bird of mass 50 g flies from the ground to a branch 10 m above. The force of gravity on an object of mass m close to the surface of the Earth is mg, where g is the ‘acceleration of free fall’: g = 9.81 m s−2. Therefore, the work it has to do against gravity is w = mgd = (0.050 kg) × (9.81 m s−2) × (10 m) = 4.9 kg m2 s−2 We would report this value as 4.9 J, where J = 1 kg m2 s−2.

As implied in the brief illustration, the SI unit of energy is the joule (J), named after the nineteenth-century scientist James Joule, who helped to establish the concept of energy (see Chapter 1). It is defined as 1 J = 1 N m = 1 kg m2 s−2. A joule is quite a small unit, and in chemistry we often deal with energies of the order of kilojoules (1 kJ = 103 J). (a) Varieties of energy

We need to distinguish the energies possessed by matter and due to radiation. The kinetic energy, Ek, is the energy of a body due to its motion. For a body of mass m moving at a speed v, Ek = 12 mv 2

Definition of kinetic energy

(F.12)

That is, a heavy object moving at the same speed as a light object has a higher kinetic energy, and doubling the speed of any object increases its kinetic energy by a factor of 4. A ball of mass 1 kg traveling at 1 m s−1 has a kinetic energy of 0.5 J. The potential energy, Ep (and commonly V), of a body is the energy it possesses due to its position. The precise dependence on position depends on the type of force acting on the body. An important type of potential energy is the Coulombic potential energy of interaction between two electric charges Q1 and Q2 separated by a distance r: Ep =

Q1Q2 4pε0r

Coulombic potential energy

(F.13)

The fundamental constant ε0 is called the vacuum permittivity; its value (and those of other fundamental constants) is given inside the front cover. With the charges in coulombs (C) and the distance in meters, the energy is obtained in joules. Equation F.13 is based on the convention of taking the potential energy to be zero when the charges are infinitely apart. The Coulombic potential energy will inform our discussion of a range of topics, from atomic structure to the nature of interactions that give rise to levels of structure in biological assemblies.

11

12

FUNDAMENTALS

A mass m close to the surface of the Earth has a potential energy that is proportional to its height above the ground, h: Ep = mgh

Gravitational potential energy

(F.14)

The constant g = 9.81 m s−2 is called the acceleration of free fall. It depends on the location on the Earth’s surface, but the variation is quite small. In this case, the arbitrary zero of potential energy is taken as being at the surface of the Earth (at h = 0).

A brief illustration

The potential energy of the 50-g bird mentioned in the preceding brief illustration is higher by 4.9 J when it is on the branch than when it is on the ground. The potential energy of this book (of mass about 1 kg) is higher by about 10 J when it is on a table 1 m above the floor than when it is on the floor.

The total energy, E, of a material body is the sum of its kinetic and potential energies: E = Ek + Ep

Total energy

(F.15)

Provided no external forces are acting on the body, its total energy is constant. This remark is elevated to a central statement of classical physics known as the law of the conservation of energy. Potential and kinetic energy may be freely interchanged, for instance a falling ball loses potential energy but gains kinetic energy as it accelerates, but its total energy remains constant provided the body is isolated from external influences, such as air resistance. Energy may also be present even in the absence of matter in the form of electromagnetic radiation, a wave of electric and magnetic fields traveling through a vacuum at the ‘speed of light’, c = 2.998 × 108 m s−1. The wave is characterized by its amplitude, frequency, and wavelength. The amplitude of the wave is the maximum displacement, and the perceived intensity of the wave is proportional to the square of the amplitude. The frequency, n (nu), is a measure of the rate at which the field goes through a complete cycle of orientations. The SI unit of frequency is 1 hertz (1 Hz), which corresponds to one cycle per second: 1 Hz = 1 s−1. The wavelength, l (lambda), is the distance between neighboring peaks of the wave (Fig. F.6). The frequency and wavelength are related by ln = c

An electromagnetic wave is characterized by its amplitude, A, wavelength, l, and frequency, n; the frequency is related to the wavelength by n = c/l.

Fig. F.6

Relation between frequency and wavelength

(F.16)

That is, high frequencies correspond to short wavelengths, and vice versa. This expression also applies to sound waves, with c interpreted as the speed of sound. The electromagnetic spectrum runs—as far as we know—over all frequencies. Each range of frequencies is classified as shown in Fig. F.7. The boundaries between each region are only approximate. The visible region of the spectrum, the region to which our eyes are sensitive, occupies a very narrow band between 400 and 700 nm. As we shall see in later chapters, each region of the spectrum excites, or is excited by, different types of nuclear, atomic, or molecular transition. For instance, electronic excitations, where electrons are redistributed into

F.3 ENERGY

different regions of the molecule, are stimulated by or give rise to visible and ultraviolet radiation. Because the separation between energy levels is dictated by the arrangement of atoms in a molecule, measuring the frequencies of transitions facilitates the study of molecular structure and reactivity. Ultraviolet radiation can also cause such extreme electron redistributions that bonds are broken. We need to be aware that electromagnetic energy is delivered in packets known as photons. The energy of a photon of electromagnetic radiation is related to the frequency of the radiation by E = hn

Energy of a photon

13

The regions of the electromagnetic spectrum and some of the spectroscopic techniques that make use of them.

Fig. F.7

(F.17)

where h is a fundamental constant known as Planck’s constant (h = 6.626 × 10−34 J s). In terms of photons, an intense ray of light consists of numerous photons, each of the same energy and each moving at the speed c. The higher the frequency of the radiation, the greater is the energy carried by each photon. Photons of visible light are sufficiently energetic to stimulate the processes of vision; photons of ultraviolet radiation are so energetic that they can destroy tissue.

A brief illustration

The energy of a photon of 350 nm ultraviolet radiation is E=

hc (6.626 × 10−34 J s) × (2.998 × 108 m s−1) = 5.68 × 10−19 J = l 3.50 × 10−7 m

corresponding to 0.568 aJ. To know the energy per mole of photons, which helps us to assess the chemical potency of the radiation, we multiply by Avogadro’s constant: E=

hcNA (6.626 × 10−34 J s) × (2.998 × 108 m s−1) × (6.022 × 1023 mol−1) = l 3.50 × 10−7 m = 342 kJ mol−1

(b) The Boltzmann distribution

One of the most important expressions in science, the ‘Boltzmann distribution’, helps to elucidate the concept of temperature as well as underlying virtually all the bulk properties and reactions of matter and their variation with temperature.

A note on good practice It is

best to carry out a numerical calculation in one step or at least to avoid rounding at an intermediate stage.

14

FUNDAMENTALS

Mathematical toolkit F.2

Exponential functions

In preparation for the large number of occurrences of exponential functions throughout the text, it will be useful to know the shape of exponential functions. Here we deal with two types, e−ax and e−ax . An exponential function of the form e−ax starts off at 1 when x = 0 and decays toward zero, which it reaches as x approaches infinity (see the illustration). This function approaches zero more rapidly as a increases. The Boltzmann distribution is an example of an exponential function. The function e−ax is called a Gaussian 2

function. It also starts off at 1 when x = 0 and decays to zero as x increases, however, its decay is initially slower but then plunges down more rapidly than e−ax. Gaussian functions will appear several times through the text. The illustration also shows the behavior of the two functions for negative values of x. The exponential function e−ax rises rapidly to infinity, but the Gaussian function falls back to zero and traces out a bell-shaped curve.

2

The exponential function, e−x, and the bell-shaped Gaussian function, e−x . Note that both are equal to 1 at x = 0, but the exponential function rises to infinity as x → −∞. 2

It should be familiar from introductory courses, and will be explained in detail later in the text, that atoms and molecules can possess only discrete amounts of energy. For instance, an electron in a hydrogen atom can possess only the energies 2.17 aJ, 0.54 aJ, 0.24 aJ, . . . (where 1 aJ, 1 attojoule = 10−18 J) below that of a widely separated proton and electron, and a C–H bond in a molecule can vibrate only with the energies 0.029 aJ, 0.086 aJ, 0.144 aJ, . . . Intermediate values of the energy are simply not allowed. The precise values of the allowed energies depend on the details of molecular structure, but it is generally the case that electronic energy levels are most widely spaced, then the energies of molecular vibration, and then the energies with which molecules rotate (Fig. F.8). The energies of translational motion are so close together even on an atomic scale (for instance, of the order of 10−44 J for a CO2 molecule in a region 10 cm wide) that they may be treated as continuous. The apparently random motion that molecules undergo at T > 0 is called thermal motion. The energy associated with this motion is the energy of thermal motion, but is commonly called simply thermal energy. A useful rule of thumb is

F.3 ENERGY

that the order of magnitude of the energy that a molecule possesses as a result of its thermal motion is kT, where k = 1.381 × 10−23 J K−1 is a fundamental constant called Boltzmann’s constant. The gas constant R is simply the ‘molar’ form of Boltzmann’s constant: Relation between the gas constant and Boltzmann’s constant

R = NAk

(F.18)

Thermal motion ensures that molecules will be found spread over the energy levels available to them such that their mean energy is of order kT. The population of each energy level depends on the temperature, and a very important result is that in a system at a temperature T, the ratio of populations N2 and N1 in states with energies E1 and E2 is given by the Boltzmann distribution, one form of which is N2 −(E −E )/kT =e N1 2

The Boltzmann distribution

1

(F.19a)

This form of the distribution applies when the Ei are actual energies (in joules, for instance); when the Ei are molar quantities (in joules or kilojoules per mole, for instance), we use N2 −(E −E )/RT =e N1 2

(F.19b)

1

with R in place of k. We see that the greater the energy separation E2 − E1, the smaller the ratio of populations. Alternatively, for a given separation, the ratio becomes smaller as the temperature is lowered. In other words, as the temperature is lowered, more and more molecules are found in their lowest energy levels and fewer are found in high energy levels. The temperature, we see, is the single parameter we need in order to state the relative populations of energy levels.

A brief illustration

Suppose that two conformations of neighboring peptide groups in a polypeptide differ in energy by 7.5 kJ mol−1, with conformation A higher in energy than conformation B. At body temperature (37°C, corresponding to 310 K) the ratio of populations of the two conformations is NA −(7500 J mol =e NB

−1

)/(8.3145 J K −1 mol−1×310 K)

= 0.054

That is, conformation B is about 18 times more abundant than conformation A.

The importance of the Boltzmann distribution will become apparent as the following chapters unfold. We shall see that it accounts for the stability of matter, for very few molecules are found in highly excited states at ordinary temperatures, but it allows for the possibility of reaction, as some molecules will be found with sufficient energy to react, and the proportion that can react increases as the temperature is raised. Already we are beginning to see why chemical reactions proceed more quickly as the temperature is raised.

Fig. F.8 The energy level separations (in joules) typical of four types of motion.

15

16

FUNDAMENTALS

We can obtain insight into the molecular origins of temperature by using the simple but powerful kinetic model of gases (also called the ‘kinetic molecular theory,’ KMT, of gases), which is based on a model of a gas that we mentioned earlier, in which the molecules are in ceaseless random motion, do not interact with one another except during collisions, and are much smaller than the average distance traveled between collisions (Fig. F.9). Different speeds correspond to different energies, so the Boltzmann formula can be used to predict the proportions of molecules having a specific speed at a particular temperature. The expression giving the fraction of molecules that have a particular speed is called the Maxwell distribution (sometimes the Maxwell–Boltzmann distribution) and has the features summarized in Figs F.10 and F.11. The Maxwell distribution, which is discussed more fully in Further information 7.1, can be used to show that the mean speed, C, of the molecules depends on the temperature T and their molar mass M as C=

A 8RT D 1/2 C pM F

Mean speed according to the Maxwell distribution

(F.20)

Thus, the mean or average speed is high for light molecules at high temperatures. The distribution itself gives more information. For instance, the tail towards high speeds is longer at high temperatures than at low, which indicates that at high temperatures more molecules in a sample have speeds much higher than average.

The model used for discussing the molecular basis of the physical properties of a perfect gas. The pointlike molecules move randomly with a wide range of speeds and in random directions, both of which change when they collide with the walls or with other molecules.

Fig. F.9

Fig. F.10 The Maxwell distribution of speeds and its variation with the temperature. Note the broadening of the distribution and the shift of the mean speed (denoted by the locations of the vertical dotted lines) to higher values as the temperature is increased.

Fig. F.11 The Maxwell distribution of speeds also depends on the molar mass of the molecules. Molecules of low molar mass have a broad spread of speeds, and a significant fraction may be found traveling much faster than the mean speed. The distribution is much narrower for heavy molecules, and most of them travel with speeds close to the mean value (denoted by the locations of the vertical dotted lines).

CHECKLIST OF KEY EQUATIONS

17

Checklist of key concepts 1. In the nuclear model, an atom of atomic number Z consists of a nucleus of charge +Ze surrounded by Z electrons each of charge −e.

10. The total energy of an isolated system is conserved, but kinetic and potential energy may be interchanged. 11. Electromagnetic radiation is characterized by its amplitude, frequency, and wavelength.

2. Proteins, nucleic acids, and polysaccharides are long molecular chains with different levels of threedimensional structure.

12. Electromagnetic radiation consists of photons, packets of energy of magnitude hn and traveling at the speed of light.

3. Cell membranes are formed by the stacking of lipid molecules into a bilayer structure.

13. The Boltzmann distribution gives the relative numbers of molecules in the energy levels available to them.

4. The states of matter are gas, liquid, and solid. 5. An equation of state is an equation relating pressure, volume, temperature, and amount of a substance.

14. The mean speed of molecules is proportional to the square root of the (absolute) temperature and inversely proportional to the square root of the molar mass.

6. The perfect gas equation of state is a limiting law applicable as p → 0. 7. Energy is the capacity to do work.

15. The properties of the Maxwell distribution of speeds are summarized in Figs F.10 and F.11.

8. Work is done when a body is moved against an opposing force. 9. The contributions to the energy of matter are the kinetic energy (the energy due to motion) and the potential energy (the energy due to position).

Checklist of key equations Property or process

Equation

Comment

Relation between number and amount

N = nNA

NA is Avogadro’s constant

Molar quantity

Xm = X/n

Molar mass is denoted M

Temperature conversion

q/°C = T/K − 273.15

273.15 is exact

Equation of state

p = nRT/V

Perfect (ideal) gas

Molar volume

Vm = RT/p

Perfect (ideal) gas Perfect (ideal) gas

Dalton’s law

p = pA + pB + · · ·

Newton’s second law

F = ma

Work

w = Fd

Kinetic energy

Ek = mv

Coulomb potential energy

Ep = Q1Q2/4pε0r

Charges in a vacuum

Gravitational potential energy

Ep = mgh

Close to surface of the Earth

Relation between wavelength and frequency

ln = c

c is speed of propagation (e.g. speed of light)

1 2

F is the opposing force 2

Energy of a photon

E = hn

Boltzmann distribution

N2 /N1 = e−(E −E )/kT

k is Boltzmann’s constant, R = NAk

Mean speed of molecules

C = (8RT/pM)1/2

Perfect (ideal) gas

h is Planck’s constant 2

1

18

FUNDAMENTALS

Discussion questions F.1 Distinguish between ionic bonds, covalent bonds, hydrogen

bonds, and van der Waals interactions.

F.6 Define the terms force, work, energy, kinetic energy, potential energy, and the energy of thermal motion.

F.2 Distinguish between polypeptides, polynucleotides, and

F.7 Distinguish between mechanical and thermal equilibrium.

polysaccharides.

F.8 Describe the main features of electromagnetic radiation and the

F.3 Describe the main structural features of a lipid bilayer.

electromagnetic spectrum.

F.4 Distinguish between primary, secondary, tertiary, and quaternary

F.9 Use the Boltzmann distribution to provide a molecular

levels of structure in biological macromolecules.

interpretation of temperature.

F.5 Explain the differences between gases, liquids, and solids at

macroscopic and microscopic levels.

Exercises Treat all gases as perfect unless instructed otherwise. F.10 You will see Lewis structures throughout the text. Using your knowledge of introductory chemistry, draw the Lewis structures of (a) SO32−, (b) XeF4, (c) P4, (d) O3, (e) ClF 3+, and (f) N3−. F.11 Using your knowledge of VSEPR theory from introductory chemistry, predict the shapes of (a) PCl3, (b) PCl5, (c) XeF2, (d) XeF4, (e) H2O2, (f) FSO3−, (g) KrF2, and (h) PCl 4+. F.12 Express (a) 110 kPa in torr, (b) 0.997 bar in atmospheres, (c) 2.15 × 104 Pa in atmospheres, and (d) 723 Torr in pascals. F.13 Given that the Celsius and Fahrenheit temperature scales are related by qCelsius/°C = 59 (qFahrenheit/°F − 32), what is the temperature of absolute zero (T = 0) on the Fahrenheit scale?

at 101 kPa is compressed at constant temperature from 7.20 dm3 to 4.21 dm3. Calculate the final pressure of the gas. F.19 Hot-air balloons gain their lift from the lowering of density

of air that occurs when the air in the envelope is heated. To what temperature should you heat a sample of air, initially at 340 K, to increase its volume by 14 per cent? F.20 At sea level, where the pressure was 104 kPa and the temperature 21.1°C, a certain mass of air occupied 2.0 m3. To what volume will the region expand when it has risen to an altitude where the pressure and temperature are (a) 52 kPa, −5.0°C and (b) 880 Pa, −52.0°C? F.21 A diving bell has an air space of 3.0 m3 when on the deck of a boat. What is the volume of the air space when the bell has been lowered to a depth of 50 m? Take the mean density of seawater to be 1.025 g cm−3 and assume that the temperature is the same as on the surface.

F.14 Imagine that Pluto is inhabited and that its scientists use a temperature scale in which the freezing point of liquid nitrogen is 0°P (degrees Plutonium) and its boiling point is 100°P. The inhabitants of Earth report these temperatures as −209.9°C and −195.8°C, respectively. What is the relation between temperatures on (a) the Plutonium and Kelvin scales, and (b) the Plutonium and Fahrenheit scales?

F.22 Calculate the work that a person of mass 65 kg must do to climb between two floors of a building separated by 3.5 m.

F.15 Much to everyone’s surprise, nitrogen monoxide (nitric oxide,

F.24 A car of mass 1.5 t (1 t = 103 kg) traveling at 50 km h−1 must be brought to a stop. How much kinetic energy must be dissipated?

NO) has been found to act as a neurotransmitter. To prepare to study its effect, a sample was collected in a container of volume 250.0 cm3. At 19.5°C its pressure is found to be 24.5 kPa. What amount (in moles) of NO has been collected? F.16 The effect of high pressure on organisms, including humans, is studied to gain information about deep-sea diving and anesthesia. A sample of air occupies 1.00 dm3 at 25°C and 1.00 atm. What pressure is needed to compress it to 100 cm3 at this temperature? F.17 You are warned not to dispose of pressurized cans by throwing them onto a fire. The gas in an aerosol container exerts a pressure of 125 kPa at 18°C. The container is thrown on a fire, and its temperature rises to 700°C. What is the pressure at this temperature? F.18 Until we find an economical way of extracting oxygen from seawater or lunar rocks, we have to carry it with us to inhospitable places and do so in compressed form in tanks. A sample of oxygen

F.23 What is the kinetic energy of a tennis ball of mass 58 g served at 30 m s−1?

F.25 Consider a region of the atmosphere of volume 25 dm3, which at 20°C contains about 1.0 mol of molecules. Take the average molar mass of the molecules as 29 g mol−1 and their average speed as about 400 m s−1. Estimate the energy stored as molecular kinetic energy in this volume of air. F.26 Calculate the minimum energy that a bird of mass 25 g must expend in order to reach a height of 50 m. F.27 The potential energy of a charge Q1 in the presence of another

charge Q2 can be expressed in terms of the Coulomb potential, f (phi): V = Q 1f

f=

Q2 4pe0r

The units of potential are joules per coulomb, J C −1, so when f is multiplied by a charge in coulombs, the result is in joules. The

PROJECT

combination joules per coulomb occurs widely and is called a volt (V), with 1 V = 1 J C −1. Calculate the Coulomb potential due to the nuclei at a point in a LiH molecule located 200 pm from the Li nucleus and 150 pm from the H nucleus. Hint: Use Q = +Ze, where Z is the atomic number and e is the elementary charge. F.28 Plot the Coulomb potential (see Exercise F.27) due to the nuclei

at a point in a Na+Cl− ion pair located on a line half-way between the nuclei (the internuclear separation is 283 pm) as the point approaches from infinity and ends at the mid point between the nuclei. F.29 What is the wavelength of the radiation used by an FM radio transmitter broadcasting at 92.0 MHz? F.30 What is the energy of (a) a single photon and (b) 1.00 mol of

photons of wavelength 670 nm?

19

F.31 Suppose that a macromolecule can exist either as a random coil or fully stretched out, with the latter conformation 2.4 kJ mol−1 higher in energy. What is the ratio of the two conformations at 20°C? F.32 An electron spin can adopt either of two orientations in a magnetic field, and its energies are ±mBB, where mB = 9.274 × 10−24 J T −1 is the Bohr magneton and B is the intensity of the magnetic field, often reported in teslas (1 T = 1 kg s−2 A−1). Calculate the relative populations of the spin states at (a) 4.0 K and (b) 298 K, when B = 1.0 T. F.33 The composition of planetary atmospheres is determined in part by the speeds of the molecules of the constituent gases because the faster-moving molecules can reach escape velocity and leave the planet. Calculate the mean speed of (a) He atoms and (b) CH4 molecules at (i) 77 K, (ii) 298 K, and (iii) 1000 K.

Project F.34 You will now explore the gravitational potential energy in some detail, with an eye toward discovering the origin of the value of the constant g, the acceleration of free fall, and the magnitude of the gravitational force experienced by all organisms on the Earth.

(a) The gravitational potential energy of a body of mass m at a distance r from the center of the Earth is −GmmE/r, where mE is the mass of the Earth and G is the gravitational constant (see inside front cover). Consider the difference in potential energy of the body when it is moved from the surface of the Earth (radius rE ) to a height h above the surface, with h 0) taking place at constant pressure results in an increase in enthalpy (DH > 0) because energy enters the system as heat. On the other hand, an exothermic process (q < 0) taking place at constant pressure corresponds to a decrease in enthalpy (DH < 0) because energy leaves the system as heat. In summary: exothermic process DH < 0

endothermic process DH > 0

Because all combustion reactions, including the controlled ‘combustions’ that contribute to respiration, are exothermic, they are accompanied by a decrease in enthalpy. These relations are consistent with the name enthalpy, which is derived from the Greek words meaning ‘heat inside’: the ‘heat inside’ the system is increased if the process is endothermic and absorbs energy as heat from the surroundings; it is decreased if the process is exothermic and releases energy as heat into the surroundings.5 5 But heat does not actually ‘exist’ inside: only energy exists in a system; heat is a means of recovering that energy or increasing it. Heat is energy in transit, not a form in which energy is stored.

1.6 THE ENTHALPY

41

(c) The temperature dependence of the enthalpy

We have seen that the internal energy of a system rises as the temperature is increased. The same is true of the enthalpy, which also rises when the temperature is increased (Fig. 1.15). For example, the enthalpy of 100 g of water is greater at 80°C than at 20°C. We can measure the change by monitoring the energy that we must supply as heat to raise the temperature through 60°C when the sample is open to the atmosphere (or subjected to some other constant pressure); it is found that DH ≈ +25 kJ in this instance. Just as we saw that the constant-volume heat capacity tells us about the temperature-dependence of the internal energy at constant volume, so the constant-pressure heat capacity tells us how the enthalpy of a system changes as its temperature is raised at constant pressure. To derive the relation, we combine the definition of heat capacity in eqn 1.5 (C = q/DT) with eqn 1.13 and obtain Cp =

DH DT

Definition of constantpressure heat capacity

(1.14a)

That is, the constant-pressure heat capacity is the slope of a plot of enthalpy against temperature of a system kept at constant pressure. Because the plot might not be a straight line, in general we interpret Cp as the slope of the tangent to the curve at the temperature of interest (Fig. 1.16, Table 1.1). That is, the constant-pressure heat capacity is the derivative of the function H with respect to the variable T at a specified pressure or Cp =

A brief comment

dH dT

(1.14b)

A brief illustration

Provided the heat capacity is constant over the range of temperatures of interest, we can write eqn 1.14a as DH = Cp DT. This relation means that when the temperature of 100 g of water (5.55 mol H2O) is raised from 20°C to 80°C (so DT = +60 K) at constant pressure, the enthalpy of the sample changes by DH = Cp DT = nCp,mDT = (5.55 mol) × (75.29 J K−1 mol−1) × (60 K) = +25 kJ The greater the temperature rise, the greater the change in enthalpy and therefore the greater the heating that is required to bring it about. Note that this calculation is only approximate because the heat capacity depends on the temperature and we have used an average value for the temperature range of interest.

The difference between Cp,m and CV,m is significant for gases (for oxygen, CV,m = 20.8 J K−1 mol−1 and Cp,m = 29.1 J K−1 mol−1), which undergo large changes of volume when heated, but is negligible for most solids and liquids. For a perfect gas, you are invited to show in Exercise 1.19 that Cp,m − CV,m = R

Fig. 1.15 The enthalpy of a system increases as its temperature is raised. Note that the enthalpy is always greater than the internal energy of the system and that the difference increases with temperature.

Difference between the molar heat capacities of a perfect gas

(1.15)

The molar heat capacity of a substance at constant pressure is always greater than the molar heat capacity at constant volume. The reason is that when a system is

Again more precisely the constant-pressure heat capacity is the partial deriVatiVe of the function H with respect to the variable T, denoted as Cp =

A ∂H D C ∂T F p

and calculated by holding the variable p constant.

42

1 THE FIRST LAW

free to expand, some of the energy supplied as heat is free to escape back into the surroundings as work. Therefore, the rise in temperature at constant pressure is not as great as at constant volume (when no expansion work can be done), and the heat capacity is correspondingly greater.

In the laboratory 1.1

Calorimetry

Calorimetry is the study of heat transfer during physical and chemical processes. A calorimeter 6 is a device for measuring energy transferred as heat. Here we explore three common types of calorimeters used in investigations of nutrients, fuels, and biological processes. (a) Bomb calorimeters The heat capacity at constant pressure is the slope of the curve showing how the enthalpy varies with temperature; the heat capacity at constant volume is the corresponding slope of the internal energy curve. Note that the heat capacity varies with temperature (in general) and that Cp is greater than CV. Fig. 1.16

The most common device for measuring DU is an adiabatic bomb calorimeter (Fig. 1.17). The process under study is initiated inside a constant-volume container, the ‘bomb’. The bomb is immersed in a stirred water bath, and the whole device is the calorimeter. The calorimeter is also immersed in an outer water bath. The water in the calorimeter and of the outer bath are both monitored and adjusted to the same temperature. This arrangement ensures that there is no net loss of heat from the calorimeter to the surroundings (the bath) and hence that the calorimeter is adiabatic. The change in temperature, DT, of the calorimeter is proportional to the energy that the process releases or absorbs as heat. Therefore, by measuring DT we can determine qV and hence find DU. The conversion of DT to qV is best achieved by calibrating the calorimeter using a process of known energy output and determining the calorimeter constant, the constant C in the relation q = CDT

Calorimeter constant

(1.16)

The calorimeter constant may be measured electrically by passing a constant current, I, from a source of known potential difference, V, through a heater for a known period of time, t, for then q = IVt

(1.17)

A brief comment

Electrical charge is measured in coulombs, C. The motion of charge gives rise to an electric current, I, measured in coulombs per second, or amperes, A, where 1 A = 1 C s−1. If a constant current I flows through a potential difference V (measured in volts, V), the total energy supplied in an interval t is IV t. Because 1 A V s = 1 (C s−1) V s = 1 C V = 1 J, the energy is obtained in joules with the current in amperes, the potential difference in volts, and the time in seconds.

A brief illustration

If we pass a current of 10.0 A from a 12 V supply for 300 s, then from eqn 1.17 the energy supplied as heat is q = (10.0 A) × (12 V) × (300 s) = 3.6 × 104 A V s = 36 kJ because 1 A V s = 1 J. If the observed rise in temperature is 5.5 K, then the calorimeter constant is C = (36 kJ)/(5.5 K) = 6.5 kJ K−1. Alternatively, C may be determined by using a reaction of known heat output, such as the combustion of benzoic acid (C6H5COOH), for which the heat output is 3227 kJ per mole of C6H5COOH consumed. With C known, it is simple to interpret an observed temperature rise as a release of energy as heat.

6

The word calorimeter comes from ‘calor’, the Latin word for heat.

1.6 THE ENTHALPY

43

Bomb calorimetry is used in nutritional studies to determine the total energy content of a nutrient, also called its gross energy (G.E.) content. The results may be expressed in a number of units, but common in nutritional studies is the large calorie or nutritional calorie (abbreviation: Cal), which is defined as 1 Cal = 4.184 kJ exactly. The large calorie is the unit of energy used colloquially and on labels on packages of food products. This unit is distinct from the calorie (abbreviation: cal), or ‘small calorie’, still encountered in the scientific literature, 1 cal = 4.184 J exactly. It follows that 1 Cal = 1 kcal.

Example 1.2

Calibrating a calorimeter and measuring the energy content of a nutrient

In an experiment to measure the heat released by the combustion of a sample of nutrient, the compound was burned in a calorimeter and the temperature rose by 3.22°C. When a current of 1.23 A from a 12.0 V source flowed through a heater in the same calorimeter for 156 s, the temperature rose by 4.47°C. What is the energy content of the nutrient, taken as the heat released by the combustion reaction? Strategy We calculate the heat supplied electrically by using eqn 1.17 and 1 A V s = 1 J. Then we use the observed rise in temperature to find the heat capacity of the calorimeter. Finally, we use this heat capacity to convert the temperature rise observed for the combustion into a heat output by writing q = CDT (or q = CDq if the temperature is given on the Celsius scale). Solution The heat supplied during the calibration step is

q = IV t = (1.23 A) × (12.0 V) × (156 s)

A constant-volume adiabatic bomb calorimeter. The ‘bomb’ is the central vessel, which is strong enough to withstand high pressures. The calorimeter (for which the heat capacity must be known) is the entire assembly shown here. To ensure adiabaticity, the calorimeter is immersed in a water bath with a temperature continuously adjusted to that of the calorimeter at each stage of the combustion.

Fig. 1.17

= 1.23 × 12.0 × 156 A V s = 1.23 × 12.0 × 156 J This product works out as 2.30 kJ, but to avoid rounding errors we save the numerical work to the final stage. The heat capacity of the calorimeter is C=

q 1.23 × 12.0 × 156 J 1.23 × 12.0 × 156 = = J °C −1 Dq 4.47°C 4.47

A note on good practice

The numerical value of C is 515 J °C −1, but we don’t evaluate it yet in the actual calculation. The heat output of the combustion is therefore q = CDq =

A 1.23 × 12.0 × 156 D J °C −1 × 3.22°C = 1.66 kJ C F 4.47

= 0.397 Cal Self-test 1.3 In an experiment to measure the heat released by the combustion of a sample of fuel, the compound was burned in an oxygen atmosphere inside a calorimeter and the temperature rose by 2.78°C. When a current of 1.12 A from an 11.5 V source flowed through a heater in the same calorimeter for 162 s, the temperature rose by 5.11°C. What is the heat released by the combustion reaction?

Answer: 1.1 kJ

As well as keeping the numerical evaluation to the final stage (or at least not rounding intermediate values obtained with a calculator), show the units at each stage of the calculation.

44

1 THE FIRST LAW

(b) Isobaric calorimeters

An enthalpy change can be measured calorimetrically by monitoring the temperature change that accompanies a physical or chemical change occurring at constant pressure. A calorimeter for studying processes at constant pressure is called an isobaric calorimeter. A simple example is a thermally insulated vessel open to the atmosphere: the heat released in the reaction is monitored by measuring the change in temperature of the contents. For a combustion reaction an adiabatic flame calorimeter may be used to measure DT when a given amount of substance burns in a supply of oxygen (Fig. 1.18). The relative efficiencies of fuels may be evaluated by this method (see Section 1.9). (c) Differential scanning calorimeters

An adiabatic flame calorimeter, an example of an isobaric calorimeter, consists of this component immersed in a stirred water bath. Combustion occurs as a known amount of reactant is passed through to fuel the flame and the rise of temperature is monitored.

Fig. 1.18

A brief comment

The rate of change of energy is the power, expressed as joules per second, or watts, W: 1 W = 1 J s−1. Because 1 J = 1 A V s, in terms of electrical units 1 W = 1 A V. We write the electrical power, P, as P = (energy supplied)/t = IV t/t = IV.

A differential scanning calorimeter (DSC) is more sophisticated than the calorimeters discussed so far. The term ‘differential’ refers to the fact that the behavior of the sample is compared to that of a reference material that does not undergo a physical or chemical change during the analysis. The term ‘scanning’ refers to the fact that the temperatures of the sample and reference material are increased, or scanned, systematically during the analysis. A DSC consists of two small compartments that are heated electrically at a constant rate (Fig. 1.19). The temperature, T, at time t during a linear scan is T = T0 + at, where T0 is the initial temperature and a is the temperature scan rate (in kelvins per second, K s−1). A computer controls the electrical power output in order to maintain the same temperature in the sample and reference compartments throughout the analysis. The temperature of the sample changes significantly relative to that of the reference material if a chemical or physical process that involves heating occurs in the sample during the scan. To maintain the same temperature in both compartments, excess energy is transferred as heat to the sample during the process. For example, an endothermic process lowers the temperature of the sample relative to that of the reference and, as a result, the sample must be supplied with more energy (as heat) than the reference in order to maintain equal temperatures. If no physical or chemical change occurs in the sample at temperature T, we can use eqn 1.5 to write qp = Cp DT, where DT = T − T0 = at and we have assumed that Cp is independent of temperature. If an endothermic process occurs in the sample, we have to supply additional ‘excess’ energy by heating, qp,ex, to achieve the same change in temperature of the sample and can express this excess energy in terms of an additional contribution to the heat capacity, Cp,ex, by writing qp,ex = Cp,exDT. It follows that Cp,ex =

A differential scanning calorimeter. The sample and a reference material are heated in separate but identical compartments. The output is the difference in power needed to maintain the compartments at equal temperatures as the temperature rises.

Fig. 1.19

qp,ex qp,ex Pex = = DT at a

where Pex = qp,ex /t is the excess electrical power necessary to equalize the temperature of the sample and reference compartments. A DSC trace, which is called a thermogram, consists of a plot of Pex or Cp,ex against T (Fig. 1.20). Broad peaks in the thermogram indicate processes requiring the transfer of energy by heating. We show in the following Justification that the enthalpy change of the process is

T2

DH =

Cp,ex dT T1

(1.18)

1.6 THE ENTHALPY

45

That is, the enthalpy change is the area under the curve of Cp,ex against T between the temperatures at which the process begins and ends. Justification 1.3 The enthalpy change of a process from DSC data

To calculate an enthalpy change from a thermogram, we begin by rewriting eqn 1.14b as dH = Cp,exdT We proceed by integrating both sides of this expression from an initial temperature T1 and initial enthalpy H1 to a final temperature T2 and enthalpy H2.

H2

T2

dH = H1

Cp,exdT

A brief comment

Infinitesimally small quantities may be treated like any other quantity in algebraic manipulations, so the expression dy/dx = a may be rewritten as dy = adx, dx/dy = a−1, and so on.

T1

Now we use the integral 2dx = x + constant to write

H2

dH = H2 − H1 = DH H1

It follows that DH =

T2

Cp,exdT T1

which is eqn 1.18. To appreciate the utility of a DSC in biochemical investigations, we consider an important type of transformation that occurs in biological macromolecules, such as proteins and nucleic acids, and aggregates, such as biological membranes. Such large systems adopt complex three-dimensional structures as a result of intra- and inter-molecular interactions (Fundamentals F.1 and Chapter 11). Denaturation, the disruption of these interactions, can be achieved by adding chemical agents (such as urea, acids, or bases) or by changing the temperature, in which case the process is called thermal denaturation. Cooking is an example of thermal denaturation. For example, when eggs are cooked, the protein albumin is denatured irreversibly. Differential scanning calorimetry is a powerful technique for the study of denaturation of biological macromolecules. Every biopolymer has a characteristic temperature, the melting temperature, Tm, at which the three-dimensional structure unravels and biological function is lost. For example, the thermogram shown in Fig. 1.20 indicates that the widely distributed protein ubiquitin retains its native structure up to about 45°C and ‘melts’ into a denatured state at higher temperatures. The area under the curve represents the heat absorbed in this process and can be identified with the enthalpy change. The thermogram also reveals the formation of new intermolecular interactions in the denatured form. The increase in heat capacity accompanying the native → denatured transition reflects the change from a more compact native conformation to one in which the more exposed amino acid side chains in the denatured form have more extensive interactions with the surrounding water molecules. Differential scanning calorimetry is a convenient method for such studies because it requires small samples, with masses as low as 0.5 mg.

A thermogram for the protein ubiquitin. The protein retains its native structure (shown as a green ribbon diagram) up to about 45°C and then undergoes an endothermic conformational change. (Adapted from B. Chowdhry and S. LeHarne, J. Chem. Educ. 74, 236 (1997).)

Fig. 1.20

46

1 THE FIRST LAW

Physical and chemical change We shall focus on the use of the enthalpy as a useful book-keeping property for tracing the flow of energy as heat during physical processes and chemical reactions at constant pressure. The discussion will lead naturally to a quantitative treatment of the factors that optimize the suitability of fuels, including ‘biological fuels’, the foods we ingest to meet the energy requirements of daily life. 1.7 Enthalpy changes accompanying physical processes To begin to understand the complex structural changes that biological macromolecules undergo when heated or cooled, we need to understand how simpler physical changes occur.

To describe changes quantitatively, we need to keep track of the numerical value of a thermodynamic property with varying conditions, such as the states of the substances involved, the pressure, and the temperature. To simplify the calculations, chemists have found it convenient to report their data for a set of standard conditions at the temperature of their choice: The standard state of a substance is the pure substance at exactly 1 bar.7

Definition of standard state

We denote the standard state value of a property by the superscript 3 on the symbol for the property, as in H m3 for the standard molar enthalpy of a substance and p3 for the standard pressure of 1 bar. For example, the standard state of hydrogen gas is the pure gas at 1 bar and the standard state of solid calcium carbonate is the pure solid at 1 bar, with either the calcite or aragonite form specified. The physical state needs to be specified because we can speak of the standard states of the solid, liquid, and vapor forms of water, for instance, which are the pure solid, the pure liquid, and the pure vapor, respectively, at 1 bar in each case. The standard states of solutions, which are never ‘pure’, need to be treated differently (Section 3.8). In older texts you might come across a standard state defined for 1 atm (101.325 kPa) in place of 1 bar. That is the old convention. In most cases, data for 1 atm differ only a little from data for 1 bar. You might also come across standard states defined as referring to 298.15 K. That is incorrect: temperature is not a part of the definition of standard state, and standard states may refer to any temperature (but it should be specified). Thus, it is possible to speak of the standard state of water vapor at 100 K, 273.15 K, or any other temperature. It is conventional, however, for data to be reported at the so-called conventional temperature of 298.15 K (25.00°C), and from now on, unless specified otherwise, all data will be for that temperature. For simplicity, we shall often refer to 298.15 K as ‘25°C’. Finally, a standard state need not be a stable state and need not be realizable in practice. Thus, the standard state of water vapor at 25°C is the vapor at 1 bar, but water vapor at that temperature and pressure would immediately condense to liquid water. (a) Phase transitions

A phase is a specific state of matter that is uniform throughout in composition and physical state. The liquid and vapor states of water are two of its phases. The term ‘phase’ is more specific than ‘state of matter’ because a substance may exist in more than one solid form, each one of which is a solid phase. There are at least 7

Remember that 1 bar = 105 Pa exactly.

1.7 ENTHALPY CHANGES ACCOMPANYING PHYSICAL PROCESSES

47

12 forms of ice. No substance has more than one gaseous phase, so ‘gas phase’ and ‘gaseous state’ are effectively synonyms. The only substance that exists in more than one liquid phase is helium, although some evidence suggests that water might also have two liquid phases. The conversion of one phase of a substance to another phase is called a phase transition. Thus, vaporization (liquid → gas) is a phase transition, as is a transition between solid phases (such as aragonite → calcite in geological processes). With a few exceptions, phase transitions are accompanied by a change of enthalpy, for the rearrangement of atoms or molecules usually requires or releases energy. (b) Enthalpies of vaporization, fusion, and sublimation

The vaporization of a liquid, such as the conversion of liquid water to water vapor when a pool of water evaporates at 20°C or a kettle boils at 100°C, is an endothermic process (DH > 0) because heating is required to bring about the change. At a molecular level, molecules are being driven apart from the grip they exert on one another, and this process requires energy. One of the body’s strategies for maintaining its temperature at about 37°C is to use the endothermic character of the vaporization of water because the evaporation8 of perspiration requires energy and withdraws it from the skin. The energy that must be supplied as heat at constant pressure per mole of molecules that are vaporized under standard conditions (that is, pure liquid at 1 bar changing to pure vapor at 1 bar) is called the standard enthalpy of vaporization of the liquid and is denoted Dvap H 3 (Table 1.2). For example, 44 kJ of heat is required to vaporize 1 mol H2O(l) at 1 bar and 25°C, so Dvap H 3 = +44 kJ mol−1. Alternatively, we can report the same information by writing the thermochemical equation9 H2O(l) → H2O(g) Table 1.2

DH 3 = +44 kJ

Standard enthalpies of transition at the transition temperature* Boiling point, Tb/K

D vap H 9/ (kJ mol−1)

5.65

239.7

23.4

1.2

87.3

6.5

278.7

9.87

353.3

30.8

158.7

4.60

351.5

43.5

3.5

0.02

Hydrogen peroxide, H2O2

272.7

12.50

Mercury, Hg

234.3 90.7

Methanol, CH3OH

175.5

Propanone, CH3COCH3 Water, H2O

Substance

Freezing point, Tfus/K

Ammonia, NH3

195.3 83.8

Benzene, C6H6 Ethanol, C2H5OH

Argon, Ar

Helium, He

Methane, CH4

D fus H 9/ (kJ mol −1)

4.22

0.08

423.4

51.6

2.292

629.7

59.30

0.94

111.7

8.2

3.16

337.2

35.3

177.8

5.72

329.4

29.1

273.15

6.01

373.2

40.7 44.02 at 25°C 45.07 at 0°C

*For values at 298.15 K, use the information in the Resource section.

8 Evaporation is virtually synonymous with vaporization but commonly denotes vaporization to dryness. 9 Unless otherwise stated, all data in this text are for 298.15 K.

A note on good practice

The attachment of the subscript vap to the D is the modern convention; however, the older convention in which the subscript is attached to the H, as in DHvap, is still widely used. All enthalpies of vaporization are positive, so the sign is not normally written explicitly in tables of data.

48

1 THE FIRST LAW

A thermochemical equation shows the standard enthalpy change (including the sign) that accompanies the conversion of an amount of reactant equal to its stoichiometric coefficient in the accompanying chemical equation (in this case, 1 mol H2O). If the stoichiometric coefficients in the chemical equation are multiplied through by 2, then the thermochemical equation would be written 2 H2O(l) → 2 H2O(g)

DH 3 = +88 kJ

This equation signifies that 88 kJ of heat is required to vaporize 2 mol H2O(l) at 1 bar and (recalling our convention) at 298.15 K. There are some striking differences in standard enthalpies of vaporization: although the value for water is 44 kJ mol−1, that for methane, CH4, at its boiling point is only 8 kJ mol−1. Even allowing for the fact that vaporization is taking place at different temperatures, the difference between the enthalpies of vaporization signifies that water molecules are held together in the bulk liquid much more tightly than methane molecules are in liquid methane. As should be recalled from introductory courses, the interaction responsible for the low volatility of water is the hydrogen bonding between neighboring H2O molecules. The high enthalpy of vaporization of water has profound ecological consequences, for it is partly responsible for the survival of the oceans and the generally low humidity of the atmosphere. If only a small amount of heat had to be supplied to vaporize the oceans, the atmosphere would be much more heavily saturated with water vapor than is in fact the case. Another common phase transition is fusion, or melting, as when ice melts to water. The change in molar enthalpy that accompanies fusion under standard conditions (pure solid at 1 bar changing to pure liquid at 1 bar) is called the standard enthalpy of fusion, DfusH 3. Its value for water at 0°C is +6.01 kJ mol−1. As for enthalpies of vaporization, all enthalpies of fusion are positive, and the sign is not written explicitly in tables. Notice that the enthalpy of fusion of water is much less than its enthalpy of vaporization. In vaporization the molecules become completely separated from each other, whereas in melting the molecules are merely loosened without separating completely (Fig. 1.21). The reverse of vaporization is condensation and the reverse of fusion (melting) is freezing. The molar enthalpy changes are, respectively, the negative of the enthalpies of vaporization and fusion because the energy that is supplied (during heating) to vaporize or melt the substance is released when it condenses or

When a solid (a) melts to a liquid (b), the molecules separate from one another only slightly, the intermolecular interactions are reduced only slightly, and there is only a small change in enthalpy. When a liquid vaporizes (c), the molecules are separated by a considerable distance, the intermolecular forces are reduced almost to zero, and the change in enthalpy is much greater.

Fig. 1.21

1.8 BOND ENTHALPY

49

freezes.10 It is always the case that the enthalpy change of a reverse transition is the negative of the enthalpy change of the forward transition (under the same conditions of temperature and pressure): H2O(s) → H2O(l)

DH 3 = +6.01 kJ

H2O(l) → H2O(s)

DH 3 = −6.01 kJ

and in general DforwardH 3 = −D reverse H 3

(1.19)

Dsub H 3 = D fus H 3 + D vap H 3

An implication of the First Law is that the enthalpy change accompanying a reverse process is the negative of the enthalpy change for the forward process.

Fig. 1.22

This relation follows from the fact that H is a state property, so it must return to the same value if a forward change is followed by the reverse of that change (Fig. 1.22). The high standard enthalpy of vaporization of water (+44 kJ mol−1), signifying a strongly endothermic process, implies that the condensation of water (−44 kJ mol−1) is a strongly exothermic process. That exothermicity is the origin of the ability of steam to scald severely because the energy is passed on to the skin. The direct conversion of a solid to a vapor is called sublimation. The reverse process is called vapor deposition. Sublimation can be observed on a cold, frosty morning, when frost vanishes as vapor without first melting. The frost itself forms by vapor deposition from cold, damp air. The vaporization of solid carbon dioxide (‘dry ice’) is another example of sublimation. The standard molar enthalpy change accompanying sublimation is called the standard enthalpy of sublimation, Dsub H 3. Because enthalpy is a state property, the same change in enthalpy must be obtained both in the direct conversion of solid to vapor and in the indirect conversion, in which the solid first melts to the liquid and then that liquid vaporizes (Fig. 1.23): (1.20)

This result is an example of a more general statement that will prove useful time and again during our study of thermochemistry: The enthalpy change of an overall process is the sum of the enthalpy changes for the steps (observed or hypothetical) into which it may be divided.

A brief illustration

To use eqn 1.20 correctly, the two enthalpies that are added together must be for the same temperature, so to get the enthalpy of sublimation of water at 0°C, we must add together the enthalpies of fusion (6.01 kJ mol−1) and vaporization (45.07 kJ mol−1) for this temperature. Adding together enthalpies of transition for different temperatures gives a meaningless result. It follows that Dsub H 3 = D fus H 3 + D vap H 3 = 6.01 kJ mol−1 + 45.07 kJ mol−1 = 51.08 kJ mol−1

1.8 Bond enthalpy To understand bioenergetics at a molecular level we need to account for the flow of energy during chemical reactions as individual chemical bonds are broken and made.

10 This relation is the origin of the obsolescent terms ‘latent heat’ of vaporization and fusion for what are now termed the enthalpy of vaporization and fusion.

Fig. 1.23 The enthalpy of sublimation at a given temperature is the sum of the enthalpies of fusion and vaporization at that temperature. Another implication of the First Law is that the enthalpy change of an overall process is the sum of the enthalpy changes for the possibly hypothetical steps into which it may be divided.

A note on good practice

Molar quantities are expressed as a quantity per mole (as in kilojoules per mole, kJ mol−1). Distinguish them from the magnitude of a property for 1 mol of substance, which is expressed as the quantity itself (as in kilojoules, kJ). All enthalpies of transition, denoted Dtrs H, are molar quantities.

50

1 THE FIRST LAW

Table 1.3

Selected bond enthalpies, H(A–B)/(kJ mol−1)

Diatomic molecules H–H

436

O=O

497

F–F

155

H–F

565

N≡N O–H

945

Cl–Cl

242

H–Cl

431

428

Br–Br

193

H–Br

C≡O

1074

366

I–I

151

H–I

299

Polyatomic molecules H–CH3

435

H–NH2

H–C6H5

469

O2N–NO2

H3C–CH3

368

O=CO

H2C=CH2

699

HC≡CH

962

431

H–OH

492

57

HO–OH

213

531

HO–CH3

377

Cl–CH3

352

Br–CH3

293

I–CH3

234

The thermochemical equation for the dissociation, or breaking, of a chemical bond can be written with the hydroxyl radical OH(g) as an example: HO(g) → H(g) + O(g)

A brief comment

Recall that a radical is a very reactive species containing one or more unpaired electrons. To emphasize the presence of an unpaired electron in a radical, it is common to use a dot (·) when writing the chemical formula. For example, the chemical formula of the hydroxyl radical may be written as ·OH. Hydroxyl radicals and other reactive species containing oxygen can be produced in organisms as undesirable by-products of electron transfer reactions and have been implicated in the development of cardiovascular disease, cancer, stroke, inflammatory disease, and other conditions.

DH 3 = +428 kJ

The corresponding standard molar enthalpy change is called the bond enthalpy, so we would report the H–O bond enthalpy as 428 kJ mol−1. All bond enthalpies are positive, so bond dissociation is an endothermic process. Some bond enthalpies are given in Table 1.3. Note that the bond in molecular nitrogen, N2, is very strong, at 945 kJ mol−1, which helps to account for the chemical inertness of nitrogen and its ability to dilute the oxygen in the atmosphere without reacting with it. In contrast, the bond in molecular fluorine, F2, is relatively weak, at 155 kJ mol−1; the weakness of this bond contributes to the high reactivity of elemental fluorine. However, bond enthalpies alone do not account for reactivity because, although the bond in molecular iodine is even weaker, I2 is less reactive than F2, and the bond in CO is stronger than the bond in N2, but CO forms many carbonyl compounds, such as Ni(CO)4. The types and strengths of the bonds that the elements can make to other elements when a new substance is formed from reactants are additional factors. A complication when dealing with bond enthalpies is that their values depend on the molecule in which the two linked atoms occur. For instance, the total standard enthalpy change for the atomization (the complete dissociation into atoms) of water: H2O(g) → 2 H(g) + O(g)

DH 3 = +927 kJ

is not twice the O–H bond enthalpy in H2O even though two O–H bonds are dissociated. There are in fact two different dissociation steps. In the first step, an O–H bond is broken in an H2O molecule: H2O(g) → HO(g) + H(g)

DH 3 = +492 kJ

In the second step, the O–H bond is broken in an OH radical: HO(g) → H(g) + O(g)

DH 3 = +428 kJ

1.8 BOND ENTHALPY

Table 1.4

H H

436

C

412

Mean bond enthalpies, DHB /(kJ mol−1)* C

N

O

F

Cl

Br

I

S

P

Si

348 (1) 612 (2) 838 (3) 518 (a)†

N

O

388

463

305 (1)

163 (1)

613 (2)

409 (2)

890 (3)

945 (3)

360 (1)

157

146 (1)

743 (2)

497 (2)

F

565

484

270

185

155

Cl

431

338

200

203

254

Br

366

276

I

299

238

S

338

259

P

322

Si

318

242 219

496

193

210

178

250

212

151 264 200

374

466

226

*Values are for single bonds except where otherwise stated (in parentheses). † (a) Denotes aromatic.

The sum of the two steps is the atomization of the molecule. As can be seen from this example, the O–H bonds in H2O and HO have similar but not identical bond enthalpies. Although accurate calculations must use bond enthalpies for the molecule in question and its successive fragments, when such data are not available, there is no choice but to make estimates by using mean bond enthalpies, DHB, which are the averages of bond enthalpies over a related series of compounds (Table 1.4). For example, the mean HO bond enthalpy, DHB(H–O) = 463 kJ mol−1, is the mean of the O–H bond enthalpies in H2O and several other similar compounds, including methanol, CH3OH. Example 1.3

Using mean bond enthalpies

Use information from the Resource section and bond enthalpy data from Tables 1.3 and 1.4 to estimate the standard enthalpy change for the reaction 2 H2O2(l) → 2 H2O(l) + O2(g) in which liquid hydrogen peroxide decomposes into O2 and water at 25°C. In the aqueous environment of biological cells, hydrogen peroxide—a very reactive species—is formed as a result of some processes involving O2. The enzyme catalase helps rid organisms of toxic hydrogen peroxide by accelerating its decomposition. The enthalpy of vaporization of H2O2(l) at 298 K is 51.5 kJ mol−1. Strategy In calculations of this kind, the procedure is to break the overall process down into a sequence of steps such that their sum is the chemical equation required.

51

52

1 THE FIRST LAW

Always ensure, when using bond enthalpies, that all the species are in the gas phase. That may mean including the appropriate enthalpies of vaporization or sublimation. One approach is to atomize all the reactants and then to build the products from the atoms so produced. When explicit bond enthalpies are available (that is, data are given in the tables available), use them; otherwise, use mean bond enthalpies to obtain estimates. Solution The following steps are required:

DH 9/kJ Vaporization of 2 mol H2O2(l), 2 H2O2(l) → 2 H2O2(g) Dissociation of 4 mol O–H bonds Dissociation of 2 mol O–O bonds in HO–OH Overall, so far: 2 H2O2(l) → 4 H(g) + 4 O(g)

2 × (+51.5) 4 × (+463) 2 × (+213) +2381

We have used the mean bond enthalpy value from Table 1.4 for the O–H bond and the exact bond enthalpy value for the O–O bond in HO–OH from Table 1.3. In the second step, four O–H bonds and one O=O bond are formed. The standard enthalpy change for bond formation (the reverse of dissociation) is the negative of the bond enthalpy. We can use exact values for the enthalpy of the O–H bond in H2O(g) and for the O=O bond in O2(g): DH 9/kJ Formation of 4 mol O–H bonds Formation of 1 mol O2 Overall, in this step: 4 O(g) + 4 H(g) → 2 H2O(g) + O2(g)

4 × (−492) −497 −2465

The final stage of the reaction is the condensation of 2 mol H2O(g) 2 H2O(g) → 2 H2O(l)

DH 3 = 2 × (−44 kJ) = −88 kJ

The sum of the enthalpy changes is DH 3 = (+2381 kJ) + (−2465 kJ) + (−88 kJ) = −172 kJ The experimental value is −196 kJ. Estimate the enthalpy change for the reaction between 1 mol C2H5OH as liquid ethanol, a fuel made by fermenting corn, and O2(g) to yield CO2(g) and H2O(l) under standard conditions by using the bond enthalpies, mean bond enthalpies, and the appropriate standard enthalpies of vaporization.

Self-test 1.4

Answer: −1305 kJ; the experimental value is −1368 kJ

1.9 Thermochemical properties of fuels We need to understand the molecular origins of the energy content of biological fuels, the carbohydrates, fats, and proteins.

We saw in Case study 1.1 that photosynthesis and the oxidation of organic molecules are the most important processes that supply energy to organisms. In this section we begin our quantitative study of biological energy conversion by assessing the thermochemical properties of fuels.

1.9 THERMOCHEMICAL PROPERTIES OF FUELS

Table 1.5

Standard enthalpies of combustion

Substance

D c H 9/(kJ mol−1) −394

Carbon, C(s, graphite)

−283

Carbon monoxide, CO(g) Citric acid, C6H8O7(s)

−1985

Ethanol, C2H5OH(l)

−1368

Glucose, C6H12O6(s)

−2808 −969

Glycine, CH2(NH2)COOH(s)

−286

Hydrogen, H2(g) iso-Octane,* C8H18(l)

−5461 −890

Methane, CH4(g)

−726

Methanol, CH3OH(l) Methylbenzene, C6H5CH3(l)

−3910

Octane, C8H18(l)

−5471

Propane, C3H8(g)

−2220 −950

Pyruvic acid, CH3(CO)COOH(l) Sucrose, C12H22O11(s)

−5645

Urea, CO(NH2)2(s)

−632

*2,2,4-Trimethylpentane.

The consumption of a fuel in a furnace or an engine is the result of a combustion. An example is the combustion of methane in a natural gas flame: CH4(g) + 2 O2(g) → CO2(g) + 2 H2O(l)

DH 3 = −890 kJ

The standard enthalpy of combustion, DcH 3, is the standard change in enthalpy per mole of combustible molecules. In this example, we would write DcH 3(CH4, g) = −890 kJ mol−1. Some typical values are given in Table 1.5. Note that DcH 3 is a molar quantity and is obtained from the value of DH 3 by dividing by the amount of organic reactant consumed (in this case, by 1 mol CH4). According to the discussion in Sections 1.5 and 1.6, and the relation DU = qV, the energy transferred as heat at constant volume is equal to the change in internal energy, DU, not DH. To convert from DU to DH, we need to note that the molar enthalpy of a substance is related to its molar internal energy by Hm = Um + pVm (eqn 1.12a). For condensed phases, pVm is so small that it may be ignored. For example, the molar volume of liquid water is 18 cm3 mol−1, and at 1.0 bar pVm = (1.0 × 105 Pa) × (18 × 10−6 m3 mol−1) = 1.8 Pa m3 mol−1 = 1.8 J mol−1 However, the molar volume of a gas, and therefore the value of pVm , is about 1000 times greater and cannot be ignored. For gases treated as perfect, pVm may be replaced by RT. Therefore, if in the chemical equation the difference (products – reactants) in the stoichiometric coefficients of gas phase species is Dngas, we can write Dc H = DcU + Dngas RT Note that ngas (where n is nu) is a dimensionless number.

(1.21)

53

54

1 THE FIRST LAW

A brief illustration

The energy released at constant volume as heat by the combustion of the amino acid glycine is −969.6 kJ mol−1 at 298.15 K, so DcU = −969.6 kJ mol−1. From the chemical equation NH2CH2COOH(s) + 94 O2(g) → 2 CO2(g) + 52 H2O(l) + 12 N2(g) we find that Dngas = (2 + 12 ) − 94 = 14 . Therefore, Dc H = DcU + 14 RT = −969.6 kJ mol−1 + 14 × (8.3145 × 10−3 kJ K−1 mol−1) × (298.15 K) = −969.6 kJ mol−1 + 0.62 kJ mol−1 = −969.0 kJ mol−1

We shall see in Chapter 2 that the best assessment of the ability of a compound to act as a fuel to drive many of the processes occurring in the body makes use of the ‘Gibbs energy’. However, a useful guide to the resources provided by a fuel, and the only one that matters when energy transferred as heat is being considered, is the enthalpy, particularly the enthalpy of combustion. The thermochemical properties of fuels and foods are commonly discussed in terms of their specific enthalpy, the magnitude of the enthalpy of combustion divided by the mass of the sample (typically in kilojoules per gram) or the enthalpy density, the magnitude of the enthalpy of combustion divided by the volume of the sample (typically in kilojoules per cubic decimeter). If the standard enthalpy of combustion is Dc H 3 and the molar mass of the compound is M, then the specific enthalpy is Dc H 3/M. Similarly, the enthalpy density is Dc H 3/Vm, where Vm is the molar volume of the material. Table 1.6 lists the specific enthalpies and enthalpy densities of several fuels. The most suitable fuels are those with high specific enthalpies, as the advantage of a high molar enthalpy of combustion may be eliminated if a large mass of fuel is to be transported. We see that H2 gas compares very well with more traditional fuels such as methane (natural gas), octane (gasoline), and methanol. Furthermore, the combustion of H2 gas does not generate CO2 gas, a pollutant implicated in the mechanism of global warming. As a result, H2 gas has been proposed as an efficient, clean alternative to fossil fuels, such as natural gas and petroleum. However, we also see that H2 gas has a very low enthalpy density, which arises from the fact

Table 1.6

Thermochemical properties of some fuels D c H 9/(kJ mol−1)

Fuel

Combustion equation

Hydrogen

2 H2(g) + O2(g) → 2 H2O(l)

−286

142

13

Methane

CH4(g) + 2 O2(g) → CO2(g) + 2 H2O(l)

−890

55

40

iso-Octane†

2 C8H18(l) + 25 O2(g) → 16 CO2(g) + 18 H2O(l)

−5461

48

3.3 × 104

Methanol

2 CH3OH(l) + 3 O2(g) → 2 CO2(g) + 4 H2O(l)

−726

23

1.8 × 104

*At atmospheric pressures and room temperature. † 2,2,4-Trimethylpentane.

Specific enthalpy/ (kJ g−1)

Enthalpy density*/ (kJ dm−3)

1.9 THERMOCHEMICAL PROPERTIES OF FUELS

55

that hydrogen is a very light gas. So, the advantage of a high specific enthalpy is undermined by the large volume of fuel to be transported and stored. Strategies are being developed to solve the storage problem. For example, the small H2 molecules can travel through holes in the crystalline lattice of a sample of metal, such as titanium, where they bind as metal hydrides. In this way it is possible to increase the effective density of hydrogen atoms to a value that is higher than that of liquid H2. Then the fuel can be released on demand by heating the metal. We now assess the factors that optimize the enthalpy of combustion of carbonbased fuels, with an eye toward understanding such biological fuels as carbohydrates, fats, and proteins. The combustion of 1 mol CH4(g) releases 890 kJ of energy as heat per mole of C atoms: CH4(g) + 2 O2(g) → CO2(g) + 2 H2O(l)

DH 3 = −890 kJ

Now consider the combustion of 1 mol CH3OH(g): CH3OH(g) + 32 O2(g) → CO2(g) + 2 H2O(l)

DH 3 = −765 kJ

This reaction is also exothermic, but now only 765 kJ of energy is released as heat per mole of C atoms. Much of the observed change in energy output between the reactions can be explained by noting that the replacement of a C–H bond by a C–O bond renders the carbon in methanol more oxidized than the carbon in methane, so it is reasonable to expect that less energy is released to complete the oxidation of carbon in methanol to CO2. In general, the presence of partially oxidized C atoms (that is, carbon atoms bonded to oxygen atoms) in a material makes it a less suitable fuel than a similar material containing less oxidized C atoms. Another factor that determines the enthalpy of combustion is the number of carbon atoms in hydrocarbon compounds. For example, whereas the enthalpy of combustion of methane is −890 kJ mol−1, that of iso-octane (C8H18, 2,2,4-trimethylpentane (1), a typical component of gasoline) is −5461 kJ mol−1 (Table 1.6). The much larger value for iso-octane is a consequence of each molecule having eight C atoms to contribute to the formation of carbon dioxide, whereas methane has only one. Case study 1.2

Biological fuels

A typical 18- to 20-year-old man requires a daily energy input of about 12 MJ (1 MJ = 106 J) or about 2870 Cal; a woman of the same age needs about 9 MJ or about 2150 Cal. If the entire consumption were in the form of glucose, which has a specific enthalpy of 16 kJ g−1, meeting energy needs would require the consumption of 750 g of glucose by a man and 560 g by a woman. In fact, the complex carbohydrates more commonly found in our diets have slightly higher specific enthalpies (17 kJ g−1 = 4 Cal g−1) than glucose itself, so a carbohydrate diet is slightly less daunting than a pure glucose diet, as well as being more appropriate in the form of fibre, the indigestible cellulose that helps move digestion products through the intestine. The specific enthalpy of fats, which are long-chain esters such as tristearin (2), is much greater than that of carbohydrates, at around 38 kJ g−1 (9 Cal g−1), slightly less than the value for the hydrocarbon oils used as fuel (48 kJ g−1 = 11 Cal g−1). The reason for this difference lies in the fact that many of the carbon atoms in carbohydrates are bonded to oxygen atoms and are already partially oxidized, whereas most of the carbon atoms in fats are bonded to hydrogen and other carbon atoms and hence have lower oxidation numbers.

A brief comment

The concept of oxidation numbers, familiar from introductory chemistry, clarifies the point made in this paragraph. The formation of CO2 from CH4 involves an increase in the oxidation number—that is, an oxidation —of carbon from −4 in CH4 to +4 in CO2. By contrast, the carbon atom in CH3OH has an oxidation number of −2 and is in a higher oxidation state than the carbon in methane.

56

1 THE FIRST LAW

As we have seen, the presence of partially oxidized carbons lowers the energy output of a fuel. Fats are commonly used as an energy store, to be used only when the more readily accessible carbohydrates have fallen into short supply. In Arctic species, the stored fat also acts as a layer of insulation; in desert species (such as the camel), the fat is also a source of water, one of its oxidation products. Proteins are also used as a source of energy, but their components, the amino acids, are also used to construct other proteins. When proteins are oxidized (to urea, CO(NH2)2), the equivalent specific enthalpy is comparable to that of carbohydrates (about 17 kJ g−1 = 4 Cal g−1).

A brief illustration

A lunch consisting of a hamburger (about 350 Cal), potato chips (1 serving = 108 Cal), and a milk shake (about 502 Cal) would sum to about 960 Cal.11,12 By contrast, a lighter lunch of halibut (about 205 Cal for a 14 -lb serving), a raw carrot (about 42 Cal), a large apple (101 Cal), and a glass of orange juice (about 120 Cal) would net only 468 Cal. The intake from these meals can be compared to the rates at which a 70-kg person can expend energy, depending on the nature of the activity: Level of activity Light (walking slowly) Moderate (walking fast) Heavy (running)

Rate of energy expenditure/(Cal min–1) 2.5–5.0 5.0–7.5 7.5–12.0

It follows that it would be necessary to walk slowly for about 3 to 6 hours (or run for 1 to 2 hours) to expend the energy taken in by eating the 960-Cal hamburger meal. Even though reading this textbook also requires energy, it would take about 16 hours for the hamburger meal to be ‘burned off ’ by so sedentary an activity.

11 Alarmingly, this single meal corresponds to about 33 per cent or 44 per cent of the daily energy requirements of a young man or woman, respectively. 12 The data for this brief illustration are from C.H. Snyder, The extraordinary chemistry of ordinary things, Wiley (2002).

1.10 THE COMBINATION OF REACTION ENTHALPIES

We have already remarked that not all the energy released by the oxidation of foods is used to perform work. The energy that is also released as heat needs to be discarded in order to maintain body temperature within its typical range of 35.6 to 37.8°C. A variety of mechanisms contribute to this aspect of homeostasis, the ability of an organism to counteract environmental changes with physiological responses. The general uniformity of temperature throughout the body is maintained largely by the flow of blood. When energy needs to be dissipated rapidly by heating, warm blood is allowed to flow through the capillaries of the skin, so producing flushing. Radiation is one means of heating the surroundings; another is evaporation and the energy demands of the enthalpy of vaporization of water.

A brief illustration

From the enthalpy of vaporization (D vap H 3 = 44 kJ mol−1 at 298 K), molar mass (M = 18 g mol−1), and mass density ( r = 1.0 g cm−3, corresponding to 1.0 × 103 g dm−3) of water, the energy removed as heat through evaporation per liter (cubic decimeter) of water perspired is q=

rD vapH 3 (1.0 × 103 g dm−3) × (44 kJ mol−1) = = 2.4 × 103 kJ dm−3 M 18 g mol−1 = 2.4 MJ dm−3

When vigorous exercise promotes sweating (through the influence of heat selectors on the hypothalamus), 1 to 2 dm3 of perspired water can be produced per hour, corresponding to a loss of energy of approximately 2.4 to 4.8 MJ h−1.

1.10 The combination of reaction enthalpies To make progress in our study of bioenergetics, we need to develop methods for predicting the reaction enthalpies of complex biochemical reactions.

It is often the case that a reaction enthalpy is needed but is not available in tables of data. Now the fact that enthalpy is a state function comes in handy, because it implies that we can construct the required reaction enthalpy from the reaction enthalpies of known reactions. We have already seen a primitive example when we calculated the enthalpy of sublimation from the sum of the enthalpies of fusion and vaporization. The only difference is that we now apply the technique to a sequence of chemical reactions. The procedure is summarized by Hess’s law, which in its modern form is: The standard enthalpy of a reaction is the sum of the standard enthalpies of the reactions into which the overall reaction may be divided. Although the procedure is given the status of a law, it hardly deserves the title because it is nothing more than a consequence of enthalpy being a state function, which implies that an overall enthalpy change can be expressed as a sum of enthalpy changes for each step in an indirect path. The individual steps need not be actual reactions that can be carried out in the laboratory—they may be entirely hypothetical reactions, the only requirement being that their equations should balance. Each step must correspond to the same temperature.

57

58

1 THE FIRST LAW

Example 1.4

Using Hess’s law

In biological cells that have a plentiful supply of O2, glucose is oxidized completely to CO2 and H2O (Section 1.9 and Case study 1.2). Muscle cells may be deprived of O2 during vigorous exercise and, in that case, one molecule of glucose is converted to two molecules of lactic acid (Atlas C2) by the process of glycolysis (Case study 4.3). Given the thermochemical equations for the combustions of glucose and lactic acid: C6H12O6(s) + 6 O2(g) → 6 CO2(g) + 6 H2O(l)

DH 3 = −2808 kJ

CH3CH(OH)COOH(s) + 3 O2(g) → 3 CO2(g) + 3 H2O(l) DH 3 = −1344 kJ calculate the standard enthalpy for glycolysis: C6H12O6(s) → 2 CH3CH(OH)COOH(s) Is there a biological advantage of complete oxidation of glucose compared with glycolysis? Explain your answer. Strategy We need to add or subtract the thermochemical equations so as to reproduce the thermochemical equation for the reaction required. Solution We obtain the thermochemical equation for glycolysis from the

following sum: DH 9/kJ C6H12O6(s) + 6 O2(g) → 6 CO2(g) + 6 H2O(l)

−2808

6 CO2(g) + 6 H2O(l) → 2 CH3CH(OH)COOH(s) + 6 O2(g)

2 × (+1344 kJ)

Overall: C6H12O6(s) → 2 CH3CH(OH)COOH(s)

−120

It follows that the standard enthalpy for the conversion of glucose to lactic acid during glycolysis is −120 kJ mol−1, a mere 4 per cent of the enthalpy of combustion of glucose. Therefore, full oxidation of glucose is metabolically more useful than glycolysis because in the former process more energy becomes available for performing work. Calculate the standard enthalpy of the fermentation C6H12O6(s) → 2 C2H5OH(l) + 2 CO2(g) from the standard enthalpies of combustion of glucose and ethanol (Table 1.5). Self-test 1.5

Answer: −72 kJ

1.11 Standard enthalpies of formation We need to simplify even further the process of predicting reaction enthalpies of biochemical reactions.

The standard reaction enthalpy, D r H 3, is the difference between the standard molar enthalpies of the reactants and the products, with each term weighted by the stoichiometric coefficient, n (nu), in the chemical equation D r H 3 = ∑ nH m3(products) − ∑ nH m3(reactants)

Definition of the standard reaction enthalpy

(1.22)

where ∑ (uppercase sigma) denotes a sum. Because the H m3 are molar quantities and the stoichiometric coefficients are pure numbers, the units of D r H 3 are

1.11 STANDARD ENTHALPIES OF FORMATION

kilojoules per mole. The standard reaction enthalpy is the change in enthalpy of the system when the reactants in their standard states (pure, 1 bar) are completely converted into products in their standard states (pure, 1 bar), with the change expressed in kilojoules per mole of reaction as written. The problem with eqn 1.22 is that we have no way of knowing the absolute enthalpies of the substances. To avoid this problem, we can imagine the reaction as taking place by an indirect route, in which the reactants are first broken down into the elements and then the products are formed from the elements (Fig. 1.24). Specifically, the standard enthalpy of formation, D f H 3, of a substance is the standard enthalpy (per mole of the substance) for its formation from its elements in their reference states. The reference state of an element is its most stable form under the prevailing conditions (Table 1.7). Don’t confuse ‘reference state’ with ‘standard state’: the reference state of carbon at 25°C is graphite (not diamond); the standard state of carbon is any specified phase of the element at 1 bar. For example, the standard enthalpy of formation of liquid water (at 25°C, as always in this text) is obtained from the thermochemical equation H2(g) + 12 O2(g) → H2O(l)

DH 3 = −286 kJ

and is Df H (H2O, l) = −286 kJ mol . Note that enthalpies of formation are molar quantities, so to go from DH 3 in a thermochemical equation to D f H 3 for that substance, divide by the amount of substance formed (in this instance, by 1 mol H2O). With the introduction of standard enthalpies of formation, we can write 3

−1

D r H 3 = ∑ nD f H 3(products) − ∑ nD f H 3(reactants)

Calculation of standard reaction enthalpies

(1.23)

The first term on the right is the enthalpy of formation of all the products from their elements; the second term on the right is the enthalpy of formation of all the reactants from their elements. The fact that the enthalpy is a state function means that a reaction enthalpy calculated in this way is identical to the value that would be calculated from eqn 1.22 if absolute enthalpies were available. The values of some standard enthalpies of formation at 25°C are given in Table 1.8, and a longer list is given in the Resource section. The standard enthalpies of formation of elements in their reference states are zero by definition (because their formation is the null reaction: element → element). Note, however, that the standard enthalpy of formation of an element in a state other than its reference state is not zero: C(s, graphite) → C(s, diamond)

DH 3 = +1.895 kJ

Therefore, although D f H 3(C, graphite) = 0, D f H 3(C, diamond) = +1.895 kJ mol−1.

Example 1.5

Using standard enthalpies of formation

Glucose and fructose (Atlas S3) are simple carbohydrates with the molecular formula C6H12O6. Sucrose (Atlas S5), or table sugar, is a complex carbohydrate with molecular formula C12H22O11 that consists of a glucose unit covalently linked to a fructose unit (a water molecule is released as a result of the reaction between glucose and fructose to form sucrose). Estimate the standard enthalpy of combustion of sucrose from the standard enthalpies of formation of the reactants and products.

59

An enthalpy of reaction may be expressed as the difference between the enthalpies of formation of the products and the reactants.

Fig. 1.24

Table 1.7 Reference states of some elements at 298.15 K

Element

Reference state

Arsenic

Gray arsenic

Bromine

Liquid, Br2(l)

Carbon

Graphite

Hydrogen

Gas, H2(g)

Iodine

Solid, I2(s)

Mercury

Liquid

Nitrogen

Gas, N2(g)

Oxygen

Gas, O2(g)

Phosphorus

White phosphorus, P4(s)

Sulfur

Rhombic sulfur, S8(s)

60

1 THE FIRST LAW

Table 1.8

Standard enthalpies of formation at 298.15 K*

Substance

D f H 9/(kJ mol−1)

Inorganic compounds Ammonia, NH3(g)

Substance Organic compounds

−46.11

Adenine, C5H5N5(s)

Carbon monoxide, CO(g)

−110.53

Alanine, CH3CH(NH2)COOH(s) Benzene, C6H6(l)

Carbon dioxide, CO2(g)

−393.51

Hydrogen sulfide, H2S(g)

−20.63

Butanoic acid, CH3(CH2)2COOH(l)

Nitrogen dioxide, NO2(g)

+33.18

Ethane, C2H6(g)

Nitrogen monoxide, NO(g)

D f H 9/(kJ mol−1)

+90.25

Ethanoic acid, CH3COOH(l)

Sodium chloride, NaCl(s)

−411.15

Ethanol, C2H5OH(l)

Water, H2O(l)

−285.83

a-d-Glucose, C6H12O6(s)

H2O(g)

−241.82

+96.9 −604.0 +49.0 −533.8 −84.68 −484.3 −277.69 −1268

Guanine, C5H5N5O(s)

−183.9

Glycine, CH2(NH2)COOH(s)

−528.5

N-Glycylglycine, C4H8N2O3(s)

−747.7

Hexadecanoic acid, CH3(CH2)14COOH(s)

−891.5

Leucine, (CH3)2CHCH2CH(NH2)COOH(s)

−637.4

Methane, CH4(g) Methanol, CH3OH(l) Sucrose, C12H22O11(s)

−74.81 −238.86 −2222

Thymine, C5H6N2O2(s)

−462.8

Urea, (NH2)2CO(s)

−333.1

*A longer list is given in the Resource section.

Strategy We write the chemical equation, identify the stoichiometric numbers of the reactants and products, and then use eqn 1.23. Note that the expression has the form ‘products – reactants’. Numerical values of standard enthalpies of formation are given in the Resource section. The standard enthalpy of combustion is the enthalpy change per mole of substance, so we need to interpret the enthalpy change accordingly. Solution The chemical equation is

C12H22O11(s) + 12 O2(g) → 12 CO2(g) + 11 H2O(l) It follows that A note on good practice

The standard enthalpy of formation of an element in its reference state (oxygen gas in this example) is written 0, not 0 kJ mol−1, because it is zero whatever units we happen to be using.

D r H 3 = {12Df H 3(CO2,g) + 11D f H 3(H2O,l)} − {D f H 3(C12H22O11,g) + 12Df H 3(O2,g)} = {12 × (−393.51 kJ mol−1) + 11 × (−285.83 kJ mol−1)} − {(−2222 kJ mol−1) + 0} = −5644 kJ mol−1 Inspection of the chemical equation shows that, in this instance, the ‘per mole’ is per mole of sucrose, which is exactly what we need for an enthalpy of combustion. It follows that the estimate for the standard enthalpy of combustion of sucrose is −5644 kJ mol−1. The experimental value is −5645 kJ mol−1.

1.12 ENTHALPIES OF FORMATION AND COMPUTATIONAL CHEMISTRY

61

Use standard enthalpies of formation to calculate the enthalpy of combustion of solid glycine to CO2(g), H2O(l), and N2(g).

Self-test 1.6

Answer: −973 kJ mol−1, in agreement with the experimental value (see the Resource section)

The reference states of the elements define a thermochemical ‘sea level’, and enthalpies of formation can be regarded as thermochemical ‘altitudes’ above or below sea level (Fig. 1.25). Compounds that have negative standard enthalpies of formation (such as water) are classified as exothermic compounds, for they lie at a lower enthalpy than their component elements (they lie below thermochemical sea level). Compounds that have positive standard enthalpies of formation (such as carbon disulfide) are classified as endothermic compounds and possess a higher enthalpy than their component elements (they lie above sea level). 1.12 Enthalpies of formation and computational chemistry Table 1.4 is useful for many calculations, but its data cannot be used to estimate the differences between the standard enthalpies of formation of conformational isomers. For example, we would obtain the same enthalpy of formation for the equatorial and axial conformers of methylcyclohexane (3 and 4, respectively) if we were to use mean bond enthalpies. However, it has been observed experimentally that these conformers have different standard enthalpies of formation due to the steric repulsions in the axial conformer, which raise its energy relative to that of the equatorial conformer. Computational chemistry is becoming the technique of choice for estimating standard enthalpies of formation of molecules with complex three-dimensional structures. Commercial software packages use the principles developed in Chapter 10 to calculate the standard enthalpy of formation of a conformer drawn on a computer screen. The difference between calculated standard enthalpies of formation of two conformers is then an estimate of the conformational energy difference. In the case of methylcyclohexane, the calculated conformational energy difference ranges from 5.9 to 7.9 kJ mol−1, with the equatorial conformer having a lower standard enthalpy of formation than the axial conformer. These estimates compare favorably with the experimental value of 7.5 kJ mol−1. However, good agreement between calculated and experimental values is relatively rare. Computational methods almost always predict correctly which conformer is more stable but do not always predict the correct magnitude of the conformational energy difference. The computational approach also makes it possible to gain insight into the effect of solvation on the enthalpy of formation without conducting experiments. A calculation performed in the absence of solvent molecules estimates the properties of the molecule of interest in the gas phase. Computational methods are available that allow for the inclusion of several solvent molecules around a solute molecule, thereby taking into account the effect of molecular interactions with the solvent on the enthalpy of formation of the solute. Again, the numerical results are only estimates, and the primary purpose of the calculation is to predict whether interactions with the solvent increase or decrease the enthalpy of formation. As an example, consider the amino acid glycine, which can exist in a neutral (5) or zwitterionic (6) form, in which the amino group is protonated and the

The enthalpy of formation acts as a kind of thermochemical ‘altitude’ of a compound with respect to the ‘sea level’ defined by the elements from which it is made. Endothermic compounds have positive enthalpies of formation; exothermic compounds have negative energies of formation.

Fig. 1.25

62

1 THE FIRST LAW

carboxyl group is deprotonated. It is possible to show computationally that in the gas phase the neutral form has a lower enthalpy of formation than the zwitterionic form. However, in water the opposite is true because of strong interactions between the polar solvent and the charges in the zwitterion. 1.13 The variation of reaction enthalpy with temperature We need to know how to predict the reaction enthalpy of a biochemical reaction at one temperature from its value at other temperatures.

Fig. 1.26 The enthalpy of a substance increases with temperature, therefore if the total enthalpy of the reactants increases by a different amount from that of the products, the reaction enthalpy will change with temperature. The change in reaction enthalpy depends on the relative slopes of the two lines and hence on the heat capacities of the substances.

Suppose we want to know the enthalpy of a particular reaction at body temperature, 37°C, but have data available for 25°C, or suppose we want to know whether the oxidation of glucose is more exothermic when it takes place inside an Arctic fish that inhabits water at 0°C than when it takes place at mammalian body temperatures. In precise work, every attempt would be made to measure the reaction enthalpy at the temperature of interest, but it is useful to have a rapid way of estimating the sign and even a moderately reliable numerical value. Figure 1.26 illustrates the technique. As we have seen, the enthalpy of a substance increases with temperature; therefore the total enthalpy of the reactants and the total enthalpy of the products increase, as shown in the illustration. Provided the two total enthalpy increases are different, the standard reaction enthalpy (their difference) will change as the temperature is changed. The change in the enthalpy of a substance depends on the slope of the graph and therefore on the constant-pressure heat capacities of the substances (recall Fig. 1.16). We can therefore expect the temperature dependence of the reaction enthalpy to be related to the difference in heat capacities of the products and the reactants. We show in the following Justification that this is indeed the case and that, when the heat capacities do not vary with temperature, the standard reaction enthalpy at a temperature T is related to the value at a different temperature T by a special formulation of Kirchhoff ’s law: D r H 3(T′) = D r H 3(T) + (T ′ − T)D rC 3p

Kirchhoff’s law

(1.24)

where D rC 3p is the difference between the weighted sums of the standard molar heat capacities of the products and the reactants: D rC 3p = ∑ nC 3p,m(products) − ∑ nC 3p,m(reactants)

(1.25)

Values of standard molar constant-pressure heat capacities for a number of substances are given in the Resource section. Because eqn 1.24 applies only when the heat capacities are constant over the range of temperature of interest, its use is restricted to small temperature differences (of no more than 100 K or so). Justification 1.4 Kirchhoff’s law

To derive Kirchhoff ’s law, we consider the variation of the enthalpy with temperature. We begin by rewriting eqn 1.14b to calculate the change in the standard molar enthalpy Hm of each reactant and product as the temperature of the reaction mixture is increased: dH 3m = C 3p,mdT where C 3p,m is the standard molar constant-pressure heat capacity, the molar heat capacity at 1 bar. We proceed by integrating both sides of the expression

1.13 THE VARIATION OF REACTION ENTHALPY WITH TEMPERATURE for dH 3m from an initial temperature T and initial enthalpy H 3m(T ) to a final temperature T′ and enthalpy H 3m(T′): H m3 (T′)

T′

dH = 3 m

C 3p,m dT

T

H (T )

It follows that for each reactant and product (assuming that no phase transition takes place in the temperature range of interest)

冮C T′

H m3(T′) = H m3(T) +

3 p,m

dT

T

Because this equation applies to each substance in the reaction, we use it and eqn 1.23 to write the following expression for D r H 3(T′): D r H 3(T′) = D r H 3(T) +

T′

DrC 3p dT T

3 p

where DrC is given by eqn 1.25. This equation is the exact form of Kirchhoff ’s law. The special case given by eqn 1.24 can be derived readily from it by making the approximation that DrC p3 is independent of temperature. Then the integral on the right evaluates to

冮 dT = D C × (T′ − T)

T′

T′

DrC 3pdT = DrC 3p T

r

3 p

T

and we obtain eqn 1.24.

Example 1.6

Using Kirchhoff’s law

The enzyme glutamine synthetase mediates the synthesis of the amino acid glutamine (Gln, 8) from the amino acid glutamate (Glu, 7) and ammonium ion: D r H 3 = +21.8 kJ mol−1 at 25°C The process is endothermic and requires energy extracted from the oxidation of biological fuels and stored in ATP (Case study 1.1). Estimate the value of the reaction enthalpy at 60°C by using data found in this text (see the Resource section) 3 (Gln, aq) = 187.0 J K−1 mol−1 and the following additional information: C p,m 3 −1 −1 and C p,m(Glu, aq) = 177.0 J K mol . Strategy Calculate the value of DrC p3 from the available data and eqn 1.25 and

use the result in eqn 1.24. Solution From the Resource section, the standard molar constant-pressure heat

capacities of H2O(l) and NH+4(aq) are 75.3 J K−1 mol−1 and 79.9 J K−1 mol−1, respectively. It follows that 3 3 3 3 DrC p3 = {C p,m (Gln, aq) + C p,m (H2O, l)} − {C p,m (Glu, aq) + C p,m (NH+4 , aq)}

= {(187.0 J K−1 mol−1) + (75.3 J K−1 mol−1)} − {(177.0 J K−1 mol−1) + (79.9 J K−1 mol−1)} = +5.4 J K−1 mol−1 = +5.4 × 10−3 kJ K−1 mol−1

63

A note on good practice

Because heat capacities can be measured more accurately than some reaction enthalpies, the exact form of Kirchhoff ’s law, with numerical integration of DrC 3p over the temperature range of interest, sometimes gives results more accurate than a direct measurement of the reaction enthalpy at the second temperature.

64

1 THE FIRST LAW Then, because T ′ − T = +35 K, from eqn 1.24 we find Dr H 3(333 K) = (+21.8 kJ mol−1) + (5.4 × 10−3 kJ K−1 mol−1) × (35 K) = (+21.8 kJ mol−1) + (0.19 kJ mol−1) = +22.0 kJ mol−1 Estimate the standard enthalpy of combustion of solid glycine at 340 K from the data in Self-test 1.6 and the Resource section. Self-test 1.7

Answer: −9683 kJ mol−1

The calculation in Example 1.6 shows that the standard reaction enthalpy at 60°C is only slightly different from that at 25°C. The reason is that the change in reaction enthalpy is proportional to the difference between the molar heat capacities of the products and the reactants, which is usually not very large. It is generally the case that provided the temperature range is not too wide, enthalpies of reactions vary only slightly with temperature. A reasonable first approximation is that standard reaction enthalpies are independent of temperature. However, notable exceptions are processes involving the unfolding of macromolecules, such as proteins (In the laboratory 1.1). The difference in molar heat capacities between the folded and unfolded states of proteins is usually rather large, in the order of a few kilojoules per mole, so the enthalpy of protein unfolding varies significantly with temperature.

Checklist of key concepts 1. A system is classified as open, closed, or isolated. 2. The surroundings remain at constant temperature and either constant volume or constant pressure when processes occur in the system. 3. An exothermic process releases energy as heat, q, to the surroundings; an endothermic process absorbs energy as heat. 4. Metabolism is the collection of chemical reactions that trap, store, and utilize energy in biological cells. 5. Catabolism is the collection of reactions associated with the oxidation of nutrients in the cell. Anabolism is the biosynthesis of small and large molecules. 6. Maximum expansion work is achieved in a reversible change.

9. The standard state of a substance is the pure substance at 1 bar. 10. Bomb calorimetry is a useful technique for the study of nutrients. 11. Isobaric calorimetry is a useful technique for the study of fuels. 12. Differential scanning calorimetry (DSC) is a useful technique for the investigation of phase transitions, especially those observed in biological macromolecules. 13. The standard enthalpy of transition, D trs H 3, is the change in molar enthalpy when a substance in one phase changes into another phase, both phases being in their standard states.

7. The First Law of thermodynamics states that the internal energy of an isolated system is constant.

14. The standard enthalpy of the reverse of a process is the negative of the standard enthalpy of the forward process, D reverseH 3 = −D forwardH 3.

8. A change in internal energy is equal to the energy transferred as heat at constant volume (DU = qV); a change in enthalpy is equal to the energy transferred as heat at constant pressure (DH = qp).

15. The standard enthalpy of a process is the sum of the standard enthalpies of the individual processes into which it may be regarded as divided, as in D sub H 3 = D fus H 3 + D vap H 3.

EXERCISES

16. Hess’s law states that the standard enthalpy of a reaction is the sum of the standard enthalpies of the reactions into which the overall reaction can be divided. 17. The standard enthalpy of formation of a compound, D f H 3, is the standard reaction enthalpy for the

65

formation of the compound from its elements in their reference states. 18. At constant pressure, exothermic compounds are those for which D f H 3 < 0; endothermic compounds are those for which D f H 3 > 0.

Checklist of key equations Property or process

Equation

Comment

Work of expansion

w = −pexDV

Constant pressure

Heat capacity

C = q/DT

General definition

Change in internal energy

DU = w + q

Constant-volume heat capacity

CV = dU/dT

Definition

Enthalpy

H = U + pV

Definition

Enthalpy change

DH = DU + pDV

Constant pressure

Constant-pressure heat capacity

Cp = dH/dT

Definition

Difference between the molar heat capacities

Cp,m − CV,m = R

Standard reaction enthalpy

D r H = ∑ nH (products) − ∑ nH (reactants)

Definition

D r H 3 = ∑ nD f H 3(products) − ∑ nDf H 3(reactants)

Practical implementation

D r H 3(T′) = Dr H 3(T) + D rC 3p (T ′ − T)

Constant-pressure heat capacities are independent of temperature

Kirchhoff ’s law

3

3 m

Perfect gas 3 m

Discussion questions 1.1 Provide molecular interpretations of work, heat, temperature, and heat capacity. 1.2 Suggest a reason why most molecules survive for long periods at room temperature. 1.3 Describe the general patterns of energy conversion in living organisms. 1.4 Explain the difference between the change in internal energy and

the change in enthalpy of a chemical or physical process. 1.5 Explain the limitations of the following expressions:

(a) DH = DU + pDV; (b) Dr H 3(T′) = Dr H 3(T) + DrC 3p × (T′ − T).

1.6 A primitive air-conditioning unit for use in places where electrical power is not available can be made by hanging up strips of linen soaked in water. Explain why this strategy is effective. 1.7 In many experimental thermograms, such as that shown in Fig. 1.20, the baseline below T1 is at a different level from that above T2. Explain this observation. 1.8 Describe at least two calculational methods by which standard reaction enthalpies can be predicted. Discuss the advantages and disadvantages of each method. 1.9 Distinguish between (a) the standard state and the reference state of an element; (b) endothermic and exothermic compounds.

Exercises Assume all gases are perfect unless stated otherwise. All thermochemical data are for 298.15 K. 1.10 The unit 1 electronvolt (1 eV) is defined as the energy acquired

by an electron as it moves through a potential difference of 1 V. Suppose two states differ in energy by 1.0 eV. What is the ratio of their populations at (a) 300 K and (b) 3000 K?

1.11 How much metabolic energy must a bird of mass 200 g

expend to fly to a height of 20 m? Neglect all losses due to friction, physiological imperfection, and the acquisition of kinetic energy. 1.12 Calculate the work of expansion accompanying the complete

combustion of 1.0 g of glucose to carbon dioxide and (a) liquid

66

1 THE FIRST LAW

water, and (b) water vapor at 20°C when the external pressure is 1.0 atm. 1.13 We are all familiar with the general principles of operation of an

internal combustion reaction: the combustion of fuel drives out the piston. It is possible to imagine engines that use reactions other than combustions, and we need to assess the work they can do. A chemical reaction takes place in a container of cross-sectional area 100 cm2; the container has a piston at one end. As a result of the reaction, the piston is pushed out through 10.0 cm against a constant external pressure of 100 kPa. Calculate the work done by the system. 1.14 A sample of methane of mass 4.50 g occupies 12.7 dm3 at 310 K.

Calculate the work done when the gas expands (a) isobarically against a constant external pressure of 30.0 kPa until its volume has increased by 3.3 dm3, and (b) isothermally by 3.3 dm3. 1.15 The heat capacity of air is much smaller than that of water, and relatively modest amounts of heat are needed to change its temperature. This is one of the reasons why desert regions, although very hot during the day, are bitterly cold at night. The heat capacity of air at room temperature and pressure is approximately 21 J K−1 mol−1. How much energy is required to raise the temperature of a room of dimensions 5.5 m × 6.5 m × 3.0 m by 10°C? If losses are neglected, how long will it take a heater rated at 1.5 kW to achieve that increase given that 1 W = 1 J s−1? 1.16 The transfer of energy from one region of the atmosphere to another is of great importance in meteorology for it affects the weather. Calculate the heat needed to be supplied to a parcel of air containing 1.00 mol air molecules to maintain its temperature at 300 K when it expands reversibly and isothermally from 22 dm3 to 30.0 dm3 as it ascends. 1.17 A laboratory animal exercised on a treadmill, which, through

pulleys, raised a mass of 200 g through 1.55 m. At the same time, the animal lost 5.0 J of energy as heat. Disregarding all other losses and regarding the animal as a closed system, what is its change in internal energy? 1.18 A sample of a serum of mass 25 g is cooled from 290 K to 275 K

at constant pressure by the extraction of 1.2 kJ of energy as heat. Calculate q and DH and estimate the heat capacity of the sample. 1.19 (a) Show that for a perfect gas, Cp,m − CV,m = R. (b) When 229 J

of energy is supplied as heat at constant pressure to 3.00 mol CO2(g), the temperature of the sample increases by 2.06 K. Calculate the molar heat capacities at constant volume and constant pressure of the gas. 1.20 Use the information in Exercise 1.19 to calculate the change in

(a) molar enthalpy and (b) molar internal energy when carbon dioxide is heated from 15°C (the temperature when air is inhaled) to 37°C (blood temperature, the temperature in our lungs). 1.21 Suppose that the molar internal energy of a substance over a

limited temperature range could be expressed as a polynomial in T as Um(T) = a + bT + cT 2. Find an expression for the constant-volume molar heat capacity at a temperature T. 1.22 The heat capacity of a substance is often reported in the form Cp,m = a + bT + c/T 2. Use this expression to make a more accurate estimate of the change in molar enthalpy of carbon dioxide when it is heated from 15°C to 37°C (as in Exercise 1.20), given a = 44.22 J K−1 mol−1, b = 8.79 × 10−3 J K−2 mol−1, and c = −8.62 × 105 J K mol−1. Hint: You will need to integrate dH = CpdT. 1.23 Exercise 1.22 gives an expression for the temperature

dependence of the constant-pressure molar heat capacity over a

limited temperature range. (a) How does the molar enthalpy of the substance change over that range? (b) Plot the molar enthalpy as a function of temperature using the data in Exercise 1.22. 1.24 Classify as endothermic or exothermic (a) a combustion

reaction for which Dr H 3 = −2020 kJ mol−1, (b) a dissolution for which DH 3 = +4.0 kJ mol−1, (c) vaporization, (d) fusion, and (e) sublimation. 1.25 The pressures deep within the Earth are much greater than those on the surface, and to make use of thermochemical data in geochemical assessments we need to take the differences into account. (a) Given that the enthalpy of combustion of graphite is −393.5 kJ mol−1 and that of diamond is −395.41 kJ mol−1, calculate the standard enthalpy of the C(s, graphite) → C(s, diamond) transition. (b) Use the information in part (a) together with the densities of graphite (2.250 g cm−3) and diamond (3.510 g cm−3) to calculate the internal energy of the transition when the sample is under a pressure of 150 kbar. 1.26 A typical human produces about 10 MJ of energy transferred as

heat each day through metabolic activity. (a) If a human body were an isolated system of mass 65 kg with the heat capacity of water, what temperature rise would the body experience? (b) Human bodies are actually open systems, and the main mechanism of heat loss is through the evaporation of water. What mass of water should be evaporated each day to maintain constant temperature? 1.27 Use the information in Tables 1.1 and 1.2 to calculate the total

heat required to melt 100 g of ice at 0°C, heat it to 100°C, and then vaporize it at that temperature. Sketch a graph of temperature against time on the assumption that the sample is heated at a constant rate. 1.28 In preparation for a study of the metabolism of an organism,

a small, sealed calorimeter was assembled. In the initial phase of the experiment, a current of 22.22 mA from an 11.8 V source was passed for 162 s through a heater inside the calorimeter. What is the change in internal energy of the calorimeter? 1.29 Water is heated to boiling under a pressure of 1.0 atm. When

an electric current of 0.50 A from a 12 V supply is passed for 300 s through a resistance in thermal contact with it, it is found that 0.798 g of water is vaporized. Calculate the molar internal energy and enthalpy changes at the boiling point (373.15 K). 1.30 In an experiment to determine the energy content of a food,

a sample of the food was burned in an oxygen atmosphere and the temperature rose by 2.89°C. When a current of 1.27 A from a 12.5 V source flowed through the same calorimeter for 157 s, the temperature rose by 3.88°C. What energy was released as heat by the combustion? 1.31 A sample of the sugar d-ribose (C5H10O5) of mass 0.727 g was placed in a calorimeter and then ignited in the presence of excess oxygen. The temperature rose by 0.910 K. In a separate experiment in the same calorimeter, the combustion of 0.917 g of benzoic acid, for which the internal energy of combustion is −3226 kJ mol−1, gave a temperature rise of 1.940 K. Calculate the enthalpy of formation of d-ribose. 1.32 Figure 1.27 shows the experimental DSC scan of hen white

lysozyme (G. Privalov et al., Anal. Biochem. 79, 232 (1995)) converted to kilojoules (from calories). Determine the enthalpy of unfolding of this protein by integration of the curve and the change in heat capacity accompanying the transition. 1.33 The mean bond enthalpies of C–C, C–H, C=O, and O–H bonds are 348, 412, 743, and 463 kJ mol−1, respectively. The combustion of a fuel such as octane is exothermic because relatively weak bonds

EXERCISES

67

1.38 Estimate the difference between the standard enthalpy of

formation of H2O(l) as currently defined (at 1 bar) and its value using the former definition (at 1 atm). 1.39 Use information in the Resource section to calculate the standard

enthalpies of the following reactions: (a) the hydrolysis of a glycine–glycine dipeptide: +

NH3CH2CONHCH2CO2−(s) + H2O(l) → 2 +NH3CH2CO2−(aq)

(b) the combustion of solid b-d-fructose (c) the dissociation of nitrogen dioxide, which occurs in the atmosphere: NO2(g) → NO(g) + O(g) 1.40 During glycolysis, glucose is partially oxidized to pyruvic acid,

CH3COCOOH, by NAD+ (see Chapter 4) without the involvement of O2. However, it is also possible to carry out the oxidation in the presence of O2: Fig. 1.27

Experimental DSC scan of hen white lysozyme.

C6H12O6(s) + O2(g) → 2 CH3COCOOH(s) + 2 H2O(l) Dr H 3 = −480.7 kJ mol−1

break to form relatively strong bonds. Use this information to justify why glucose has a lower specific enthalpy than the lipid decanoic acid (C10H20O2) even though these compounds have similar molar masses.

From these data and additional information in the Resource section, calculate the standard enthalpy of combustion and standard enthalpy of formation of pyruvic acid.

1.34 Use bond enthalpies and mean bond enthalpies to estimate

lysozyme is +217.6 kJ mol−1 and the change in the constant-pressure molar heat capacity resulting from denaturation of the protein is +6.3 kJ K−1 mol−1. (a) Estimate the enthalpy of denaturation of the protein at (i) 351 K, the ‘melting’ temperature of the macromolecule, and (ii) 263 K. State any assumptions in your calculations. (b) Based on your answers to part (a), is denaturation of hen egg white lysozyme always endothermic?

(a) the enthalpy of the anaerobic breakdown of glucose to lactic acid in cells that are starved of O2, C6H12O6(aq) → 2 CH3CH(OH) COOH(aq), and (b) the enthalpy of combustion of glucose. Ignore the contributions of enthalpies of fusion and vaporization. 1.35 Glucose and fructose are simple sugars with the molecular

formula C6H12O6. Sucrose (table sugar) is a complex sugar with molecular formula C12H22O11 that consists of a glucose unit covalently bound to a fructose unit (a water molecule is eliminated as a result of the reaction between glucose and fructose to form sucrose). (a) Calculate the energy released as heat when a typical table sugar cube of mass 1.5 g is burned in air. (b) To what height could you climb on the energy a table sugar cube provides assuming 25 per cent of the energy is available for work? (c) The mass of a typical glucose tablet is 2.5 g. Calculate the energy released as heat when a glucose tablet is burned in air. (d) To what height could you climb on the energy a tablet provides assuming 25 per cent of the energy released by the metabolism of glucose is available for work? 1.36 Camping gas is typically propane. The standard enthalpy of

combustion of propane gas is −2220 kJ mol−1 and the standard enthalpy of vaporization of the liquid is +15 kJ mol−1. Calculate (a) the standard enthalpy and (b) the standard internal energy of combustion of the liquid. 1.37 Ethane is flamed off in abundance from oil wells because it is

unreactive and difficult to use commercially. But would it make a good fuel? The standard enthalpy of reaction for 2 C2H6(g) + 7 O2(g) → 4 CO2(g) + 6 H2O(l) is −3120 kJ. (a) What is the standard enthalpy of combustion of ethane? (b) What is the specific enthalpy of combustion of ethane? (c) Is ethane a more or less efficient fuel than methane?

1.41 At 298 K, the enthalpy of denaturation of hen egg white

1.42 Estimate the enthalpy of vaporization of water at 100°C from

its value at 25°C (+44.01 kJ mol−1) given the constant-pressure heat capacities of 75.29 J K−1 mol−1 and 33.58 J K−1 mol−1 for liquid and gas, respectively. 1.43 Is the standard enthalpy of combustion of glucose likely to be

higher or lower at blood temperature than at 25°C? 1.44 Using the fact that the enthalpy is a state function, derive a

version of Kirchhoff ’s law (eqn 1.24) by adding contributions from the following processes: (a) the enthalpy change when the reactants are cooled from a temperature T to 298 K, (b) the reaction enthalpy at 298 K, and (c) the enthalpy change when the temperature of the products is increased from 298 K to T. 1.45 Derive a version of Kirchhoff ’s law (eqn 1.24) for the

temperature dependence of the internal energy of reaction. 1.46 The formulation of Kirchhoff ’s law given in eqn 1.24 is valid

when the difference in heat capacities is independent of temperature over the temperature range of interest. Suppose instead that DrC 3p = a + bT + c/T 2. Derive a more accurate form of Kirchhoff ’s law in terms of the parameters a, b, and c. Hint: The change in the reaction enthalpy for an infinitesimal change in temperature is DrC 3p dT. Integrate this expression between the two temperatures of interest.

68

1 THE FIRST LAW

Projects 1.47 The Boltzmann distribution can be used to calculate the average energy associated with each mode of motion of a molecule. However, for certain modes of motion, such as translation, there is a short cut, called the equipartition theorem:

In a sample at a temperature T, all quadratic contributions to the total energy have the same mean value, namely 12 kT. A ‘quadratic contribution’ means a contribution that depends on the square of the position or the velocity (or momentum). (a) The kinetic energy of a particle of mass m free to undergo translation in three dimensions is Ek = 12 mv x2 + 12 mv y2 + 12 mv z2. What is the average kinetic energy of a particle free to move in three dimensions? (b) Use the equipartition theorem to show that for a monatomic perfect gas: Um(T) = Um(0) + 32 RT

CV,m = 32 R

where Um(0) is the molar internal energy at T = 0, when all translational motion has ceased. (c) When the gas consists of molecules, we need to take into account the effect of rotation and vibration. A linear molecule, such as N2 and CO2, can rotate around two axes perpendicular to the line of the atoms, so it has two rotational modes of motion, each contributing a term 12 kT to the internal energy. Show that Um(T) = Um(0) + 52 RT CV,m = 52 R (linear molecule, translation and rotation only) (d) A non-linear molecule, such as CH4 or H2O, can rotate around three axes and, again, each mode of motion contributes a term 12 kT to the internal energy. Show that Um(T) = Um(0) + 3RT CV,m = 3R (non-linear molecule, translation and rotation only) (e) Molecules do not vibrate significantly at room temperature and, as a first approximation, the contribution of molecular vibrations to the internal energy is negligible except for very large molecules such as polymers and biological macromolecules. Use the information in this problem to justify the following statement: the internal energy of a perfect gas does not change when the gas undergoes isothermal expansion. (f) Use the equipartition theorem to calculate the contribution of molecular motion to the total energy of a sample of 10.0 g of (i) argon, (ii) carbon dioxide, and (iii) methane at 20°C. Hint: For (ii) and (iii), take into account translation and rotation but not vibration. (g) We saw in part (e) that the internal energy of a perfect gas does not change when the gas undergoes isothermal expansion. What is the change in enthalpy? 1.48 It is possible to see with the aid of a powerful microscope that a

long piece of double-stranded DNA is flexible, with the distance

between the ends of the chain adopting a wide range of values. This flexibility is important because it allows DNA to adopt very compact conformations as it is packaged in a chromosome (see Chapter 11). It is convenient to visualize a long piece of DNA as a freely jointed chain, a chain of N small, rigid units of length l that are free to make any angle with respect to each other. The length l, the persistence length, is approximately 45 nm, corresponding to approximately 130 base pairs. You will now explore the work associated with extending a DNA molecule. (a) Suppose that a DNA molecule resists being extended from an equilibrium, more compact conformation with a restoring force F = −kf x, where x is the difference in the end-to-end distance of the chain from an equilibrium value and kf is the force constant. Systems showing this behavior are said to obey Hooke’s law. (i) What are the limitations of this model of the DNA molecule? (ii) Using this model, write an expression for the work that must be done to extend a DNA molecule by x. Draw a graph of your conclusion. (b) A better model of a DNA molecule is the one-dimensional freely jointed chain, in which a rigid unit of length l can make an angle of only 0° or 180° with an adjacent unit. In this case, the restoring force of a chain extended by x = nl is given by F=

kT A 1 + n D ln 2l C 1 − n F

n = n/N

where k = 1.381 × 10−23 J K−1 is Boltzmann’s constant (not a force constant). (i) What are the limitations of this model? (ii) What is the magnitude of the force that must be applied to extend a DNA molecule with N = 200 by 90 nm? (iii) Plot the restoring force against n, noting that n can be either positive or negative. How is the variation of the restoring force with end-to-end distance different from that predicted by Hooke’s law? (iv) Keeping in mind that the difference in end-to-end distance from an equilibrium value is x = nl and, consequently, dx = ldn = Nldn, write an expression for the work of extending a DNA molecule. (v) Calculate the work of extending a DNA molecule from n = 0 to n = 1.0. Hint: You must integrate the expression for w. The task can be accomplished easily with mathematical software. (c) Show that for small extensions of the chain, when n Ti, Tf /Ti > 1, which implies that the logarithm is positive, that DS > 0, and therefore that the entropy increases (Fig. 2.4). Note that the relation also shows a less obvious point, that the higher the heat capacity of the substance, the greater the change in entropy for a given rise in temperature. A moment’s thought shows this conclusion to be reasonable too: a high heat capacity implies that a lot of heat is required to produce a given change in temperature, so the ‘sneeze’ (in terms of the analogy mentioned earlier) must be more powerful than when the heat capacity is low, and the entropy increase is correspondingly high.

Calculate the change in molar entropy when water vapor is heated from 160°C to 170°C at constant volume. (CV,m = 26.92 J K−1 mol−1.) Self-test 2.1

Answer: +0.615 J K−1 mol−1

When we cannot assume that the heat capacity is constant over the temperature range of interest, which is the case for all solids at low temperatures, we have to allow for the variation of C with temperature. In Justification 2.1 we found, before making the assumption that the heat capacity is constant, that The experimental determination of the change in entropy of a sample that has a heat capacity that varies with temperature, as shown in (a), involves measuring the heat capacity over the range of temperatures of interest, then plotting C/T against T and determining the area under the curve (the tinted area shown), as shown in (b). The heat capacity of all solids decreases toward zero as the temperature is reduced. Fig. 2.5

Tf

DS =

Ti

CdT T

(2.3)

All we need to recognize is the standard result from calculus, that the integral of a function between two limits is the area under the graph of the function between the two limits. In this case, the function is C/T, the heat capacity at each temperature divided by that temperature, and it follows that DS = area under the graph of C/T plotted against T, between Ti and Tf

Experimental basis of determining an entropy change

This rule is illustrated in Fig. 2.5. To use eqn 2.3, we measure the heat capacity throughout the range of temperatures of interest. Then we divide each measurement by the corresponding temperature to get C/T at each temperature, plot these

2.2 ENTROPY AND THE SECOND LAW

75

C/T against T, and evaluate the area under the graph between the temperatures Ti and Tf . In practice, mathematical software is used to fit a curve to the variation of C with T and the integration carried out automatically.

A brief illustration

The molar heat capacity of chloroform between 20°C and 37°C was found to fit the following expression: Cp,m(T) = 91.74 J K−1 mol−1 + 0.075T J K−2 mol−1 The change in entropy over this range is therefore DSm = Sm(310 K) − Sm(293 K) =

310 K

293 K

A (91.74 J K−1 mol−1 D + 0.075 J K−2 mol−1 dT C F T

= (91.74 J K−1 mol−1)ln

310 K + (0.075 J K−2 mol−1)(310 K − 293 K) 293 K

= +6.45 J K−1 mol−1

(c) The entropy change accompanying a phase transition

We can suspect that the entropy of a substance increases when it melts and when it vaporizes because its molecules become distributed in a more disorderly way as it changes from solid to liquid and from liquid to vapor. Likewise, we expect the unfolding of a protein from a compact, active three-dimensional conformation to a more flexible conformation, a process discussed in In the laboratory 1.1, to be accompanied by an increase of entropy because the secondary structure of the polypeptide chain is lost. The transfer of energy as heat occurs reversibly when a solid is at its melting temperature. If the temperature of the surroundings is infinitesimally lower than that of the system, then energy flows out of the system as heat and the substance freezes. If the temperature is infinitesimally higher, then energy flows into the system as heat and the substance melts. Moreover, because the transition occurs at constant pressure, we can identify the energy transferred by heating per mole of substance with the enthalpy of fusion (melting). Therefore, the entropy of fusion, DfusS, the change of entropy per mole of substance, at the melting temperature, Tfus, is At the melting temperature: D fusS =

D fusH(Tfus) Tfus

Entropy of fusion

(2.4)

Notice how we must use the enthalpy of fusion at the melting temperature and that this expression applies only at the melting temperature. We get the standard entropy of fusion, D fusS 3, if the solid and liquid are both at 1 bar; we use the melting temperature at 1 bar and the corresponding standard enthalpy of fusion at that temperature. All enthalpies of fusion are positive (melting is endothermic: it requires heat), so all entropies of fusion are positive too: disorder increases on melting. The entropy of water, for example, increases when it melts because the orderly structure of ice collapses as the liquid forms (Fig. 2.6).

Fig. 2.6 (a) When a solid, here a highly stylized version of ice, melts, the molecules form a liquid. (b) As a result, the entropy of the sample increases.

76

2 THE SECOND LAW

A brief illustration

The protein lysozyme, an enzyme that breaks down bacterial cell walls, unfolds at a transition temperature of 75.5°C, and the standard enthalpy of transition as determined using differential scanning calorimetry is +509 kJ mol−1. It follows that DtrsS 3 =

D trsH 3(Ttrs) +509 kJ mol−1 = +1.46 kJ K−1 mol−1 = Ttrs (273.15 + 75.5) K

At the molecular level, the positive entropy change can be explained by the dispersal of matter and energy that accompanies the unraveling of the compact three-dimensional structure of lysozyme into a long, flexible chain that can adopt many different conformations as it writhes about in solution. Calculate the standard entropy of fusion of ice at 0°C from the information in Table 1.2. Self-test 2.2 Table 2.1 Entropies of vaporization at 1 atm and the normal boiling point

Substance

D vapS/ (J K−1 mol−1)

Ammonia, NH3

97.4

Benzene, C6H6

87.2

Bromine, Br2

88.6

Carbon tetrachloride, CCl4

85.9

Cyclohexane, C6H12 Ethanol, CH3CH2OH Hydrogen sulfide, H2S Water, H2O

85.1 109.7 87.9 109.1

Answer: +22 J K−1 mol−1

The entropy of other types of transition may be discussed similarly. Thus, the entropy of vaporization, DvapS, at the boiling temperature, Tb, of a liquid is related to its enthalpy of vaporization at that temperature by At the boiling temperature: DvapS =

DvapH(Tb) Tb

Enthalpy of vaporization

(2.5)

Note that to use this formula we use the enthalpy of vaporization at the boiling temperature. Table 2.1 lists the entropy of vaporization of several substances at 1 atm. For the standard value, DvapS, we use data corresponding to 1 bar. Because vaporization is endothermic for all substances (with one exception of little relevance to biology: helium), all entropies of vaporization are positive. The increase in entropy accompanying vaporization is in line with what we should expect when a compact liquid turns into a gas. To calculate the entropy of phase transition at a temperature other than the transition temperature, we have to do additional calculations, as shown in the following brief illustration.

A brief illustration

Suppose we want to calculate the entropy of vaporization of water at 25°C. We need to perform three calculations (Fig. 2.7). First, we calculate the entropy change for heating liquid water from 25°C to 100°C (using eqn 2.2 with data for the liquid from Table 1.1): DS1 = Cp,m(H2O, liquid) ln

The cycle of steps used to calculate the entropy of transition at a temperature other than the transition temperature. Fig. 2.7

Tf 373 K = (75.29 J K−1 mol−1) × ln Ti 298 K −1 −1 = +16.9 J K mol

Then, we use eqn 2.5 and data from Table 1.2 to calculate the entropy of transition at 100°C: DS2 =

DvapH(Tb) 4.07 × 104 J mol−1 = = +109 J K−1 mol−1 Tb 373 K

2.3 ABSOLUTE ENTROPIES AND THE THIRD LAW OF THERMODYNAMICS

Finally, we calculate the change in entropy for cooling the vapor from 100°C to 25°C (using eqn 2.2 again, but now with data for the vapor from Table 1.1): DS3 = Cp,m(H2O, vapor) ln

Tf 298 K = (33.58 J K−1 mol−1) × ln 373 K Ti −1 −1 = −7.5 J K mol

The sum of the three entropy changes is the entropy of transition at 25°C: DvapS(298 K) = DS1 + DS2 + DS3 = +118 J K−1 mol−1

(d) Entropy changes in the surroundings

We can use the definition of entropy in eqn 2.1 to calculate the entropy change of the surroundings in contact with the system at the temperature T: DSsur = qsur,rev /T. However, surroundings are so extensive that the spread of heat through them is effectively reversible, so the ‘rev’ subscript can be dropped and we can write DSsur = qsur /T. Moreover, the heat entering the surroundings is lost from the system, so qsur = −q. (For instance, if q = +100 J, an influx of 100 J into the system, then qsur = −100 J, indicating that the surroundings have lost that 100 J.) Therefore, at this stage we can write DSsur = −q/T. Finally, if the change in the system is taking place at constant pressure, we can identify q with the change of enthalpy DH, and so obtain for a process at constant pressure: DSsur = −

DH T

Entropy change of the surroundings

(2.6)

This enormously important expression will lie at the heart of our discussion of bioenergetics and the structural consequences of the Second Law. We see that it is consistent with common sense: if the process is exothermic, DH is negative and therefore DSsur is positive. The entropy of the surroundings increases if heat is released into them. If the process is endothermic (DH > 0), then the entropy of the surroundings decreases.

A brief illustration

The enthalpy of vaporization of water at 20°C is 44 kJ mol−1. When 10 cm3 of water (corresponding to 10 g or 0.55 mol H2O) in an open vessel evaporates at that temperature, the change in entropy of the surroundings is DSsur = −

(0.55 mol) × (44 kJ mol−1) = −83 J K−1 293 K

The entropy of the surroundings decreases because heat flows out of them into the water.

2.3 Absolute entropies and the Third Law of thermodynamics To calculate the entropy changes associated with biological processes, we need to see how to compile tables that list the values of the entropies of substances.

The graphical procedure summarized by Fig. 2.5 and eqn 2.3 for the determination of the difference in entropy of a substance at two temperatures has a very

77

78

2 THE SECOND LAW

The molar entropies of monoclinic and rhombic sulfur vary with temperature as shown here. Initially we do not know their values at T = 0. When we slide the two curves together by matching their separation to the measured entropy of transition at the transition temperature, we find that the entropies of the two forms are the same at T = 0.

Fig. 2.8

important application. If Ti = 0, then the area under the graph between T = 0 and some temperature T gives us the value of DS = S(T) − S(0). However, at T = 0, all the motion of the atoms has been eliminated and there is no thermal disorder. Moreover, if the substance is perfectly crystalline, with every atom in a welldefined location, then there is no spatial disorder either. We can therefore suspect that at T = 0, the entropy is zero. The thermodynamic evidence for this conclusion is based on observations like the following. Sulfur undergoes a phase transition from its rhombic form to its monoclinic polymorph at 96°C (369 K) and the enthalpy of transition is +402 J mol−1. The entropy of transition is therefore +1.09 J K−1 mol−1 at this temperature. We can also measure the molar entropy of each phase relative to its value at T = 0 by determining the heat capacity from T = 0 up to the transition temperature (Fig. 2.8). At this stage, we do not know the values of the entropies at T = 0. However, as we see from the illustration, to match the observed entropy of transition at 369 K, the molar entropies of the two crystalline forms must be the same at T = 0. We cannot say that the entropies are zero at T = 0, but from the experimental data we do know that they are the same. This observation is generalized into the Third Law of thermodynamics: The entropies of all perfectly crystalline substances are the same at T = 0. The Third Law

For convenience (and in accordance with our understanding of entropy as a measure of disorder), we take this common value to be zero. Then, with this convention, according to the Third Law, S(0) = 0 for all perfectly ordered crystalline materials.

Fig. 2.9 The absolute entropy (or Third-Law entropy) of a substance is calculated by extending the measurement of heat capacities down to T = 0 (or as close to that value as possible) and then determining the area of the graph of C/T against T up to the temperature of interest. The area is equal to the absolute entropy at the temperature T.

The Third-Law entropy, which is commonly called simply ‘the entropy’, at any temperature, S(T), is based on setting S(0) = 0. The entropy of a substance depends on the pressure; we therefore select a standard pressure (1 bar) and report the standard molar entropy, S m3 , the molar entropy of a substance in its standard state at the temperature of interest. Some values at 298.15 K (the conventional temperature for reporting data) are given in Table 2.2. It is worth taking a moment to look at the values in Table 2.2 to see that they are consistent with our understanding. All standard molar entropies are positive because raising the temperature of a sample above T = 0 invariably increases its entropy above the value S(0) = 0 because there is more thermal disorder. Another feature that we can understand in terms of disorder is illustrated by the standard molar entropy of diamond (2.4 J K−1 mol−1), which is lower than that of graphite (5.7 J K−1 mol−1). This difference is consistent with the atoms being linked less rigidly in graphite than in diamond and their thermal motion being correspondingly greater. The standard molar entropies of ice, water, and water vapor at 25°C are, respectively, 45, 70, and 189 J K−1 mol−1, and the increase in values corresponds to the increasing molecular disorder on going from a solid to a liquid and then to a gas. In the laboratory 2.1

The measurement of entropies

The Third-Law entropy at any temperature, S(T), is equal to the area under the graph of C/T between T = 0 and the temperature T (Fig. 2.9). If there are any phase transitions (for example, melting) in that range, then the entropy of each transition at the transition temperature is calculated like that in eqn 2.4 and

2.3 ABSOLUTE ENTROPIES AND THE THIRD LAW OF THERMODYNAMICS

Table 2.2

79

Standard molar entropies of some substances at 298.15 K* Sm9 /(J K−1 mol−1)

Substance Gases Ammonia, NH3

192.5

Carbon dioxide, CO2

213.7

Hydrogen, H2

130.7

Nitrogen, N2

191.6

Oxygen, O2

205.1

Water vapor, H2O

188.8

Liquids Acetic acid, CH3COOH

159.8

Ethanol, CH3CH2OH

160.7

Water, H2O

69.9

Solids Calcium carbonate, CaCO3

92.9

Diamond, C

2.4

Glycine, CH2(NH2)COOH

103.5

Graphite, C

5.7

Sodium chloride, NaCl

72.1

Sucrose, C12H22O11

360.2

Urea, CO(NH2)2

104.60

*See the Resource section for more values.

its contribution added to the contributions from each of the phases, as shown in Fig. 2.10. The entropies of gas-phase species may also be calculated from spectroscopic data about bond lengths and angles using the techniques of statistical thermodynamics, but few biologically interesting substances can be treated in this way. To implement the calorimetric procedure the heat capacity of the substance is measured (for instance, by using a differential scanning calorimeter (DSC)) down to as low a temperature as feasible and then using eqn 2.3. In practice, a polynomial in T is fitted to the experimental data and then Cp/T is integrated from the lowest temperature attainable up to the temperature of interest. Thus, if the function Cp(T) = a + bT + cT 2 + · · · is fitted (for instance, by using a leastsquares procedure in a software package) to the data between Tlowest and Ttrs, where Ttrs is the temperature of a phase transition, the entropy just before the phase transition is T trs

S(Ttrs) = S(Tlowest) +

T lowest

Cp(T) dT T

Then another polynomial is fitted to the heat capacities for the new phase up to the temperature of interest (or the next phase transition and a similar integral is evaluated). At each phase transition the enthalpy of transition is measured (once again, typically with a DSC), the entropy of transition is calculated as D trsH(Ttrs)/Ttrs by analogy with eqn 2.4, and this value is added to the value calculated by integrating the heat capacity.

The determination of entropy from heat capacity data. (a) Variation of C/T with the temperature of the sample. (b) The entropy, which is equal to the area beneath the upper curve up to the temperature of interest plus the entropy of each phase transition between T = 0 and the temperature of interest.

Fig. 2.10

80

2 THE SECOND LAW

There remains the experimental problem of determining S(Tlowest ), the entropy at the lowest attainable temperature. If very low temperatures (within a few kelvins of T = 0) can be reached and reliable measurements of Cp made, it is possible to use an extrapolation based on the observation that many nonmetallic substances have a heat capacity that obeys the Debye T 3-law: At temperatures close to T = 0, Cp = aT 3

Debye T 3-law

(2.7a)

where a is a constant that depends on the substance and is found by fitting this equation to a series of measurements of the heat capacity close to T = 0. With a determined, the entropy at low temperatures is simply At temperatures close to T = 0, S(T) = 13 Cp(T)

Entropy at low temperatures

(2.7b)

(See Exercise 2.21.) That is, the molar entropy at the low temperature T (which can be identified as Tlowest ) is equal to one-third of the constant-pressure molar heat capacity at that temperature. Other extrapolation techniques have been developed that do not require reaching such low temperatures as those required for the Debye approximation to be reliable and are described in textbooks of laboratory procedures.

2.4 The molecular interpretation of the Second and Third Laws To gain insight into the thermodynamic properties of biological assemblies and a deeper understanding of what drives a spontaneous change, we need to develop a molecular view of entropy.

The entry point into the molecular interpretation of the Second Law of thermodynamics is Boltzmann’s insight into the manner in which molecules are distributed over their available energy levels, which we explored in Fundamentals F.3 and Section 1.2. (a) The Boltzmann formula

Boltzmann made the link between the distribution of molecules over energy levels and the entropy. He proposed that the entropy of a system is given by S = k ln W

Boltzmann formula for the entropy

(2.8)

where k is Boltzmann’s constant and W is the number of microstates, the ways in which the molecules of a system can be arranged for the same total energy. At T = 0, all the molecules must be in the lowest energy state, and there is only one way of achieving that arrangement, so W = 1 and S(0) = 0 (because ln 1 = 0), in accord with the Third Law. As the temperature is raised, more arrangements correspond to the same energy, so W increases and S rises. Suppose we raise the temperature just enough for two molecules of a 100molecule system to be able to leave their lowest energy state and occupy the first excited state. Two possible microstates are ([3,4, . . . ,100]in state 0[1,2]in state 1) and ([1,3,4 . . . 42,44, . . . . 100]in state 0[2,43]in state 1)

2.4 THE MOLECULAR INTERPRETATION OF THE SECOND AND THIRD LAWS

81

where molecules 1 and 2 are excited in the first microstate and molecules 2 and 43 are excited in the second. Each microstate lasts only for an instant and corresponds to a particular distribution of molecules over the available energy levels. These two microstates and a large number of others all correspond to the same configuration, in this case the configuration {98,2,0, . . .} we introduced in Section 1.2(c). In this case, there are W = 4950 possible microstates (that corresponds to the number of ways of choosing two molecules from 100). As we saw in Section 1.2(c), there is a dominating configuration of the system—the one corresponding to the greatest number of microstates for a given total energy—and the properties of the system are those of this most probable configuration. That configuration is the one with populations given by the Boltzmann distribution. To use Boltzmann’s formula for the entropy, we set the W that occurs in it equal to the W of this dominating configuration.

A brief illustration

Suppose that a protein molecule of 100 amino acid residues denatured into a random coil can adopt 1.0 × 1031 different conformations of the same energy. We set W = 1.0 × 1031 and calculate the entropy as S = (1.38 × 10−23 J K−1) × ln(1.0 × 1031) = 9.9 × 10−22 J K−1 The corresponding molar entropy of the protein is 600 J K−1 mol−1 (to 2 significant figures; that is, 6.0 × 102 J K−1 mol−1).

(b) The relation between thermodynamic and statistical entropy

The concept of the number of microstates makes quantitative the ill-defined qualitative concepts of ‘disorder’ and ‘the dispersal of matter and energy’ that we have used to introduce the concept of entropy: a more ‘disorderly’ distribution of energy and matter corresponds to a greater number of microstates associated with the same total energy. For instance, when a perfect gas expands, the available translational energy levels get closer together (Fig. 2.11), so it is possible to distribute the molecules over them in more ways than when the volume of the container is small and the energy levels are further apart. Therefore, as the container expands, W and therefore S increase, just as for thermodynamic entropy. The Boltzmann approach also illuminates the thermodynamic definition itself (eqn 2.1) and in particular the role of the temperature. Molecules in a system at high temperature can occupy a large number of the available energy levels, so a small additional transfer of energy as heat will lead to a relatively small change in the number of accessible energy levels. Consequently, the number of microstates does not increase appreciably and neither does the entropy of the system. In contrast, the molecules in a system at low temperature have access to far fewer energy levels (at T = 0, only the lowest level is accessible), and the transfer of the same quantity of energy by heating will increase the number of accessible energy levels and the number of microstates significantly. Hence, the change in entropy on heating will be greater when the energy is transferred to a cold body than when it is transferred to a hot body. This argument suggests that the change in entropy should be inversely proportional to the temperature at which the transfer takes place, as in eqn 2.1.

When the size of a container is increased (shown here in two dimensions), the energy levels available to the molecules inside it move closer together so more are accessible at a given temperature (as indicated by the levels colored red).

Fig. 2.11

82

2 THE SECOND LAW

(c) The residual entropy

In most cases, W = 1 at T = 0 because there is only one way of putting all the molecules into the same, lowest state. Therefore, as we have seen, S = 0 at T = 0, in accord with the Third Law of thermodynamics. In certain cases, however, W may differ from 1 at T = 0. This is the case if disorder survives down to absolute zero because there is no energy advantage in adopting a particular orientation. For instance, there may be no energy difference between the arrangements . . . AB AB AB . . . and . . . BA AB BA . . . , so W > 1 even at T = 0. If S > 0 at T = 0 we say that the substance has a residual entropy. Ice has a residual entropy of 3.4 J K−1 mol−1. It stems from the disorder in the hydrogen bonds between neighboring water molecules: a given O atom has two short O–H bonds and two long O···H bonds to its neighbors, but there is a degree of randomness in which two bonds are short and which two are long. The six possible arrangements of H atoms around a central O atom in ice. Occupied locations are indicated by black dots and unoccupied locations by a grey outline.

Fig. 2.12

A brief illustration

Consider a sample of ice of N H2O molecules. Each of the 2N H atoms can be either close to or relatively far from an O atom, resulting in 22N possible arrangements. However, of the 24 = 16 possible arrangements around a single O atom, only 6 have two short and two long bonds (Fig. 2.12) and hence are acceptable. Therefore W = 22N(166 )N = (32)N and the residual entropy is S(0) = k ln W = k ln (32)N = Nk ln 32 The molar residual entropy (replace N by NA and use NAk = R) is therefore Sm(0) = R ln 32 = 3.4 J K−1 mol−1

2.5 Entropy changes accompanying chemical reactions To move into the arena of biochemistry, where reactants are transformed into products, we need to establish procedures for using the tabulated values of absolute entropies to calculate entropy changes associated with chemical reactions; to assess the spontaneity of a biological process, we need to see how to take into account entropy changes in both the system and the surroundings.

Once again, we can sometimes use our intuition to predict the sign of the entropy change associated with a chemical reaction. When there is a net formation of a gas in a reaction, as in a combustion or the equivalent but controlled oxidations characteristic of organisms, we can usually anticipate that the entropy increases. When there is a net consumption of gas, as in the fixation of N2 by certain microorganisms, it is usually safe to predict that the entropy decreases. However, for a quantitative value of the change in entropy and to predict the sign of the change when no gases are involved, we need to do an explicit calculation. (a) Standard reaction entropies

The difference in molar entropy between the products and the reactants in their standard states is called the standard reaction entropy, D rS 3. It can be expressed in terms of the molar entropies of the substances in much the same way as we have already used for the standard reaction enthalpy: D rS 3 = ∑ nS m3 (products) − ∑ nS m3 (reactants)

The standard entropy of reaction

where the v are the stoichiometric coefficients in the chemical equation.

(2.9)

2.5 ENTROPY CHANGES ACCOMPANYING CHEMICAL REACTIONS

A brief illustration

A note on good practice

The enzyme carbonic anhydrase catalyses the hydration of CO2 gas in red blood cells: CO2(g) + H2O(l) → H2CO3(aq). We expect a negative entropy of reaction because a gas is consumed. To find the explicit value at 25°C, we use the information from the Resource section to write DrS 3 = S m3 (H2CO3, aq) − {S m3 (CO2, g) + S m3 (H2O, 1)} = (187.4 J K−1 mol−1) − {(213.74 J K−1 mol−1) + (69.91 J K−1 mol−1)} = −96.3 J K−1 mol−1 (a) Predict the sign of the entropy change associated with the complete oxidation of solid sucrose, C12H22O11(s), by O2 gas to CO2 gas and liquid H2O. (b) Calculate the standard reaction entropy at 25°C. Self-test 2.3

Answer: (a) Positive; (b) +512 J K−1 mol−1

(b) The spontaneity of chemical reactions

A process may be spontaneous even though the entropy change of the system itself is negative. Consider the binding of oxidized nicotinamide adenine dinucleotide (NAD+; Atlas N4), an important electron carrier in metabolism (Case studies 1.1 and 1.2), to the enzyme lactate dehydrogenase, which plays a role in the catabolism and anabolism of carbohydrates. Experiments show that DrS 3 = −16.8 J K−1 mol−1 for binding at 25°C and pH = 7.0. The negative sign of the entropy change is expected because the association of two reactants gives rise to a more compact structure. The reaction results in a more organized structure, yet it is spontaneous! The resolution of this apparent paradox underscores a feature of entropy that recurs throughout chemistry and biology: it is essential to consider the entropy of both the system and its surroundings when deciding whether or not a process is spontaneous. The reduction in entropy by 16.8 J K−1 mol−1 relates only to the system, the reaction mixture. To apply the Second Law correctly, we need to calculate the total entropy, the sum of the changes in the system and the surroundings that jointly compose the entire ‘isolated system’ referred to in the Second Law. It may well be the case that the entropy of the system decreases when a change takes place, but there may be a more than compensating increase in entropy of the surroundings, so that overall the entropy change is positive. The opposite may also be true: a large decrease in the entropy of the surroundings may occur when the entropy of the system increases. In that case we would be wrong to conclude from the increase in the system alone that the change is spontaneous. Whenever considering the implications of entropy, we must always consider the total change of the system and its surroundings. A brief illustration

To calculate the entropy change in the surroundings when a reaction takes place at constant pressure, we use eqn 2.6, interpreting the DH in that expression as the reaction enthalpy. For example, for the formation of the NAD+-enzyme complex discussed above, with DrH 3 = −24.2 kJ mol−1, the change in entropy of the surroundings (which are maintained at 25°C, the same temperature as the reaction mixture) is

Do not make the mistake of setting the standard molar entropies of elements equal to zero: they have non-zero values (provided T > 0), as we have already discussed.

83

84

2 THE SECOND LAW

DrSsur = −

DrH (−24.2 kJ mol−1) =− = +81.2 J K−1 mol−1 T 298 K

Now we can see that the total entropy change is positive: DrStotal = (−16.8 J K−1 mol−1) + (81.2 J K−1 mol−1) = +64.4 J K−1 mol−1 This calculation confirms that the reaction is spontaneous. In this case, the spontaneity is a result of the dispersal of energy that the reaction generates in the surroundings: the complex is dragged into existence, even though it has a lower entropy than the separated reactants, by the tendency of energy to disperse into the surroundings.

The Gibbs energy One of the problems with entropy calculations is already apparent: we have to work out two entropy changes, the change in the system and the change in the surroundings, and then consider the sign of their sum. The great American theoretician J.W. Gibbs, who laid the foundations of chemical thermodynamics toward the end of the nineteenth century, discovered how to combine the two calculations into one. The combination of the two procedures in fact turns out to be of much greater relevance than just saving a little labor, and throughout this text we shall see consequences of the procedure he developed. 2.6 Focusing on the system To simplify the discussion of the role of the total change in the entropy, we need to introduce a new state function, the Gibbs energy, which will be used extensively in our study of bioenergetics and biological structure.

The total entropy change that accompanies a process is DStotal = DS + DSsur

Total entropy change

(2.10)

where DS is the entropy change for the system; for a spontaneous change, DStotal > 0. If the process occurs at constant pressure and temperature, we can use eqn 2.6 to express the change in entropy of the surroundings in terms of the enthalpy change of the system, DH. When the resulting expression is inserted into this one, we obtain At constant temperature and pressure: DStotal = DS −

DH T

(2.11)

The great advantage of this formula is that it expresses the total entropy change of the system and its surroundings in terms of the properties of the system alone. The only restriction is that the expression is confined to changes at constant pressure and temperature. (a) The definition of the Gibbs energy

Now we take a very important step. First, we introduce the Gibbs energy, G, which is defined as3 3

The Gibbs energy is still commonly referred to by its older name, the ‘free energy’.

2.6 FOCUSING ON THE SYSTEM

G = H − TS

Definition of Gibbs energy

85

(2.12)

Because H, T, and S are state functions, G is a state function too. A change in Gibbs energy, DG, at constant temperature arises from changes in enthalpy and entropy and is At constant temperature: DG = DH − TDS

Change in G at constant T

(2.13)

By comparing eqns 2.11 and 2.13, we obtain At constant temperature and pressure: DG = −TDStotal

(2.14)

We see that at constant temperature and pressure, the change in the Gibbs energy of a system is proportional to the overall change in the entropy of the system plus its surroundings. (b) Spontaneity and the Gibbs energy

The difference in sign between DG and DStotal in eqn 2.14 implies that the condition for a process being spontaneous changes from DStotal > 0 in terms of the total entropy (which is universally true) to DG < 0 in terms of the Gibbs energy (for processes occurring at constant temperature and pressure). That is, in a spontaneous change at constant temperature and pressure, the Gibbs energy decreases (Fig. 2.13). It may seem more natural to think of a system as falling to a lower value of some property. However, it must never be forgotten that to say that a system tends to fall toward lower Gibbs energy is only a modified way of saying that a system and its surroundings jointly tend toward a greater total entropy. The only criterion of spontaneous change is the total entropy of the system and its surroundings; the Gibbs energy merely contrives a way of expressing that total change in terms of the properties of the system alone and is valid only for processes that occur at constant temperature and pressure. Case study 2.1

Life and the Second Law

Every chemical reaction that is spontaneous under conditions of constant temperature and pressure, including those that drive the processes of growth, learning, and reproduction, is a reaction that proceeds in the direction of lower Gibbs energy, or—another way of expressing the same thing—results in the overall entropy of the system and its surroundings becoming greater. With these ideas in mind, it is easy to explain why life, which can be regarded as a collection of biological processes, proceeds in accord with the Second Law of thermodynamics. It is not difficult to imagine conditions in the cell that may render spontaneous many of the reactions of catabolism described briefly in Case study 1.1. After all, the breakdown of large molecules, such as sugars and lipids, into smaller molecules leads to the dispersal of matter in the cell. Energy is also dispersed, as it is released on reorganization of bonds in foods when they are oxidized. More difficult to rationalize is life’s requirement of the organization of a very large number of molecules into biological cells, which in turn assemble into organisms. To be sure, the entropy of the system—the organism—is very low because matter becomes less dispersed when molecules assemble to form cells, tissues, organs, and so on. However, the lowering of the system’s entropy comes at the expense of an increase in the entropy of the surroundings.

The criterion of spontaneous change is the increase in total entropy of the system and its surroundings. Provided we accept the limitation of working at constant pressure and temperature, we can focus entirely on the properties of the system and express the criterion as a tendency to move to lower Gibbs energy.

Fig. 2.13

86

2 THE SECOND LAW

To understand this point, recall from Case study 1.1 and Case study 1.2 that cells grow by converting energy from the Sun or oxidation of foods partially into work. The remaining energy is released as heat into the surroundings, so qsur > 0 and DSsur > 0. As with any process, life is spontaneous and organisms thrive as long as the increase in the entropy of the organism’s environment compensates for decreases in the entropy arising from the assembly of the organism. Alternatively, we may say that DG < 0 for the overall sum of physical and chemical changes that we call life.

2.7 The hydrophobic interaction To gain insight into the thermodynamic factors that contribute to the spontaneous assembly of biological macromolecules, we need to examine in detail some of the interactions that bring molecular building blocks together.

When a hydrophobic molecule (in shades of gray) is surrounded by water, the H2O molecules (with their oxygen atoms shown in red) form a cage, of which a cross-section is shown here. As a result of this acquisition of structure, the entropy of water decreases, so the dispersal of the hydrophobic molecule into the water is entropy opposed; its coalescence is entropy favored. Fig. 2.14

Throughout the text we shall see how concepts of physical chemistry can be used to establish some of the known ‘rules’ for the assembly of complex biological structures. Here, we describe how the Second Law can account for the formation of such organized assemblies as proteins and biological cell membranes. As remarked in the Prologue, we do not know all the rules that govern the folding of proteins into well-defined three-dimensional structures. However, a number of general conclusions from experimental studies give some insight into the origin of tertiary and quaternary structure in proteins. Here we focus on the observation that, in an aqueous environment (including the interior of biological cells), the chains of a protein fold in such a way as to place hydrophobic groups (water-repelling, non-polar groups such as –CH2CH(CH3)2) in the interior, which is often not very accessible to solvent, and hydrophilic groups (water-loving, polar or charged groups such as –NH3+) on the surface, which is in direct contact with the polar solvent. A species with both hydrophobic and hydrophilic regions is called amphipathic.4 Phospholipids also are amphipathic molecules that can group together to form bilayer structures and cell membranes (recall Fig. F.1). To understand the process in more detail, imagine a hypothetical initial state in which a polypeptide chain is immersed in water and has not acquired its final structure. Each hydrophobic group is surrounded by a cage of water molecules (Fig. 2.14). Now consider the actual final state in which hydrophobic groups are clustered together. Although the clustering together results in a negative contribution to the change in entropy of the system (the solution), fewer (albeit larger) cages are required and more solvent molecules are free to move. The net effect of the formation of clusters of hydrophobic groups is then a decrease in the organization of the solvent and a net increase in entropy of the system. This increase in entropy of the solvent is large enough to result in the association of hydrophobic groups in an aqueous environment being spontaneous. The process that drives the spontaneous clustering of hydrophobic groups in the presence of water is called the hydrophobic interaction.

4 The amphi- part of the name is from the Greek word for ‘both’ and the -pathic part is from the same root (meaning ‘feeling’) as sympathetic.

2.7 THE HYDROPHOBIC INTERACTION

Self-test 2.4 Two long-chain hydrophobic polypeptides can associate endto-end so that only the ends meet or side-by-side so that the entire chains are in contact. Which arrangement would produce a larger entropy change when they come together?

Answer: The side-by-side arrangement

To understand the hydrophobic interaction more completely we need to know more about the energetics of the interaction of hydrophobic groups and water. Experiments indicate that the dissolution of a largely hydrophobic molecule in water is commonly endothermic (DdissH > 0) but that the entropy change is positive (DdissS > 0): D dissG 3/kJ mol−1 D diss H 3/kJ mol−1 D dissS 3/J K−1 mol−1 CH3CH2CH2CH2OH −10 +9 +65 CH3CH2CH2CH2CH2OH −13 +8 +72

The positive entropy of dissolving is consistent with the solvent water becoming more disorganized in the presence of the hydrophobic tails of the alkanol molecules, as the hydrophobic interaction requires. These experimental values are consistent with a general rule that each additional –CH2– group contributes a further −3 kJ mol−1 to the Gibbs energy of dissolving. An important consequence of this analysis is that low temperatures disfavor the hydrophobic interaction. Thus, from DG = DH − TDS, lowering the temperature reduces the effect of DS and the DG can change from negative to positive. This is the reason why some proteins and viruses dissociate into their individual subunits as the temperature is lowered to 0°C. A further aspect of this discussion is that we can set up a scale of hydrophobicities. The hydrophobicity of a small molecular group R is reported by defining the hydrophobicity constant, p, as p = log

s(RX) s(HX)

Definition of hydrophobicity constant

(2.15)

where s(RX) is the ratio of the molar solubility (the maximum chemical amount that can be dissolved to form 1 dm3 of solution) of the compound RX in octan-1-ol, a non-polar solvent, to that in water, and s(HX) is the ratio of the molar solubility of the compound HX in octan-1-ol to that in water. Therefore, positive values of p indicate hydrophobicity and negative values indicate hydrophilicity, the thermodynamic preference for water as a solvent. It is observed experimentally that the p values of most groups do not depend on the nature of X. However, measurements do suggest group additivity of p values: –R

–CH3

–CH2CH3

–(CH2)2CH3

–(CH2)3CH3

–(CH2)4CH3

p

0.5

1

1.5

2

2.5

We see that acyclic saturated hydrocarbons become more hydrophobic as the carbon chain length increases. This trend can be rationalized by D diss H becoming more positive and D diss S more negative as the number of carbon atoms in the chain increases.

87

88

2 THE SECOND LAW

2.8 Work and the Gibbs energy change To understand how biochemical reactions can be used to release energy as work in the cell, we need to gain deeper insight into the Gibbs energy.

An important feature of the Gibbs energy is that the value of DG for a process gives the maximum non-expansion work that can be extracted from the process at constant temperature and pressure. By non-expansion work, wnon-exp , we mean any work other than that arising from the expansion of the system. It may include electrical work, if the process takes place inside an electrochemical or biological cell, or other kinds of mechanical work, such as the winding of a spring or the contraction of a muscle. As we show in the following Justification, At constant temperature and pressure: DG = wmax,non-exp Gibbs energy and non-expansion work

(2.16)

Justification 2.3 Maximum non-expansion work

We need to consider infinitesimal changes because dealing with reversible processes is then much easier. Our aim is to derive the relation between the infinitesimal change in Gibbs energy, dG, accompanying a process and the maximum amount of non-expansion work that the process can do, dwnon-exp. We start with the infinitesimal form of eqn 2.13, at constant temperature: dG = dH − TdS where, as usual, d denotes an infinitesimal difference. A good rule in the manipulation of thermodynamic expressions is to feed in definitions of the terms that appear. We do this twice. First, we use the expression for the change in enthalpy at constant pressure (eqn 1.11b, written as dH = dU + pdV ) and obtain at constant temperature and pressure: dG = dU + pdV − TdS Then we replace dU in terms of infinitesimal contributions from work and heat (dU = dw + dq): dG = dw + dq + pdV − TdS The work done on the system consists of expansion work, −pexdV, and nonexpansion work, dwnon-exp. Therefore, dG = −pexdV + dwnon-exp+ dq + pdV − TdS This derivation is valid for any process taking place at constant temperature and pressure. Now we specialize to a reversible change. For expansion work to be reversible, we need to match p and pex, in which case the first and fourth terms on the right cancel. Moreover, because the transfer of energy as heat is also reversible, we can replace dq by TdS, in which case the third and fifth terms also cancel. We are left with at constant temperature and pressure, for a reversible process: dG = dwnon-exp,rev Maximum work is done during a reversible change (Section 1.3(c)), so another way of writing this expression is

2.8 WORK AND THE GIBBS ENERGY CHANGE at constant temperature and pressure: dG = dwmax,non-exp Because this relation holds for each infinitesimal step between the specified initial and final states, it applies to the overall change too. Therefore, we obtain eqn 2.16.

Example 2.1

Estimating a change in Gibbs energy for a metabolic process

Suppose a certain small bird has a mass of 30 g. What is the minimum mass of glucose that it must consume to fly up to a branch 10 m above the ground? The change in Gibbs energy that accompanies the oxidation of 1.0 mol C6H12O6(s) to carbon dioxide gas and liquid water at 25°C is −2808 kJ. Strategy First, we need to calculate the work needed to raise a mass m through

a height h on the surface of the Earth. As we saw in eqn 1.2, this work is equal to mgh, where g is the acceleration of free fall. This work, which is nonexpansion work, can be identified with DG. We need to determine the amount of substance that corresponds to the required change in Gibbs energy and then convert that amount to a mass by using the molar mass of glucose. Solution The non-expansion work to be done is

wnon-exp = (30 × 10−3 kg) × (9.81 m s−2) × (10 m) = 3.0 × 9.81 × 1.0 × 10−1 J (because 1 kg m2 s−2 = 1 J). The amount, n, of glucose molecules required for oxidation to give a change in Gibbs energy of this value given that 1 mol provides 2808 kJ is n=

3.0 × 9.81 × 1.0 × 10−1 J 3.0 × 9.81 × 1.0 × 10−7 = mol 2.808 × 106 J mol−1 2.808

Therefore, because the molar mass, M, of glucose is 180 g mol−1, the mass, m, of glucose that must be oxidized is m = nM =

A 3.0 × 9.81 × 1.0 × 10−7 D mol × (180 g mol−1) = 1.9 × 10−4 g C F 2.808

That is, the bird must consume at least 0.19 mg of glucose for the mechanical effort (and more if it thinks about it). A hardworking human brain, perhaps one that is grappling with physical chemistry, operates at about 25 J s−1. What mass of glucose must be consumed to sustain that metabolic rate for an hour?

Self-test 2.5

Answer: 5.8 g

The great importance of the Gibbs energy in chemistry is becoming apparent. At this stage, we see that it is a measure of the non-expansion work resources of chemical reactions: if we know DG, then we know the maximum non-expansion work that we can obtain by harnessing the reaction in some way. In some cases, the non-expansion work is extracted as electrical energy. This is the case when electrons are transferred across cell membranes in some key reactions of photosynthesis and respiration (see Sections 5.10 and 5.11).

89

90

2 THE SECOND LAW

Some insight into the physical significance of G itself comes from its definition as H − TS. The enthalpy is a measure of the energy that can be obtained from the system as heat. The term TS is a measure of the quantity of energy stored in the random motion of the molecules making up the sample. Work, as we have seen, is energy transferred in an orderly way, so we cannot expect to obtain work from the energy stored randomly. The difference between the total stored energy and the energy stored randomly, H − TS, is available for doing work, and we recognize that difference as the Gibbs energy. In other words, the Gibbs energy is the energy stored in the uniform motion and arrangement of the molecules in the system.

Case study 2.2

The action of adenosine triphosphate

In biological cells, the energy released by the oxidation of foods (Case study 1.1) is stored in adenosine triphosphate (ATP or ATP4−, Atlas N3). The essence of ATP’s action is its ability to lose its terminal phosphate group by hydrolysis and to form adenosine diphosphate (ADP or ADP3−, Atlas N2): + ATP4−(aq) + H2O(l) → ADP3−(aq) + HPO 2− 4 (aq) + H3O (aq)

At pH = 7.0 and 37°C (310 K, blood temperature) the enthalpy and Gibbs energy of hydrolysis are DrH = −20 kJ mol−1 and DrG = −31 kJ mol−1, respectively. Under these conditions, the hydrolysis of 1 mol ATP4 −(aq) results in the extraction of up to 31 kJ of energy that can be used to do non-expansion work, such as the synthesis of proteins from amino acids, muscular contraction, and the activation of neuronal circuits in our brains, as we shall see in Chapter 5. If no attempt is made to extract any energy as work, then 20 kJ (in general, DH) of heat will be produced.

Checklist of key concepts 1. A spontaneous change is a change that has a tendency to occur without work having to be done to bring it about.

6. The Boltzmann formula expresses the statistical entropy in terms of the number of microstates of a system.

2. Matter and energy tend to disperse.

7. The Gibbs energy is defined as G = H − TS and is a state function.

3. The Second Law states that the entropy of an isolated system tends to increase. 4. In general, the entropy change accompanying the heating of a system is equal to the area under the graph of C/T against T between the two temperatures of interest. 5. The Third Law of thermodynamics states that the entropies of all perfectly crystalline substances are the same at T = 0 (and may be taken to be zero).

8. At constant temperature and pressure, a system tends to change in the direction of decreasing Gibbs energy. 9. The hydrophobic interaction is a process that leads to the organization of solute molecules and is driven by a tendency toward greater dispersal of solvent molecules. 10. At constant temperature and pressure, the change in Gibbs energy accompanying a process is equal to the maximum non-expansion work the process can do.

EXERCISES

91

Checklist of key equations Property or process

Equation

Comment

Entropy change

DS = qrev /T

Definition

Entropy change of surroundings

DSsur = −q/T

q is heat supplied to the system

DSsur = −DH/T

DH is the enthalpy change of the system; constant pressure process

Boltzmann formula for the entropy

S = k ln W

W is the number of microstates

Entropy of transition

D trsS (Ttrs) = D trsH(Ttrs )/Ttrs

At the transition temperature

Entropy change due to a change in temperature

DS = C ln(T2/T1)

Heat capacity constant in the range of interest

Standard reaction entropy

Dr S 3 = ∑nSm3(products) − ∑ nSm3(reactants)

Definition

Gibbs energy

G = H − TS

Definition

Change in Gibbs energy

DG = DH − TDS

At constant temperature

Relation to maximum non-expansion work

DG = wmax,non-exp

At constant temperature and pressure

Discussion questions 2.1 The following expressions have been used to establish criteria for spontaneous change: DSisolated system > 0 and DG < 0. Discuss the origin, significance, and applicability of each criterion.

2.5 Explain the origin of the residual entropy. 2.6 Without performing a calculation, predict whether the standard entropies of the following reactions are positive or negative:

2.2 Explain the limitations of the following expressions: (a) DS = C ln(Tf /Ti ), (b) DG = DH − TDS, and (c) DG = wmax,non-exp.

(a) Ala–Ser–Thr–Lys–Gly–Arg–Ser ffg Ala–Ser–Thr–Lys– + Gly–Arg

2.3 Suggest a procedure for the measurement of the entropy of

(b) N2(g) + 3 H2(g) ffg 2 NH3(g)

unfolding of a protein with differential scanning calorimetry (see In the laboratory 1.1).

(c) ATP4−(aq) + H2O(1) ffg ADP3−(aq) + HPO2− 4 (aq) + H3O+(aq)

2.4 Justify the identification of the statistical entropy with the thermodynamic entropy.

trypsin

trypsin

trypsin

2.7 Provide a molecular interpretation of the hydrophobic interaction.

Exercises 2.8 A goldfish swims in a bowl of water at 20°C. Over a period of time, the fish transfers 120 J to the water as a result of its metabolism. What is the change in entropy of the water?

energy at a constant rate, and sketch a graph showing (a) the change in temperature of the system, (b) the enthalpy of the system, and (c) the entropy of the system as a function of time.

2.9 Suppose that when you exercise, you consume 100 g of glucose and that all the energy released as heat remains in your body at 37°C. What is the change in entropy of your body?

2.12 What is the change in entropy of 100 g of water when it is heated

2.10 Suppose you put a cube of ice of mass 100 g into a glass of water at

2.13 Estimate the molar entropy of potassium chloride at 5.0 K given

just above 0°C. When the ice melts, about 33 kJ of energy is absorbed from the surroundings as heat. What is the change in entropy of (a) the sample (the ice) and (b) the surroundings (the glass of water)?

that its molar heat capacity at that temperature is 1.2 mJ K−1 mol−1.

2.11 Calculate the change in entropy of 100 g of ice at 0°C as it is

melted, heated to 100°C, and then vaporized at that temperature. Suppose that the changes are brought about by a heater that supplies

from room temperature (20°C) to body temperature (37°C)? Use Cp,m = 75.5 J K−1 mol−1.

2.14 Equation 2.2 is based on the assumption that the heat capacity is

independent of temperature. Suppose, instead, that the heat capacity depends on temperature as C = a + bT + a/T 2. Find an expression for the change of entropy accompanying heating from Ti to Tf . Hint: See Justification 2.1.

92

2 THE SECOND LAW

2.15 Calculate the change in entropy when 100 g of water at 80°C is

2.23 Calculate the standard reaction entropy at 298 K

poured into 100 g of water at 10°C in an insulated vessel given that Cp,m = 75.5 J K−1 mol−1.

of the fermentation of glucose to ethanol: C6H12O6(s) → 2 C2H5OH(l) + 2 CO2(g).

2.16 The protein lysozyme unfolds at a transition temperature of 75.5°C, and the standard enthalpy of transition is 509 kJ mol−1. Calculate the entropy of unfolding of lysozyme at 25.0°C, given that the difference in the constant-pressure heat capacities on unfolding is 6.28 kJ K−1 mol−1 and can be assumed to be independent of temperature. Hint: Imagine that the transition at 25.0°C occurs in three steps: (i) heating of the folded protein from 25.0°C to the transition temperature, (ii) unfolding at the transition temperature, and (iii) cooling of the unfolded protein to 25.0°C. Because the entropy is a state function, the entropy change at 25.0°C is equal to the sum of the entropy changes of the steps.

2.24 The constant-pressure molar heat capacities of linear gaseous molecules are approximately 72 R and those of non-linear gaseous molecules are approximately 4R. Estimate the change in standard reaction entropy of the following two reactions when the temperature is increased by 10 K at constant pressure:

2.17 The enthalpy of the graphite → diamond phase transition,

which under 100 kbar occurs at 2000 K, is +1.9 kJ mol . Calculate the entropy of transition at that temperature. −1

2.18 The enthalpy of vaporization of methanol is 35.27 kJ mol−1

at its normal boiling point of 64.1°C. Calculate (a) the entropy of vaporization of methanol at this temperature and (b) the entropy change of the surroundings. 2.19 Trouton’s rule summarizes the results of experiments showing

that the entropy of vaporization measured at the boiling point, DvapS = Dvap H(Tb)/Tb, is approximately the same and equal to about 85 J K−1 mol−1 for all liquids except when hydrogen bonding or some other kind of specific molecular interaction is present. (a) Provide a molecular interpretation for Trouton’s rule. (b) Estimate the entropy of vaporization and the enthalpy of vaporization of octane, which boils at 126°C. (c) Trouton’s rule does not apply to water because in the liquid, water molecules are held together by an extensive network of hydrogen bonds. Provide a molecular interpretation for the observation that Trouton’s rule underestimates the value of the entropy of vaporization of water. 2.20 Calculate the entropy of fusion of a compound at 25°C given that

its enthalpy of fusion is 32 kJ mol−1 at its melting point of 146°C and the molar heat capacities (at constant pressure) of the liquid and solid forms are 28 J K−1 mol−1 and 19 J K−1 mol−1, respectively. 2.21 Show that at temperatures close to T = 0, S(T) = Cp(T). 1 3

2.22 Calculate the residual molar entropy of a solid in which the

molecules can adopt (a) three, (b) five, and (c) six orientations of equal energy at T = 0.

(a) 2 H2(g) + O2(g) → 2 H2O(l) (b) CH4(g) + 2 O2(g) → CO2(g) + 2 H2O(g) 2.25 Use the information in Exercise 2.24 to calculate the standard

Gibbs energy of reaction of N2 + 3 H2(g) → 2 NH3(g). 2.26 In a particular biological reaction taking place in the body at

37°C, the change in enthalpy was −125 kJ mol−1 and the change in entropy was −126 J K−1 mol−1. (a) Calculate the change in Gibbs energy. (b) Is the reaction spontaneous? (c) Calculate the total change in entropy of the system and the surroundings. 2.27 The change in Gibbs energy that accompanies the oxidation of C6H12O6(s) to carbon dioxide and water vapor at 25°C is −2808 kJ mol−1. How much glucose does a person of mass 65 kg need to consume to climb through 10 m? 2.28 A non-spontaneous reaction may be driven by coupling it to

a reaction that is spontaneous. The formation of glutamine from glutamate and ammonium ions requires 14.2 kJ mol−1 of energy input. It is driven by the hydrolysis of ATP to ADP mediated by the enzyme glutamine synthetase. (a) Given that the change in Gibbs energy for the hydrolysis of ATP corresponds to DG = −31 kJ mol−1 under the conditions prevailing in a typical cell, can the hydrolysis drive the formation of glutamine? (b) How many moles of ATP must be hydrolyzed to form 1 mol glutamine? 2.29 The hydrolysis of acetyl phosphate has DG = −42 kJ mol−1 under typical biological conditions. If the phosphorylation of acetic acid were to be coupled to the hydrolysis of ATP, what is the minimum number of ATP molecules that would need to be involved? 2.30 Suppose that the radius of a typical cell is 10 mm and that inside

it 106 ATP molecules are hydrolyzed each second. What is the power density of the cell in watts per cubic meter (1 W = 1 J s−1)? A computer battery delivers about 15 W and has a volume of 100 cm3. Which has the greater power density, the cell or the battery? (For data, see Exercise 2.28.)

Projects 2.31 The following is an example of a structure–activity relation (SAR), in which it is possible to correlate the effect of a structural change in a compound with its biological function. The use of SARs can improve the design of drugs for the treatment of disease because it facilitates the prediction of the biological activity of a compound before it is synthesized. The binding of non-polar groups of amino acid to hydrophobic sites in the interior of proteins is governed largely by hydrophobic interactions.

(a) Consider a family of hydrocarbons R–H. The hydrophobicity constants, p, for R = CH3, CH2CH3, (CH2)2CH3, (CH2)3CH3, and (CH2)4CH3 are, respectively, 0.5, 1.0, 1.5, 2.0, and 2.5. Use these data to predict the p value for (CH2)6CH3.

(b) The equilibrium constants KI for the dissociation of inhibitors (1) from the enzyme chymotrypsin (Atlas P3) were measured for different substituents R:

PROJECTS

R p log KI

CH3CO −0.20 −1.73

CN −0.025 −1.90

NO2 0.33 −2.43

CH3 0.50 −2.55

Cl 0.90 −3.40

Plot log KI against p. Does the plot suggest a linear relationship? If so, what are the slope and intercept to the log KI axis of the line that best fits the data? (c) Predict the value of Ki for the case R = H. 2.32 An exergonic reaction is a reaction for which DG < 0, and an

endergonic reaction is a reaction for which DG > 0. Here we investigate the molecular basis for the observation first discussed in Case study 2.2 that the hydrolysis of ATP is exergonic at pH = 7.0 and 310 K: ATP4−(aq) + H2O(l) → ADP3−(aq) + HPO42−(aq) + H3O+(aq) DrG = −31 kJ mol−1 (a) It is thought that the exergonicity of ATP hydrolysis is due in part to the fact that the standard entropies of hydrolysis of polyphosphates are positive. Why would an increase in entropy accompany the hydrolysis of a triphosphate group into a diphosphate and a phosphate group?

93

(b) Under identical conditions, the Gibbs energies of hydrolysis of H4ATP and MgATP2−, a complex between the Mg 2+ ion and ATP4−, are less negative than the Gibbs energy of hydrolysis of ATP4−. This observation has been used to support the hypothesis that electrostatic repulsion between adjacent phosphate groups is a factor that controls the exergonicity of ATP hydrolysis. Provide a rationale for the hypothesis and discuss how the experimental evidence supports it. Do these electrostatic effects contribute to the D rH or D r S terms that determine the exergonicity of the reaction? Hint: In the MgATP2− complex, the Mg 2+ ion and ATP4− anion form two bonds: one that involves a negatively charged oxygen belonging to the terminal phosphate group of ATP4− and another that involves a negatively charged oxygen belonging to the phosphate group adjacent to the terminal phosphate group of ATP4−. (c) Stabilization due to resonance in ATP4− and the HPO42− ion is thought to be one of the factors that controls the exergonicity of ATP hydrolysis. Provide a rationale for the hypothesis. Does stabilization through resonance contribute to the Dr H or Dr S terms that determine the exergonicity of the reaction?

3 The thermodynamics of transition 3.1 3.2 3.3

3.4

The condition of stability The variation of Gibbs energy with pressure The variation of Gibbs energy with temperature Phase diagrams

Phase transitions in biopolymers and aggregates 3.5 3.6

Phase equilibria

94

94 95

98 99

106

The stability of nucleic acids and proteins 106 Phase transitions of biological membranes 108

Case study 3.1 The use of phase diagrams in the study of proteins

Boiling, freezing, the unfolding of proteins, and the unzipping of a DNA double helix are all examples of phase transitions, or changes of phase without change of chemical composition. Many phase changes are common everyday phenomena, and their description is an important part of physical chemistry. They occur whenever a solid changes into a liquid, as in the melting of ice, or a liquid changes into a vapor, as in the vaporization of water in our lungs. They also occur when one solid phase changes into another, as in the conversion of one phase of a biological membrane into another as it is heated. The thermodynamics of phase changes of pure materials is also important because it prepares us first for the study of mixtures and then for the study of chemical equilibria (Chapter 4). Some of the thermodynamic concepts developed in this chapter also form the basis of important experimental techniques in biochemistry, such as the measurement of molar masses of proteins and nucleic acids and the investigation of the binding of small molecules to proteins.

The thermodynamics of transition 109

Gas solubility and breathing Case study 3.3 The Donnan equilibrium

117

Because the Gibbs energy, G = H − TS, provides a signpost of spontaneous change when the pressure and temperature are constant, and we need to know the conditions under which a transition from one state to another becomes spontaneous, it is at the centre of all that follows. In particular, we need to know how G depends on the pressure and temperature. As we work out these dependencies, we shall acquire deep insight into the thermodynamic properties of biologically important substances and the transitions they can undergo.

119

3.1 The condition of stability

Colligative properties

122

The modification of boiling and freezing points 3.10 Osmosis

123 125

The thermodynamic description of mixtures 110 3.7 3.8

The chemical potential 110 Ideal and ideal–dilute solutions 111

Case study 3.2

3.9

In the laboratory 3.1

Osmometry

127

Checklist of key concepts Checklist of key equations Further information 3.1 The phase rule Further information 3.2 Measures of concentration Discussion questions Exercises Projects

128 129 129 130 132 132 134

To understand processes ranging from the melting of ice to the denaturation of biopolymers, we need to understand the relative thermodynamic stabilities of the phases of a substance.

First, we need to establish the importance of the molar Gibbs energy, Gm = G/n, in the discussion of phase transitions of a pure substance. The molar Gibbs energy, an intensive property, is characteristic of the phase of the substance. For instance, the molar Gibbs energy of liquid water is in general different from that of water vapor at the same temperature and pressure. When an amount n of the substance changes from phase 1 (for instance, liquid) with molar Gibbs energy Gm(1) to phase 2 (for instance, vapor) with molar Gibbs energy Gm(2), the change in Gibbs energy is

3.2 THE VARIATION OF GIBBS ENERGY WITH PRESSURE DG = nGm(2) − nGm(1) = n{Gm(2) − Gm(1)} We know that a negative value of DG indicates that the change from phase 1 to phase 2 is spontaneous at constant temperature and pressure. It follows that the change from phase 1 to phase 2 is spontaneous if the molar Gibbs energy of phase 2 is lower than that of phase 1. In other words, a substance has a spontaneous tendency to change into the phase with the lower molar Gibbs energy. If at a certain temperature and pressure the solid phase of a substance has a lower molar Gibbs energy than its liquid phase, then the solid phase is thermodynamically more stable and the liquid will (or at least has a tendency to) freeze. If the opposite is true, the liquid phase is thermodynamically more stable and the solid will melt. For example, at 1 atm, ice has a lower molar Gibbs energy than liquid water when the temperature is below 0°C, and under these conditions water converts spontaneously to ice. 3.2 The variation of Gibbs energy with pressure To discuss how phase transitions depend on the pressure and to lay the foundation for understanding the behavior of solutions of biological macromolecules, we need to know how the molar Gibbs energy varies with pressure.

Why should biologists be interested in the variation of the Gibbs energy with the pressure since in most cases their systems are at pressures close to 1 atm? You should recall the discussion in Chapter 1, where we pointed out that to study the thermodynamic properties of a liquid (in which biochemists do have an interest), we can explore the properties of a vapor which, as a gas, are easy to formulate, and then imagine bringing the vapor into equilibrium with the liquid. Then the properties of the liquid mirror those of the vapor. That is the strategy we adopt throughout this chapter. First we establish equations that apply to gases. Then we consider equilibria between gases and liquids and adapt the gas-phase expressions to describe what really interests us, the properties of liquids. We show in the following Justification that when the temperature is held constant and the pressure is changed from pi to pf , the molar Gibbs energy of an incompressible liquid becomes Gm(pf ) = Gm(pi ) + (pf − pi )Vm

Variation of the Gibbs energy with pressure (for an incompressible liquid)

(3.1)

where Vm is the molar volume of the substance. This expression is valid when the molar volume is constant in the pressure range of interest, which is true of most liquids (and solids) under normal circ*mstances. Even though gases are far from incompressible, we can also use eqn 3.1 for the qualitative discussion of the pressure dependence of Gm of a gas provided the change in pressure is small.

Justification 3.1 The variation of G of an incompressible liquid with pressure

When the temperature, volume, and pressure of a substance are changed by infinitesimal amounts, H changes to H + dH, T changes to T + dT, and S changes to S + dS. As a result, G changes to G + dG, where G + dG = (H + dH) − (T + dT)(S + dS) = H + dH − TS − TdS − SdT − dTdS The G on the left cancels the H − TS on the right, the doubly infinitesimal dTdS can be neglected, and we are left with

95

96

3 PHASE EQUILIBRIA dG = dH − TdS − SdT In a similar way, from the definition H = U + pV, letting U change to U + dU, and so on, and neglecting the doubly infinitesimal term dpdV, we can write dH = dU + pdV + Vdp At this point we need to know how the internal energy changes, and write dU = dq + dw If initially we consider only reversible changes, we can replace dq by TdS (because dS = dqrev /T) and dw by −pdV (because dw = −pexdV and pex = p for a reversible change) and obtain dU = TdS − pdV Now we substitute this expression into the expression for dH and that expression into the expression for dG and obtain dG = TdS − pdV + pdV + Vdp − TdS − SdT It follows that

A note on good practice

When confronted with a proof in thermodynamics, go back to fundamental definitions (as we did three times in succession in this derivation: first of G, then of H, and finally of U).

dG = Vdp − SdT

Variation of the Gibbs energy with pressure and temperature

(3.2)

Now here is a subtle but important point. To derive this result we have supposed that the changes in conditions have been made reversibly. However, G is a state function and so the change in its value is independent of path. Therefore, the expression is valid for any change within a system of known composition, not just a reversible change. At this point we decide to keep the temperature constant and set dT = 0; this leaves dG = Vdp and, for molar quantities, dGm = Vmdp. This expression is exact but applies only to an infinitesimal change in the pressure. For an observable change, we replace dGm by Gm(pf ) − Gm(pi) and dp by pf − pi, respectively, and obtain eqn 3.1, provided the molar volume is constant over the range of interest.

The variation of molar Gibbs energy with pressure. The region of stability of each phase is indicated in the band at the bottom of the illustration.

Fig. 3.1

Equation 3.1 tells us that, because all molar volumes are positive, the molar Gibbs energy increases (Gm(pf ) > Gm(pi)) when the pressure increases (pf > pi ). We also see that, for a given change in pressure, the resulting change in molar Gibbs energy is greatest for substances with large molar volumes. Again bearing in mind that we can apply eqn 3.1 qualitatively to gases over small changes in pressure, we see that because the molar volume of a gas is much larger than that of a condensed phase (a liquid or a solid), the dependence of Gm on p is much greater for a gas than for a condensed phase. For most substances (water is an important exception), the molar volume of the liquid phase is greater than that of the solid phase. Therefore, for most substances, the slope of a graph of Gm against p is greater for a liquid than for a solid. These characteristics are illustrated in Fig. 3.1. As we see from Fig. 3.1, when we increase the pressure on a substance, the molar Gibbs energy of the gas phase rises above that of the liquid, then the molar Gibbs energy of the liquid rises above that of the solid. Because the system has a

3.2 THE VARIATION OF GIBBS ENERGY WITH PRESSURE

Mathematical toolkit 3.1

97

Integration

The area under a graph of any function f is found by the techniques of integration. For instance, the area under the graph of the function f(x) between x = a and x = b is denoted by b

area between a and b =

冮 f(x)dx a

The elongated S symbol on the right is called the integral of the function f. When written as 2 alone, it is the indefinite integral of the function. When written with limits (as in the expression above), it becomes the definite integral of the function. The definite integral is the indefinite integral evaluated at the upper limit (b) minus the indefinite integral evaluated at the lower limit (a).

A very useful integral in physical chemistry is

冮 dxx = ln x + constant where ln x is the natural logarithm of x. To evaluate the integral between the limits x = a and x = b, we write b

dx = (ln x + constant) x a

b

a

= (ln b + constant) − (ln a + constant) = ln b − ln a = ln

b a

We see that the constant cancels. For instance, the area under the graph of 1/x lying between a = 2 and b = 3 is ln( 32 ) = 0.41.

tendency to convert into the state of lowest molar Gibbs energy, the graphs show that at low pressures the gas phase is the most stable, then at higher pressures the liquid phase becomes the most stable, followed by the solid phase. In other words, under pressure the substance condenses to a liquid, and then further pressure can result in the formation of a solid. We can use eqn 3.1 to predict the actual shape of graphs like those in Fig. 3.1. For a solid or liquid, the molar volume is almost independent of pressure, so eqn 3.1 is an excellent approximation to the change in molar Gibbs energy. It shows that the molar Gibbs energy of a solid or liquid increases linearly with pressure. However, because the molar volume of a condensed phase is so small, the dependence is very weak, and for typical ranges of pressure of interest to us, we can ignore the pressure dependence of G. The molar Gibbs energy of a gas, however, does depend on the pressure, and because the molar volume of a gas is large, the dependence is significant. To find a quantitative expression for the pressure dependence that is valid over a substantial pressure range, we have to take into account the fact that a gas is compressible and that the molar volume decreases as pressure is applied. We therefore expect the Gibbs energy to increase with pressure, but for it to become less sensitive to pressure as the pressure rises (because the molar volume is decreasing). We show in the following Justification that Gm(pf ) = Gm(pi ) + RT ln

pf pi

(3.3)

This equation shows that the molar Gibbs energy increases logarithmically (as ln p) with the pressure (Fig. 3.2). The flattening of the curve at high pressures reflects the fact that, as we anticipated, as Vm gets smaller, Gm becomes less responsive to pressure. Justification 3.2 The pressure variation of the Gibbs energy of a perfect gas

We start with the exact expression for the effect of an infinitesimal change in pressure obtained in Justification 3.1, that dGm = Vmdp. For a change in pressure

Fig. 3.2 The variation of the molar Gibbs energy of a perfect gas with pressure.

98

3 PHASE EQUILIBRIA

from pi to pf, we need to add together (integrate) all these infinitesimal changes and write pf

DGm =

冮 V dp m

pi

To evaluate the integral, we must know how the molar volume depends on the pressure. The easiest case to consider is a perfect gas, for which Vm = RT/p. Then pf

DGm =

pi

RT dp = RT p

pf

pi

dp p = RT ln f p pi

We have used the standard integral described in Mathematical toolkit 3.1. Finally, with DGm = Gm(pf ) − Gm(pi), we get eqn 3.3.

3.3 The variation of Gibbs energy with temperature To understand why phase transitions, including the denaturation of a biopolymer, occur at a specific temperature, we need to know how molar Gibbs energy varies with temperature.

For small changes in temperature, we show in the following Justification that the change in molar Gibbs energy at constant pressure may be written as Gm(Tf ) = Gm(Ti) − (Tf − Ti )Sm

Variation of the Gibbs energy with temperature

(3.4)

This expression is valid provided the entropy of the substance is unchanged over the range of temperatures of interest. Justification 3.3 The variation of the Gibbs energy with temperature

The starting point for this short derivation is eqn 3.2 in Justification 3.1, which we rewrite as the change in molar Gibbs energy when both the pressure and the temperature are changed by infinitesimal amounts: dGm = Vmdp − SmdT If we hold the pressure constant, dp = 0, and dGm = −SmdT This expression is exact. If we suppose that the molar entropy is unchanged in the range of temperatures of interest, we can replace the infinitesimal changes by observable changes and so obtain eqn 3.4. The variation of molar Gibbs energy with temperature. All molar Gibbs energies decrease with increasing temperature. The regions of temperature over which the solid, liquid, and gaseous forms of a substance have the lowest molar Gibbs energy are indicated in the band at the top of the illustration.

Fig. 3.3

Equation 3.4 tells us that, because molar entropy is positive, an increase in temperature (Tf > Ti ) results in a decrease in Gm (Gm(Tf ) < Gm(Ti )). Moreover, for a given change of temperature, the change in molar Gibbs energy is proportional to the molar entropy. For a given substance, because the molar entropy of the gas phase is greater than that for a condensed phase, the molar Gibbs energy falls more steeply with temperature for a gas than for a condensed phase. The molar entropy of the liquid phase of a substance is greater than that of its solid phase, so the slope is least steep for a solid. Figure 3.3 summarizes these characteristics.

3.4 PHASE DIAGRAMS

Figure 3.3 reveals the thermodynamic reason why substances melt and vaporize as the temperature is raised. At low temperatures, the solid phase has the lowest molar Gibbs energy and is therefore the most stable. However, as the temperature is raised, the molar Gibbs energy of the liquid phase falls below that of the solid phase and the substance melts. At even higher temperatures, the molar Gibbs energy of the vapor plunges down below that of the liquid phase, and the vapor becomes the most stable phase. In other words, above a certain temperature, the liquid vaporizes. We can also start to understand why some substances, such as solid carbon dioxide, sublime to a vapor without first forming a liquid. There is no fundamental requirement for the three lines to lie exactly in the positions we have drawn them in Fig. 3.3: the liquid line, for instance, could lie where we have drawn it in Fig. 3.4. Now we see that at no temperature (at the given pressure) does the liquid phase have the lowest molar Gibbs energy. Such a substance converts spontaneously directly from the solid to the vapor. That is, the substance sublimes. The transition temperature between two phases, such as between liquid and solid or between conformations of a protein, is the temperature, at a given pressure, at which the two phases are in equilibrium and therefore their molar Gibbs energies are equal. At 1 atm, for instance, ice and liquid water are in equilibrium at 0°C and Gm(H2O,l) = Gm(H2O,s). As always when using thermodynamic arguments, it is important to keep in mind the distinction between the spontaneity of a phase transition and its rate. Spontaneity is a tendency, not necessarily an actuality. A phase transition predicted to be spontaneous may occur so slowly as to be unimportant in practice. For instance, at normal temperatures and pressures the molar Gibbs energy of graphite is 3 kJ mol−1 lower than that of diamond, so there is a thermodynamic tendency for diamond to convert into graphite. However, for this transition to take place, the carbon atoms of diamond must change their locations, and because the bonds between the atoms are so strong and large numbers of bonds must change simultaneously, this process is immeasurably slow except at high temperatures. In gases and liquids the mobilities of the molecules normally allow phase transitions to occur rapidly, but in solids thermodynamic instability may be frozen in and a thermodynamically unstable phase may persist for thousands of years. The molecules of liquids are mobile, so this ‘metastability’ is much less likely to occur. Nevertheless, even a liquid may persist above its boiling point as a superheated liquid if it is heated carefully and there are no so-called nucleation centers, such as scratches on the interior of the containing vessel, at which the vapor can form.

99

Fig. 3.4 If the line for the Gibbs energy of the liquid phase does not cut through the line for the solid phase (at a given pressure) before the line for the gas phase cuts through the line for the solid, the liquid is not stable at any temperature at that pressure. Such a substance sublimes.

3.4 Phase diagrams To prepare for being able to describe phase transitions in biological macromolecules, first we need to explore the conditions for equilibrium between phases of simpler substances.

The phase diagram of a substance is a map showing the conditions of temperature and pressure at which its various phases are thermodynamically most stable (Fig. 3.5). For example, at point A in the illustration, the vapor phase of the substance is thermodynamically the most stable, but at C the liquid phase is the most stable. The boundaries between regions in a phase diagram, which are called phase boundaries, show the values of p and T at which the two neighboring phases are in equilibrium. For example, if the system is arranged to have a pressure and

Fig. 3.5 A typical phase diagram, showing the regions of pressure and temperature at which each phase is the most stable. The phase boundaries (three are shown here) show the values of pressure and temperature at which the two phases separated by the line are in equilibrium. The significance of the letters A, B, C, D, and E (also referred to in Fig. 3.8) is explained in the text.

100

3 PHASE EQUILIBRIA

Fig. 3.6 When a small volume of water is introduced into the vacuum above the mercury in a barometer (a), the mercury is depressed (b) by an amount that is proportional to the vapor pressure of the liquid. (c) The same pressure is observed however much liquid is present (provided some is present).

temperature represented by point B, then the liquid and its vapor are in equilibrium (like liquid water and water vapor at 1 atm and 100°C). If the temperature is reduced at constant pressure, the system moves to point C, where the liquid is stable (like water at 1 atm and at temperatures between 0°C and 100°C). If the temperature is reduced still further to D, then the solid and liquid phases are in equilibrium (like ice and water at 1 atm and 0°C). A further reduction in temperature to E takes the system into the region where the solid is the stable phase. Any point lying on a phase boundary represents a pressure and temperature at which there is a ‘dynamic equilibrium’ between the two adjacent phases. A state of dynamic equilibrium is one in which a reverse process is taking place at the same rate as the forward process. Although there may be a great deal of activity at a molecular level, there is no net change in the bulk properties or appearance of the sample. For example, any point on the liquid–vapor boundary represents a state of dynamic equilibrium in which vaporization and condensation continue at matching rates. Molecules are leaving the surface of the liquid at a certain rate, and molecules already in the gas phase are returning to the liquid at the same rate; as a result, there is no net change in the number of molecules in the vapor and hence no net change in its pressure. Similarly, a point on the solid–liquid curve represents conditions of pressure and temperature at which molecules are ceaselessly breaking away from the surface of the solid and contributing to the liquid. However, they are doing so at a rate that exactly matches that at which molecules already in the liquid are settling onto the surface of the solid and contributing to the solid phase. (a) Phase boundaries

The pressure of a vapor that is in equilibrium with its condensed phase is called the vapor pressure of the substance. Vapor pressure increases with temperature because, as the temperature is raised, more molecules have sufficient energy to leave their neighbors in the liquid. To determine the vapor pressure, a small amount of liquid can be introduced into the near-vacuum at the top of a mercury barometer and the depression of the column measured (Fig. 3.6). To ensure that the pressure exerted by the vapor is truly the vapor pressure, enough liquid must be added for some to remain after the vapor forms, for only then are the liquid and vapor phases in equilibrium. The temperature can be changed to determine another point on the curve, and so on (Fig. 3.7). The plot of the vapor pressure against temperature is also the liquid–vapor boundary in a phase diagram. To appreciate that interpretation, suppose we have a liquid in a cylinder fitted with a piston. If at some temperature we apply a pressure greater than the vapor pressure of the liquid, the vapor is eliminated, the piston rests on the surface of the liquid, and the system moves to one of the points in the ‘liquid’ region of the phase diagram. If instead we reduce the pressure on the system to a value below the vapor pressure at that temperature, the system moves to one of the points in the ‘vapor’ region of the diagram. At the vapor pressure itself, vapor and liquid are in equilibrium, and the state of the system is represented by a point on the phase boundary. What would be observed when a pressure of 50 Torr is applied to a sample of water in equilibrium with its vapor at 25°C, when its vapor pressure is 23.8 Torr? Self-test 3.1

Fig. 3.7 The experimental variation of the vapor pressure of water with temperature.

Answer: The sample condenses entirely to liquid.

3.4 PHASE DIAGRAMS

101

The same approach can be used to plot the solid–vapor boundary, which is a graph of the vapor pressure of the solid against temperature. The sublimation vapor pressure of a solid, the pressure of the vapor in equilibrium with a solid at a particular temperature, is usually much lower than that of a liquid because the molecules are more strongly bound together in the solid than in the liquid. A more sophisticated procedure is needed to determine the locations of solid– solid phase boundaries like that between the different forms of ice, for instance, because the transition between two solid phases is more difficult to detect. One approach is to use thermal analysis, which takes advantage of the heat released during a transition. In a typical thermal analysis experiment, a sample is allowed to cool and its temperature is monitored. When the transition occurs, energy is released as heat and the cooling stops until the transition is complete (Fig. 3.8). The transition temperature is obvious from the shape of the graph and is used to mark a point on the phase diagram. The pressure can then be changed and the corresponding transition temperature determined. (b) The location of phase boundaries

Thermodynamics provides us with a way of predicting the location of the phase boundaries and relating their location and shape to the thermodynamic properties of the system. For instance, the shape of the vapor pressure curve (the liquid–vapor boundary) is related to the enthalpy of vaporization of the liquid. Suppose two phases, such as liquid and vapor, are in equilibrium at a given pressure and temperature. As we have seen, at a given temperature, the pressure corresponding to equilibrium is the vapor pressure of the liquid. If we change the temperature, the vapor pressure changes to a different value. That is, there is a relation between the change in temperature, dT, and the accompanying change in vapor pressure, dp. If we were considering the equilibrium between a solid and a liquid, the focus would be different: in this case we would typically be interested in the change in melting point as the pressure is increased. We show in the following Justification that the relation between dT and dp that ensures that in either case the two phases remain in equilibrium is given by the Clapeyron equation for the slope of the phase boundary at any temperature dp D trsH = dT TD trsV

Clapeyron equation

(3.5)

where D trsH is the enthalpy of transition and D trsV is the volume of transition (the change in molar volume that accompanies the transition) at the temperature of interest. For the liquid–vapor equilibrium, the equation in the form dp = (D vap H/ TD vapV)dT gives the change in vapor pressure when the temperature is changed; for the solid–liquid equilibrium, the equation in the form dT = (TD fusV/D fus H)dp gives the change in melting point caused by a change in pressure. For the solid–liquid phase boundary, the enthalpy of fusion is positive because melting is endothermic for all substances of interest in biology. For most substances, the molar volume increases slightly on melting, so D fusV is positive but small. It follows that the melting temperature changes very little when the pressure is changed. In other words, the slope of the phase boundary is large and positive (up from left to right). Water, however, is quite different, for although its melting is endothermic, its molar volume decreases on melting (liquid water is denser than ice at 0°C, which is why ice floats on water), so DfusV is small but negative. Consequently, an increase in pressure brings about a decrease in the melting point of ice.

Fig. 3.8 The cooling curve for the B–E section of the horizontal line in Fig. 3.5. The halt at D corresponds to the pause in cooling while the liquid freezes and releases its enthalpy of transition. The halt lets us locate Tf even if the transition cannot be observed visually.

102

3 PHASE EQUILIBRIA

A brief illustration

For water at 1 bar, D fus H 3 = 6.008 × 103 J mol−1 and DfusV 3 = −1.634 × 10−6 m3 mol−1. It follows from eqn 3.5 that at T = 273.15 K, the melting point of ice, dp 6.008 × 103 J mol−1 = = −1.346 × 107 Pa K−1 dT (273.15 K) × (−1.634 × 10−6 m3 mol−1) where we have used 1 Pa = 1 N m−2 and 1 J = 1 N m to write 1 Pa = 1 J m−3. In other words, the slope of the ice–water phase boundary is steep but negative (down from left to right).

For the liquid–vapor boundary (the vapor pressure curve), both the enthalpy and volume of vaporization are invariably positive, so the vapor pressure invariably increases with temperature (dp is positive if dT is positive). However, we have to be cautious because although the enthalpy of vaporization is not very sensitive to temperature, the volume of vaporization depends strongly on the temperature (through the effect of temperature on the volume of a gas). If we suppose that the vapor behaves as a perfect gas, then we show in the following Justification that the relation between a change in temperature and a change in vapor pressure is given by the Clausius–Clapeyron equation: d ln p D vap H = dT RT 2

Clausius–Clapeyron equation

(3.6)

A brief illustration

For water at 1 bar and 373.2 K (the boiling point of water), Dvap H 3 = 4.07 × 104 J mol−1, and it follows that d ln p 4.07 × 104 J mol−1 = = 3.51 × 10−2 K−1 dT (8.314 J K−1 mol−1) × (373.2 K)2 The liquid–vapor boundary of the phase diagram for water has a positive slope, and we shall see in Section 3.4(d) that the slope of the liquid–vapor phase boundary is much less steep than the slope of the ice–water phase boundary.

Justification 3.4 The Clapeyron and Clausius–Clapeyron equations

Fig. 3.9 At equilibrium, two phases have the same molar Gibbs energy. When the temperature is changed by dT, for the two phases to remain in equilibrium, the pressure must be changed by dp so that the Gibbs energies of the two phases remain equal.

The derivation of the Clapeyron equations is based on eqn 3.2, written as dGm = Vmdp − SmdT. At a certain pressure and temperature two phases, which we call 1 and 2 but can imagine to be a liquid and a vapor, respectively, are in equilibrium and Gm(1) = Gm(2) (Fig. 3.9). When the temperature is changed by dT and the pressure changes by dp, the molar Gibbs energies change as follows: dGm(1) = Vm(1)dp − Sm(1)dT

dGm(2) = Vm(2)dp − Sm(2)dT

The two phases are in equilibrium before the change and remain in equilibrium after the change, so the two changes in molar Gibbs energy must be equal, dGm(1) = dGm(2). It follows that Vm(1)dp − Sm(1)dT = Vm(2)dp − Sm(2)dT

3.4 PHASE DIAGRAMS

103

and therefore {Vm(2) − Vm(1)}dp = {Sm(2) − Sm(1)}dT With DtrsV = Vm(2) − Vm(1) and DtrsS = Sm(2) − Sm(1), this equation becomes DtrsVdp = DtrsSdT or dp DtrsS = dT DtrsV We saw in Section 2.2 that the transition entropy is related to the enthalpy of transition by D trsS = D trs H/Ttrs, so eqn 3.5 follows immediately. We have dropped the ‘trs’ subscript from the temperature in eqn 3.5 because all the points on the phase boundary—the only points we are considering in eqn 3.5—are transition temperatures. To move on to the Clausius–Clapeyron equation, we consider the case of vaporization. Because the molar volume of a gas is much larger than the molar volume of a liquid, we can replace D vapV = Vm(g) − Vm(l) by Vm(g) alone and write dp D H ≈ vap dT TVm(g) Next, we suppose that the vapor behaves as a perfect gas and write its molar volume as Vm(g) = RT/p. Then

A note on good practice

Keep a note of any approximations made in a derivation, for they limit the range of applicability of an expression. We have made two approximations in the derivation of the Clausius–Clapeyron equation: (1) the molar volume of a gas is much greater than that of a liquid and (2) the vapor behaves as a perfect gas.

dp D H pD H = vap = vap2 dT T(RT/p) RT and therefore 1 dp D vap H = p dT RT 2 A standard result of calculus is d ln x/dx = 1/x, and therefore (by multiplying both sides by dx), dx/x = d ln x. In this case, dp/p = d ln p, and eqn 3.6 follows.

(c) Characteristic points

We have seen that as the temperature of a liquid is raised, its vapor pressure increases. What we observe, however, depends on whether the heating takes place in a closed or an open container. First, consider what we would observe when we heat a liquid in an open vessel. At a certain temperature, the vapor pressure becomes equal to the external pressure. At this temperature, the vapor can drive back the surrounding atmosphere and expand indefinitely. Moreover, because there is no constraint on expansion, bubbles of vapor can form throughout the body of the liquid, the condition known as boiling. The temperature at which the vapor pressure of a liquid is equal to the external pressure is called the boiling temperature. When the external pressure is 1 atm, the boiling temperature is called the normal boiling point, Tb. It follows that we can predict the normal boiling point of a liquid by noting the temperature on the phase diagram at which its vapor pressure is 1 atm. Now consider what happens when we heat the liquid in a closed vessel. Because the vapor cannot escape, its density increases as the vapor pressure rises and in due course the density of the vapor becomes equal to that of the remaining liquid. At this stage the surface between the two phases disappears (Fig. 3.10).

Fig. 3.10 When a liquid is heated in a sealed container, the density of the vapor phase increases and that of the liquid phase decreases, as depicted here by the changing density of shading. There comes a stage at which the two densities are equal and the interface between the two fluids disappears. This disappearance occurs at the critical temperature. The container needs to be strong: the critical temperature of water is at 373°C and the vapor pressure is then 218 atm.

104

3 PHASE EQUILIBRIA

Table 3.1

Critical constants* pc /atm

Ammonia, NH3

Vc /(cm3 mol−1)

Tc /K

111

73

Argon, Ar

48

75

406 151

Benzene, C6H6

49

260

563

Carbon dioxide, CO2

73

94

304

Hydrogen, H2

13

65

33

Methane, CH4

46

99

191

Oxygen, O2

50

78

155

Water, H2O

218

55

647

*The critical volume, Vc , is the molar volume at the critical pressure and critical volume.

The significant points of a phase diagram. The liquid–vapor phase boundary terminates at the critical point. At the triple point, solid, liquid, and vapor are in dynamic equilibrium. The normal freezing point is the temperature at which the liquid freezes when the pressure is 1 atm; the normal boiling point is the temperature at which the vapor pressure of the liquid is 1 atm.

Fig. 3.11

The temperature at which the surface disappears is the critical temperature, Tc. The vapor pressure at the critical temperature is called the critical pressure, pc, and the critical temperature and critical pressure together identify the critical point of the substance (see Table 3.1). If we exert pressure on a sample that is above its critical temperature, we produce a denser fluid. However, no surface appears to separate the two parts of the sample and a single uniform phase, a supercritical fluid, continues to fill the container. That is, we have to conclude that a liquid cannot be produced by the application of pressure to a substance if it is at or above its critical temperature. That is why the liquid–vapor boundary in a phase diagram terminates at the critical point (Fig. 3.11). A supercritical fluid is not a true liquid, but it behaves like a liquid in many respects—for example, it has a density similar to that of a liquid.

A brief illustration

Supercritical carbon dioxide, scCO2, is the center of attention for an increasing number of solvent-based processes. The critical temperature of CO2, 304.2 K (31.0°C) and its critical pressure, 72.9 atm, are readily accessible, it is cheap, and it can readily be recycled. A great advantage of scCO2 is that there are no noxious residues once the solvent has been allowed to evaporate, so, coupled with its low critical temperature, scCO2 is ideally suited to food processing and the production of pharmaceuticals. It is used, for instance, to remove caffeine from coffee or fats from milk. The supercritical fluid is also increasingly being used for dry cleaning, which avoids the use of carcinogenic and environmentally damaging chlorinated hydrocarbons.

The temperature at which the liquid and solid phases of a substance coexist in equilibrium at a specified pressure is called the melting temperature of the substance. Because a substance melts at the same temperature as it freezes, the melting temperature is the same as the freezing temperature. The solid–liquid boundary therefore shows how the melting temperature of a solid varies with pressure.

3.4 PHASE DIAGRAMS

105

The melting temperature when the pressure on the sample is 1 atm is called the normal melting point or the normal freezing point, Tf. A liquid freezes when the energy of the molecules in the liquid is so low that they cannot escape from the attractive forces of their neighbors and lose their mobility. There is a set of conditions under which three different phases (typically solid, liquid, and vapor) all simultaneously coexist in equilibrium. It is represented by the triple point, where the three phase boundaries meet. The triple point of a pure substance is a characteristic, unchangeable physical property of the substance. For water the triple point lies at 273.16 K and 611 Pa, and ice, liquid water, and water vapor coexist in equilibrium at no other combination of pressure and temperature.1 At the triple point, the rates of each forward and reverse process are equal (but the three individual rates are not necessarily the same). The triple point and the critical point are important features of a substance because they act as frontier posts for the existence of the liquid phase. As we see from Fig. 3.12a, if the slope of the solid-liquid phase boundary is as shown in the diagram: The triple point marks the lowest temperature at which the liquid can exist. The critical point marks the highest temperature at which the liquid can exist. We shall see in the following section that for water, the solid–liquid phase boundary slopes in the opposite direction, and then only the second of these conclusions is relevant (see Fig. 3.12b). (d) The phase diagram of water

Figure 3.13 is the phase diagram for water. The liquid–vapor phase boundary shows how the vapor pressure of liquid water varies with temperature. We can use this curve to decide how the boiling temperature varies with changing external pressure. For example, when the external pressure is 149 Torr (at an altitude of 12 km), water boils at 60°C because that is the temperature at which the vapor pressure is 149 Torr (19.9 kPa). The solid–liquid boundary line in Fig. 3.14 shows how the melting temperature of water depends on the pressure. For example, although ice melts at 0°C at 1 atm, it melts at −1°C when the pressure is 130 atm. The very steep slope of the boundary indicates that enormous pressures are needed to bring about significant changes. Notice that the line slopes down from left to right, which—as we anticipated—means that the melting temperature of ice falls as the pressure is raised. We can trace the reason for this unusual behavior to the decrease in volume that occurs when ice melts: it is favorable for the solid to transform into the denser liquid as the pressure is raised. The decrease in volume is a result of the very open structure of the crystal structure of ice: as shown in Fig. 3.15, the water molecules are held apart, as well as together, by the hydrogen bonds between them, but the structure partially collapses on melting and the liquid is denser than the solid. Figure 3.13 shows that water has one liquid phase but many different solid phases other than ordinary ice (‘ice I’, shown in Fig. 3.15). These solid phases differ in the arrangement of the water molecules: under the influence of very high pressures, hydrogen bonds buckle and the H2O molecules adopt different 1 The triple point of water is used to define the Kelvin scale of temperatures: the triple point is defined as lying at 273.16 K exactly. The normal freezing point of water is found experimentally to lie approximately 0.01 K below the triple point, at very close to 273.15 K.

(a) For substances that have phase diagrams resembling the one shown here (which is common for most substances, with the important exception of water), the triple point and the critical point mark the range of temperatures over which the substance can exist as a liquid. The shaded areas show the regions of temperature in which a liquid cannot exist as a stable phase. (b) A liquid cannot exist as a stable phase if the pressure is below that of the triple point for normal or anomalous liquids.

Fig. 3.12

106

Fig. 3.13

3 PHASE EQUILIBRIA

The phase diagram for water showing the different solid phases.

Fig. 3.14 The solid–liquid boundary of water in more detail. The graph is schematic and not to scale.

arrangements. These polymorphs, or different solid phases, of ice may be responsible for the advance of glaciers, for ice at the bottom of glaciers experiences very high pressures where it rests on jagged rocks. The sudden apparent explosion of Halley’s comet in 1991 may have been due to the conversion of one form of ice into another in its interior. Figure 3.13 also shows that four or more phases of water (such as two solid forms, liquid, and vapor) are never in equilibrium. This observation is justified and generalized to all substances by the phase rule, which is derived in Further information 3.1.

Phase transitions in biopolymers and aggregates The structure of ice I. Each O atom is at the center of a tetrahedron of four O atoms at a distance of 276 pm. The central O atom is attached by two short O–H bonds to two H atoms and by two long hydrogen bonds to the H atoms of two of the neighboring molecules. Overall, the structure consists of planes of puckered hexagonal rings of H2O molecules (like the chair form of cyclohexane). This structure collapses partially on melting, leading to a liquid that is denser than the solid. Fig. 3.15

In Fundamentals F.1 and Chapter 2 we saw that proteins and biological membranes can exist in ordered structures stabilized by a variety of molecular interactions, such as hydrogen bonds and hydrophobic interactions. However, when certain conditions are changed, the helical and sheet structures of a polypeptide chain may collapse into a random coil and the hydrocarbon chains in the interior of bilayer membranes may become more or less flexible. These structural changes may be regarded as phase transitions in which molecular interactions in compact phases are disrupted at characteristic transition temperatures to yield phases in which the atoms can move more randomly. 3.5 The stability of nucleic acids and proteins To understand melting of proteins and nucleic acids at specific transition temperatures, we need to explore quantitatively the effect of intermolecular interactions on the stability of compact conformations of biopolymers.

3.5 THE STABILITY OF NUCLEIC ACIDS AND PROTEINS

From In the laboratory 1.1 we learned that the thermal denaturation of a biopolymer may be thought of as a kind of intramolecular melting from an organized structure to a flexible coil. This melting occurs at a specific melting temperature, Tm, which increases with the strength and number of intramolecular and intermolecular interactions in the material. Denaturation is a cooperative process in the sense that the biopolymer becomes increasingly more susceptible to denaturation once the process begins. This cooperativity is observed as a sharp step in a plot of fraction of unfolded polymer against temperature (Fig. 3.16). The melting temperature, Tm, is the temperature at which the fraction of unfolded polymer is 0.5. Closer examination of thermal denaturation reveals some of the chemical factors that determine protein and nucleic acid stability. For example, the thermal stability of DNA increases with the number of C–G base pairs in the sequence because each C–G base pair has three hydrogen bonds (1), whereas each T–A base pair has only two (2). More energy is required to unravel a double helix that has a higher proportion of hydrogen bonding interactions per base pair.

107

A protein unfolds as the temperature of the sample increases. The sharp step in the plot of fraction of unfolded protein against temperature indicates that the transition is cooperative. The melting temperature, Tm, is the temperature at which the fraction of unfolded polymer is 0.5.

Fig. 3.16

A note on good practice

Example 3.1

Predicting the melting temperature of DNA

The melting temperature of a DNA molecule can be determined by differential scanning calorimetry (In the laboratory 1.1). The following data were obtained in 0.010 m Na3PO4(aq) for a series of DNA molecules with varying base pair composition, with f the fraction of C–G base pairs: f Tm/K

0.375 339

0.509 344

0.589 348

0.688 351

0.750 354

Estimate the melting temperature of a DNA molecule containing 40.0 per cent C–G base pairs. Strategy We need to look for a quantitative relation between the melting temperature and the composition of DNA. We can begin by plotting Tm against fraction of C–G base pairs and examining the shape of the curve. If visual inspection of the plot suggests a linear relation, then the melting point at any composition can be predicted from the equation of the line that fits the data. Solution Figure 3.17 shows that Tm varies linearly with the fraction of C–G base pairs, at least in this range of composition. The equation of the line that fits the data is

Tm /K = 325 + 39.7f It follows that Tm = 341 K for 40.0 per cent C–G base pairs (at f = 0.400).

In this example we do not have a good theory to guide us in the choice of a mathematical model to describe the behavior of the system over a wide range of parameters. We are limited to finding a purely empirical relation—in this case a simple first-order polynomial equation—that fits the available data. It follows that we should not attempt to predict the property of a system that falls outside the narrow range of the data used to generate the fit because the mathematical model may have to be enhanced (for example, by using higherorder polynomial equations) to describe the system over a wider range of conditions. In the present case, we should not attempt to predict the Tm of DNA molecules outside the range 0.375 < f < 0.750.

108

3 PHASE EQUILIBRIA

The following calorimetric data were obtained in solutions containing 0.15 m NaCl(aq) for the same series of DNA molecules studied in Example 3.1. Estimate the melting temperature of a DNA molecule containing 40.0 per cent C–G base pairs under these conditions. Self-test 3.2

f Tm /K

0.375 359

0.509 364

0.589 368

0.688 371

0.750 374 Answer: 360 K

Data for Example 3.1 showing the variation of the melting temperature of DNA molecules with the fraction of C–G base pairs. All the samples also contain 1.0 × 10−2 mol dm−3 Na3PO4.

Fig. 3.17

Example 3.1 and Self-test 3.2 reveal that DNA is rather stable toward thermal denaturation, with Tm values ranging from about 340 K to 375 K, which is significantly higher than body temperature (310 K). The data also show that increasing the concentration of ions in solution increases the melting temperature of DNA. The stabilizing effect of ions can be traced to the fact that DNA has negatively charged phosphate groups decorating its surface. When the concentration of ions in solution is low, repulsive Coulomb interactions between neighboring phosphate groups destabilize the double helix and lower the melting temperature. On the other hand, positive ions, such as the Na+ ions in Self-test 3.2, bind electrostatically to the surface of DNA and mitigate repulsive interactions between phosphate groups. The result is stabilization of the double helical conformation and an increase in Tm. In contrast to DNA, proteins are relatively unstable toward thermal denaturation. For example, Tm = 320 K for ribonuclease T1 (an enzyme that cleaves RNA in the cell), which is close to body temperature. More surprisingly, the Gibbs energy for the unfolding of ribonuclease T1 at pH = 7.0 and 298 K is only +22.5 kJ mol−1, which is comparable to the energy required to break a single hydrogen bond (about 20 kJ mol−1) despite the fact that the formation of helices and sheets in proteins requires many hydrogen bonds. Therefore, unlike DNA, the stability of a protein does not increase in a simple way with the number of hydrogen bonding interactions. Although the reasons for the low stability of proteins are not known, the answer probably lies in a delicate balance of all intra- and intermolecular interactions that allow a protein to fold into its active conformation (Chapter 11). 3.6 Phase transitions of biological membranes To understand why cell membranes are sufficiently rigid to encase life’s molecular machines while being flexible enough to allow for cell division, we need to explore the factors that determine the melting temperatures of lipid bilayers.

A depiction of the variation with temperature of the flexibility of hydrocarbon chains in a lipid bilayer. (a) At physiological temperature, the bilayer exists as a liquid crystal, in which some order exists but the chains writhe. (b) At a specific temperature, the chains are largely frozen and the bilayer is said to exist as a gel.

Fig. 3.18

All lipid bilayers undergo a transition from a state of high to low chain mobility at a temperature that depends on the structure of the lipid. To visualize the transition, we consider what happens to a membrane as we lower its temperature (Fig. 3.18). There is sufficient energy available at normal temperatures for limited bond rotation to occur and the flexible chains to writhe around. However, the membrane is still highly organized in the sense that the bilayer structure does not come apart and the system is best described as a liquid crystal, a substance having liquid-like, imperfect long-range order in at least one direction in space but positional or orientational order in at least one other direction (Fig. 3.18a). At lower temperatures, the amplitudes of the writhing motion decrease until a specific temperature is reached at which motion is largely frozen. The membrane is then

3.6 PHASE TRANSITIONS OF BIOLOGICAL MEMBRANES

109

said to exist as a gel (Fig. 3.18b). Biological membranes exist as liquid crystals at physiological temperatures. Phase transitions in membranes are often observed as ‘melting’ from gel to liquid crystal by differential scanning calorimetry (In the laboratory 1.1). The data show relations between the structure of the lipid and the melting temperature. For example, the melting temperature increases with the length of the hydrophobic chain of the lipid. This correlation is reasonable, as we expect longer chains to be held together more strongly by hydrophobic interactions than shorter chains (Section 2.7). It follows that stabilization of the gel phase in the membranes of lipids with long chains results in relatively high melting temperatures. On the other hand, any structural elements that prevent alignment of the hydrophobic chains in the gel phase lead to low melting temperatures. Indeed, lipids containing unsaturated chains, those containing some C=C bonds, form membranes with lower melting temperatures than those formed from lipids with fully saturated chains, those consisting of C–C bonds only. Interspersed among the phospholipids of biological membranes are sterols, such as cholesterol (Atlas L1), which is largely hydrophobic but does contain a hydrophilic –OH group. Sterols, which are present in different proportions in different types of cells, prevent the hydrophobic chains of lipids from ‘freezing’ into a gel and, by disrupting the packing of the chains, spread the melting point of the membrane over a range of temperatures.

Organisms are capable of biosynthesizing lipids of different composition so that cell membranes have melting temperatures close to the ambient temperature. Why do bacterial and plant cells grown at low temperatures synthesize more phospholipids with unsaturated chains than do cells grown at higher temperatures?

Self-test 3.3

Answer: Insertion of lipids with unsaturated chains lowers the plasma membrane’s melting temperature to a value that is close to the lower ambient temperature.

Case study 3.1

The use of phase diagrams in the study of proteins

As in the discussion of pure substances, the phase diagram of a mixture shows which phase is most stable for the given conditions. However, composition is now a variable in addition to the pressure and temperature. Phase equilibria in binary mixtures may be explored by collecting data at constant pressure and displaying the results as a temperature–composition diagram, in which one axis is the temperature and the other axis is the mole fraction or concentration. Temperature–composition diagrams may be used to characterize intermediates in the unfolding of a protein caused by denaturation with a chemical agent. For example, urea, CO(NH2)2, competes for NH and CO groups, interferes with hydrogen bonding in a polypeptide, and disrupts the intramolecular interactions responsible for its native three-dimensional conformation. A temperature–composition diagram, such as the idealized form shown in Fig. 3.19, can reveal conditions under which different forms of the polypeptide can exist. The idealized diagram shows three structural regions, or phases: the native form, the unfolded form, and a ‘molten globule’ form, a partially unfolded but still compact form of the protein. As usual, two phases in equilibrium

An example of a temperature–composition diagram showing denaturation of a protein in a native phase into molten globule and fully unfolded phases. The concentrations marked x and y will be used in Exercise 3.39.

Fig. 3.19

110

3 PHASE EQUILIBRIA

define a line in the diagram, and a point represents a unique set of conditions under which the three phases are in equilibrium.

An example of a phase diagram in which the mole fraction of a precipitant, a substance that causes precipitation, is plotted against the mole fraction of a protein. The data help biochemists find conditions under which a protein crystallizes.

Fig. 3.20

In another type of phase diagram, the mole fraction or concentration of one component of a mixture is plotted against the mole fraction or concentration of another component, and experiments are conducted at constant temperature and pressure. Phase diagrams so constructed help biochemists find conditions under which a protein may form an ordered crystal amenable to study by X-ray diffraction techniques, which can reveal the three-dimensional arrangement of atoms in biological assemblies (see the Prologue and Chapter 11). A common crystallization technique for charged proteins consists of adding large amounts of a salt, such as (NH4)2SO4, to a buffer solution containing the biopolymer. The increase in the ionic strength of the solution decreases the solubility of the protein to such an extent that the protein precipitates. The idealized phase diagram in Fig. 3.20 shows that ordered crystals precipitate over a relatively narrow range of protein and salt concentrations. Precise knowledge of crystallization conditions is a key to the reproducibility of X-ray diffraction experiments.

The thermodynamic description of mixtures We now leave pure materials and the limited but important changes they can undergo and examine mixtures. We shall consider only hom*ogeneous mixtures, or solutions, in which the composition is uniform however small the sample. The component in smaller abundance is called the solute and that in larger abundance is the solvent. These terms, however, are normally but not invariably reserved for solids dissolved in liquids; one liquid mixed with another is normally called simply a ‘mixture’ of the two liquids. In this chapter we consider mainly nonelectrolyte solutions, where the solute is not present as ions. Examples are sucrose dissolved in water, sulfur dissolved in carbon disulfide, and a mixture of ethanol and water. Although we also consider some of the special problems of electrolyte solutions, in which the solute consists of ions that interact strongly with one another, we defer a full study until Chapter 5. The measures of concentration commonly encountered in physical chemistry are reviewed in Further information 3.2. 3.7 The chemical potential To assess the spontaneity of a biological process involving a mixture, we need to know how to compute the contribution of each substance to the total Gibbs energy of the mixture.

A partial molar property is the contribution (per mole) that a substance makes to an overall property of a mixture. The most important partial molar property for our purposes is the partial molar Gibbs energy, GJ,m, of a substance J, which is the contribution of J (per mole of J) to the total Gibbs energy of a mixture. It follows that if we know the partial molar Gibbs energies of two substances A and B in a mixture of a given composition, then we can calculate the total Gibbs energy of the mixture by using G = nAGA,m + nBGB,m

(3.7)

3.8 IDEAL AND IDEAL–DILUTE SOLUTIONS

To gain insight into the significance of the partial molar Gibbs energy, consider a mixture of ethanol and water. Ethanol has a particular partial molar Gibbs energy when it is pure (and every molecule is surrounded by other ethanol molecules), and it has a different partial molar Gibbs energy when it is in an aqueous solution of a certain composition (because then each ethanol molecule is surrounded by a mixture of ethanol and water molecules). The partial molar Gibbs energy is so important in chemistry that it is given a special name and symbol. From now on, we shall call it the chemical potential and denote it m (mu). Then eqn 3.7 becomes G = nA mA + nB mB

(3.8)

where mA is the chemical potential of A in the mixture and mB is the chemical potential of B. In the course of this chapter and the next we shall see that the name ‘chemical potential’ is very appropriate, for it will become clear that mJ is a measure of the ability of J to bring about physical and chemical change. A substance with a high chemical potential has a high ability, in a sense we shall explore, to drive a reaction or some other physical process forward. We saw in Section 3.1 that the molar Gibbs energy of a pure substance is the same in all the phases at equilibrium. We can use the same argument to show in the following Justification that a system is at equilibrium when the chemical potential of each substance has the same value in every phase in which it occurs. We can think of the chemical potential as the pushing power of each substance, and equilibrium is reached only when each substance pushes with the same strength in any phase it occupies. Justification 3.5 The uniformity of chemical potential

Suppose a substance J occurs in different phases in different regions of a system. For instance, we might have a liquid mixture of ethanol and water and a mixture of their vapors. Let the substance J have chemical potential mJ(l) in the liquid mixture and mJ(g) in the vapor. We could imagine an infinitesimal amount, dnJ, of J migrating from the liquid to the vapor. As a result, the Gibbs energy of the liquid phase falls by mJ(l)dnJ and that of the vapor rises by mJ(g) dnJ. The net change in Gibbs energy is dG = mJ(g)dnJ − mJ(l)dnJ = {mJ(g) − mJ(l)}dnJ There is no tendency for this migration (and the reverse process, migration from the vapor to the liquid) to occur, and the system is at equilibrium if dG = 0, which requires that mJ(g) = mJ(l). The argument applies to each component of the system. Therefore, for a substance to be at equilibrium throughout the system, its chemical potential must be the same everywhere, as asserted in the text.

3.8 Ideal and ideal–dilute solutions Because in biochemistry we are concerned primarily with liquid solutions, we need expressions for the chemical potentials of solutes and solvents.

We need an explicit formula for the variation of the chemical potential of a substance with the composition of the mixture. Here we use the strategy mentioned at the start of the chapter: we begin by considering the chemical potential of a gas, not because gases are particularly interesting in biology but because we can use the resulting expression to derive results for solutions.

111

112

3 PHASE EQUILIBRIA

(a) The chemical potential of a gas

Our starting point is eqn 3.3, Gm(pf ) = Gm(pi) + RT ln(pf /pi), which shows how the molar Gibbs energy of a perfect gas depends on pressure. First, we set pf = p, the pressure of interest, and pi = p3, the standard pressure (1 bar). At the latter pressure, the molar Gibbs energy has its standard value, G m3 , so we can write Gm(p) = G m3 + RT ln(p/p3)

(3.9)

Next, for a mixture of perfect gases, we interpret p as the partial pressure of the gas, and Gm is the partial molar Gibbs energy, the chemical potential. Therefore, for a mixture of perfect gases, for each component J present at a partial pressure pJ, mJ = m3J + RT ln(pJ/p3)

(3.10a)

3 J

The variation with partial pressure of the chemical potential of a perfect gas. Note that the chemical potential increases with pressure.

Fig. 3.21

In this expression, m is the standard chemical potential of the gas J, which is identical to its standard molar Gibbs energy, the value of Gm for the pure gas at 1 bar. If we adopt the convention that, whenever pJ appears in a formula, it is to be interpreted as pJ/p3 (so, if the pressure is 2.0 bar, pJ = 2.0), we can write eqn 3.10a more simply as mJ = m3J + RT ln pJ

(3.10b)

Figure 3.21 illustrates the pressure dependence of the chemical potential of a perfect gas predicted by this equation. Note that the chemical potential becomes negatively infinite as the pressure tends to zero, rises to its standard value at 1 bar (because ln 1 = 0), and then increases slowly (logarithmically, as ln p) as the pressure is increased further. Equation 3.10 tells us that the higher the partial pressure of a gas, the higher its chemical potential. This conclusion is consistent with the interpretation of the chemical potential as an indication of the potential of a substance to be active chemically: the higher the partial pressure, the more active chemically the species. In this instance the chemical potential represents the tendency of the substance to react when it is in its standard state (the significance of the term m3) plus an additional tendency that reflects whether it is at a different pressure. A higher partial pressure gives a substance more chemical ‘punch’, just like winding a spring gives a spring more physical punch (that is, enables it to do more work).

Suppose that the partial pressure of a perfect gas falls from 1.00 bar to 0.50 bar as it is consumed in a reaction at 25°C. What is the change in chemical potential of the substance? Self-test 3.4

Answer: −1.7 kJ mol−1

(b) The chemical potential of a solvent

We can anticipate that the chemical potential of a species ought to increase with concentration because the higher its concentration, the greater its chemical ‘punch’. In the following, we use J to denote a substance in general, A to denote a solvent, and B to denote a solute. The key to linking the properties of a solution to those of a gas and setting up an expression for the chemical potential of a solute is the work done by the French chemist François Raoult (1830–1901), who spent most of his life measuring the vapor pressures of solutions. He measured the partial vapor pressure, pJ, of each component in the mixture, the partial pressure of the vapor of each component in

3.8 IDEAL AND IDEAL–DILUTE SOLUTIONS

113

dynamic equilibrium with the liquid mixture, and established what is now called Raoult’s law: The partial vapor pressure of a substance in a liquid mixture is proportional to its mole fraction in the mixture and its vapor pressure when pure: pJ = xJ p*J

Raoult’s law

(3.11)

In this expression, p*J is the vapor pressure of the pure substance. A brief illustration

When the mole fraction of water in an aqueous solution is 0.90, then, provided Raoult’s law is obeyed, the partial vapor pressure of the water in the solution is 90 per cent that of pure water. This conclusion is approximately true whatever the identity of the solute and the solvent (Fig. 3.22). The partial vapor pressures of the two components of an ideal binary mixture are proportional to the mole fractions of the components in the liquid. The total pressure of the vapor is the sum of the two partial vapor pressures.

Fig. 3.22

The molecular origin of Raoult’s law is the effect of the solute on the entropy of the solution. The entropy of the solvent arises from the random locations and the thermal motion of its molecules. The vapor pressure then represents the tendency of the system and its surroundings to reach a higher entropy. When a solute is present, the molecules in the solution are more dispersed than in the pure solvent, so we cannot be sure that a molecule chosen at random will be a solvent molecule (Fig. 3.23). Because the entropy of the solution is higher than that of the pure solvent, the solution has a lower tendency to acquire an even higher entropy by the solvent vaporizing. In other words, the vapor pressure of the solvent in the solution is lower than that of the pure solvent. A hypothetical solution of a solute B in a solvent A that obeys Raoult’s law throughout the composition range from pure A to pure B is called an ideal solution. The law is most reliable when the components of a mixture have similar molecular shapes and are held together in the liquid by similar types and strengths of intermolecular forces. An example is a mixture of two structurally similar hydrocarbons. A mixture of benzene and methylbenzene (toluene) is a good approximation to an ideal solution, for the partial vapor pressure of each component satisfies Raoult’s law reasonably well throughout the composition range from pure benzene to pure methylbenzene (Fig. 3.24).

Two similar substances, in this case benzene and methylbenzene (toluene), behave almost ideally and have vapor pressures that closely resemble those for the ideal case depicted in Fig. 3.22.

Fig. 3.24

(a) In a pure liquid, we can be confident that any molecule selected from the sample is a solvent molecule. (b) When a solute is present, we cannot be sure that blind selection will give a solvent molecule, so the entropy of the system is greater than in the absence of the solute.

Fig. 3.23

114

3 PHASE EQUILIBRIA

No mixture is perfectly ideal, and all real mixtures show deviations from Raoult’s law. However, the deviations are small for the component of the mixture that is in large excess (the solvent) and become smaller as the concentration of solute decreases (Fig. 3.25). We can usually be confident that Raoult’s law is reliable for the solvent when the solution is very dilute. More formally, Raoult’s law is a limiting law (like the perfect gas law) and is strictly valid only at the limit of zero concentration of solute. The theoretical importance of Raoult’s law is that, because it relates vapor pressure to composition and we know how to relate pressure to chemical potential, we can use the law to relate chemical potential to the composition of a solution. As we show in the following Justification, the chemical potential of a solvent A present in solution at a mole fraction xA is mA = mA* + RT ln xA

Strong deviations from ideality are shown by dissimilar substances, in this case carbon disulfide and acetone (propanone). Note, however, that Raoult’s law is obeyed by propanone when only a small amount of carbon disulfide is present (on the left) and by carbon disulfide when only a small amount of propanone is present (on the right).

Fig. 3.25

Chemical potential of the solvent in an ideal solution

(3.12)

where mA* is the chemical potential of pure A. This expression is valid throughout the concentration range for either component of a binary ideal solution. It is valid for the solvent of a real solution the closer the composition approaches pure solvent (pure A). Justification 3.6 The chemical potential of a solvent

When a solvent A in a solution is in equilibrium with its vapor at a partial pressure pA, the chemical potentials of the two phases are equal and we can write mA(l) = mA(g) (Fig. 3.26). However, we have just derived an expression for the chemical potential of a vapor, eqn 3.10, so at equilibrium mA(l) = m3A(g) + RT ln pA According to Raoult’s law, pA = xA pA*, so we can use the relation ln(xy) = ln x + ln y to write mA(l) = m3A(g) + RT ln (xA pA*) = m3A(g) + RT ln pA* + RT ln xA

A note on good practice

An asterisk (*) denotes a pure substance, but not one that is necessarily in its standard state. Only if the pressure is 1 bar would mA* be the standard chemical potential of A, and it would then be written as m3A.

The first two terms on the right, m3A(g) and RT ln pA*, are independent of the composition of the mixture and can be combined into the constant mA*, the chemical potential of pure liquid A. Equation 3.12 then follows. Figure 3.27 shows the variation of the chemical potential of the solvent predicted by this expression. Note that the chemical potential has its pure value at xA = 1 (when only A is present). The essential feature of eqn 3.12 is that because xA < 1 implies that ln xA < 0, the chemical potential of a solvent is lower in a solution than when it is pure. Provided the solution is almost ideal, a solvent in which a solute is present has less chemical ‘punch’ (including a lower ability to generate a vapor pressure) than when it is pure. By how much is the chemical potential of benzene reduced at 25°C by a solute that is present at a mole fraction of 0.10? Self-test 3.5

Answer: 0.26 kJ mol−1

(c) The chemical potential of a solute

Raoult’s law provides a good description of the vapor pressure of the solvent in a very dilute solution, when the solvent A is almost pure. However, we cannot in

3.8 IDEAL AND IDEAL–DILUTE SOLUTIONS

At equilibrium, the chemical potential of a substance in its liquid phase is equal to the chemical potential of the substance in its vapor phase.

Fig. 3.26

Fig. 3.27 The variation of the chemical potential of the solvent with the composition of the solution. Note that the chemical potential of the solvent is lower in the mixture than for the pure liquid (for an ideal system). This behavior is likely to be shown by a dilute solution in which the solvent is almost pure (and obeys Raoult’s law).

general expect it to be a good description of the vapor pressure of the solute B because a solute in dilute solution is very far from being pure. In a dilute solution, each solute molecule is surrounded by nearly pure solvent, so its environment is quite unlike that in the pure solute, and except when solute and solvent are very similar (such as benzene and methylbenzene), it is very unlikely that the vapor pressure of the solute will be related in a simple manner to the vapor pressure of the pure solute. However, it is found experimentally that in dilute solutions, the vapor pressure of the solute is in fact proportional to its mole fraction, just as for the solvent. Unlike the solvent, however, the constant of proportionality is not in general the vapor pressure of the pure solute. This linear but different dependence was discovered by the English chemist William Henry (1774–1836) and is summarized as Henry’s law: The vapor pressure of a volatile solute B is proportional to its mole fraction in a solution: pB = K H′ xB

Henry’s law

(3.13)

Here K′H, which is called Henry’s law constant, is characteristic of the solute and chosen so that the straight line predicted by eqn 3.13 is tangent to the experimental curve at xB = 0 (Fig. 3.28). Henry’s law is usually obeyed only at low concentrations of the solute (close to xB = 0). Solutions that are dilute enough for the solute to obey Henry’s law are called ideal–dilute solutions. The Henry’s law constants of some gases are listed in Table 3.2. The values given there are for the law rewritten to show how the molar concentration depends on the partial pressure, rather than vice versa: [J] = KH pJ

Another version of Henry’s law

(3.14)

115

When a component (the solvent) is almost pure, it behaves in accord with Raoult’s law and has a vapor pressure that is proportional to the mole fraction in the liquid mixture and a slope p*, the vapor pressure of the pure substance. When the same substance is the minor component (the solute), its vapor pressure is still proportional to its mole fraction, but the constant of proportionality is now KH′ . Fig. 3.28

Table 3.2 Henry’s law constants for gases dissolved in water at 25°C

KH / (mol m−3 kPa−1) Carbon dioxide, CO2

3.39 × 10−1

Hydrogen, H2

7.78 × 10−3

Methane, CH4

1.48 × 10−2

Nitrogen, N2

6.48 × 10−3

Oxygen, O2

1.30 × 10−2

116

3 PHASE EQUILIBRIA

The Henry’s law constant, KH, is commonly reported in moles per cubic metre per kilopascal (mol m−3 kPa−1). This form of the law and these units make it very easy to calculate the molar concentration of the dissolved gas, simply by multiplying the partial pressure of the gas (in kilopascals) by the appropriate constant. Equation 3.14 is used, for instance, to estimate the concentration of O2 in natural waters or the concentration of carbon dioxide in blood plasma.

Example 3.2

Determining whether a natural water can support aquatic life

The concentration of O2 in water required to support aerobic aquatic life is about 4.0 mg dm−3. What is the minimum partial pressure of oxygen in the atmosphere that can achieve this concentration? A note on good practice

The number of significant figures in the result of a calculation should not exceed the number in the data.

Strategy The strategy of the calculation is to determine the partial pressure of

oxygen that, according to Henry’s law (written as eqn 3.14), corresponds to the concentration specified. Solution Equation 3.14 becomes

pO = 2

[O2] KH

We note that the molar concentration of O2 is [O2] =

4.0 × 10−3 g dm−3 4.0 × 10−3 mol 4.0 × 10−3 mol 4.0 = = = mol m−3 32 dm3 32 × 10−3 m3 32 32 g mol−1

From Table 3.2, KH for oxygen in water is 1.30 × 10−2 mol m−3 kPa−1, therefore the partial pressure needed to achieve the stated concentration is pO = 2

(4.0/32) mol m−3 = 9.6 kPa 1.30 × 10−2 mol m−3 kPa−1

The partial pressure of oxygen in air at sea level is 21 kPa (158 Torr), which is greater than 9.6 kPa (72 Torr), so the required concentration can be maintained under normal conditions.

What partial pressure of methane is needed to dissolve 21 mg of methane in 100 g of benzene at 25°C (K′H = 5.69 × 104 kPa, for Henry’s law in the form given in eqn 3.13)? Self-test 3.6

Answer: 57 kPa (4.3 × 102 Torr) The variation of the chemical potential of the solute with the composition of the solution expressed in terms of the mole fraction of solute. Note that the chemical potential of the solute is lower in the mixture than for the pure solute (for an ideal system). This behavior is likely to be shown by a dilute solution in which the solvent is almost pure and the solute obeys Henry’s law.

Fig. 3.29

Henry’s law lets us write an expression for the chemical potential of a solute in a solution. We show in the following Justification that the chemical potential of the solute when it is present at a mole fraction xB is mB = mB* + RT ln xB

Chemical potential of the solute in terms of the mole fraction

(3.15)

This expression, which is illustrated in Fig. 3.29, applies when Henry’s law is valid, in very dilute solutions. The chemical potential of the solute has its pure value when it is present alone (xB = 1, ln 1 = 0) and a smaller value when dissolved (when xB < 1, ln xB < 0).

3.8 IDEAL AND IDEAL–DILUTE SOLUTIONS

117

Justification 3.7 The chemical potential of the solute

We apply the same reasoning as in Justification 3.6. When a solute B in a solution is in equilibrium with its vapor at a partial pressure pB, we can write mB(l) = mB(g) and (from eqn 3.10) mB(l) = m3B (g) + RT ln pB According to Henry’s law, pB = K H′ xB, so it follows that mB(l) = m3B (g) + RT ln K′HxB = m3B (g) + RT ln K′H + RT ln xB The terms m3B (g) and RT ln KH are independent of the composition of the mixture and can be combined into the constant m*, B the chemical potential of pure liquid B. Equation 3.15 then follows.

We often express the composition of a solution in terms of the molar concentration of the solute, [B], rather than as a mole fraction. The mole fraction and the molar concentration are proportional to each other in dilute solutions, so we write xB = constant × [B]/c 3, where c 3 = 1 mol dm−3 is introduced to ensure that the constant is dimensionless. We shall call c 3 the standard molar concentration. Then eqn 3.15 becomes mB = m*B + RT ln(constant) + RT ln([B]/c 3)

A note on good practice

It is meaningless to take logarithms of quantities with units, so always ensure that the x of ln x is a pure number.

We can combine the first two terms into a single constant, which we denote m3B , and write this relation as mB = m3B + RT ln([B]/c 3)

Chemical potential of the solute in terms of the molar concentration

(3.16a)

This equation is the best way to write the relation, but it is cumbersome, and for the rest of the chapter we shall write [B]/c 3 simply as [B] and—to conform to the requirement stated in the note on good practice—interpret [B] as the molar concentration with the units deleted (we treated pressure similarly earlier in the chapter). Thus, if in fact [B] = 0.1 mol dm−3, so [B]/c 3= 0.1, from now on we shall write [B] = 0.1 and use eqn 3.16a in the form mB = m3B + RT ln[B]

Simplified form of eqn 3.16a

(3.16b)

Figure 3.30 illustrates the variation of chemical potential with concentration predicted by this equation. The chemical potential of the solute has its standard value when the molar concentration of the solute is c 3 = 1 mol dm−3. At this stage a summary of the results so far might be helpful: Species Gas, J Solvent, A Solute, B

Chemical potential mJ = m3J + RT ln pJ mA = mA* + RT ln xA mB = m*B + RT ln xB mB = m3B + RT ln[B]

Case study 3.2

Comment Perfect gas Dilute solution Dilute solution

Gas solubility and breathing

We inhale about 500 cm3 of air with each breath we take. The influx of air is a result of changes in volume of the lungs as the diaphragm is depressed and the chest expands, which results in a decrease in pressure of about 100 Pa

The variation of the chemical potential of the solute with the composition of the solution that obeys Henry’s law expressed in terms of the molar concentration of solute. The chemical potential has its standard value at [B] = 1 mol dm−3.

Fig. 3.30

118

3 PHASE EQUILIBRIA

relative to atmospheric pressure. Expiration occurs as the diaphragm rises and the chest contracts, and gives rise to a differential pressure of about 100 Pa above atmospheric pressure. The total volume of air in the lungs is about 6 dm3, and the additional volume of air that can be exhaled forcefully after normal expiration is about 1.5 dm3. Some air remains in the lungs at all times to prevent the collapse of the alveoli. The effect of gas exchange between blood and air inside the alveoli of the lungs means that the composition of the air in the lungs is different from that in the atmosphere, and changes throughout the breathing cycle. Alveolar gas is in fact a mixture of newly inhaled air and air about to be exhaled. The concentration of oxygen present in arterial blood is equivalent to a partial pressure of about 40 Torr (5.3 kPa), whereas the partial pressure of freshly inhaled air in the alveoli of the lungs is about 100 Torr (13.3 kPa). Arterial blood remains in the capillary passing through the wall of an alveolus for about 0.75 s, but such is the steepness of the pressure gradient that it becomes fully saturated with oxygen in about 0.25 s. If the lungs collect fluids (as in pneumonia), then the respiratory membrane thickens, diffusion is greatly slowed, and body tissues begin to suffer from oxygen starvation. Carbon dioxide moves in the opposite direction across the respiratory tissue, but the partial pressure gradient is much less, corresponding to about 5 Torr (0.7 kPa) in blood and 40 Torr (5.3 kPa) in air at equilibrium in the alveoli of the lungs. However, because carbon dioxide is much more soluble in the alveolar fluid than oxygen is, equal amounts of oxygen and carbon dioxide are exchanged in each breath. A hyperbaric oxygen chamber, in which oxygen is at an elevated partial pressure, is used to treat certain types of disease. Carbon monoxide poisoning can be treated in this way, as can the consequences of shock. Diseases that are caused by anaerobic bacteria, such as gas gangrene and tetanus, can also be treated because the bacteria cannot thrive in high oxygen concentrations.

(d) Real solutions: activities

No actual solutions are ideal, and many solutions deviate from ideal–dilute behavior as soon as the concentration of solute rises above a small value. In thermodynamics we try to preserve the form of equations developed for ideal systems so that it becomes easy to step between the two types of system.2 This is the thought behind the introduction of the activity, aJ, of a substance, which is a kind of effective concentration. The activity is defined so that the expression The chemical potential in terms of the activity

mJ = m3J + RT ln aJ

(3.17)

is true at all concentrations and for both the solvent and the solute. For ideal solutions, aJ = xJ, and the activity of each component is equal to its mole fraction. For ideal–dilute solutions using the definition in eqn 3.17, aB = [B]/c 3, and the activity of the solute is equal to the numerical value of its molar concentration. For non-ideal solutions we write For the solvent: aA = gAxA For the solute: aB = gB[B]/c 2

3

The activity in terms of the activity coefficient

An added advantage is that there are fewer equations to remember!

(3.18)

3.8 IDEAL AND IDEAL–DILUTE SOLUTIONS

Table 3.3

Activities and standard states*

Substance

Standard state

Activity, a

Solid

Pure solid, 1 bar

1

Liquid

Pure liquid, 1 bar

1

Gas

Pure gas, 1 bar

p/p3

Solute

Molar concentration of 1 mol dm−3

[J]/c3

p3 = 1 bar (= 105 Pa), c3 = 1 mol dm−3. *Activities are for perfect gases and ideal–dilute solutions; all activities are dimensionless.

where the g (gamma) in each case is the activity coefficient. Activity coefficients depend on the composition of the solution, and we should note the following: Because the solvent behaves more in accord with Raoult’s law as it becomes pure, gA → 1 as xA → 1. Because the solute behaves more in accord with Henry’s law as the solution becomes very dilute, gB → 1 as [B] → 0. Conventions concerning standard states and activities of ideal systems are summarized in Table 3.3. Activities and activity coefficients are often branded as ‘fudge factors’. To some extent that is true. However, their introduction does allow us to derive thermodynamically exact expressions for the properties of nonideal solutions. Moreover, in a number of cases it is possible to calculate or measure the activity coefficient of a species in solution. In this text we shall normally derive thermodynamic relations in terms of activities, but when we want to make contact with actual measurements, we shall set the activities equal to the ‘ideal’ values in Table 3.3. Case study 3.3

The Donnan equilibrium

The term Donnan equilibrium refers to the distribution of ions between two solutions in contact through a semipermeable membrane, in one of which there is a polyelectrolyte, such as NanP (with Pn− a polyanion), and where the membrane is not permeable to the large charged macromolecule. This arrangement is one that actually occurs in living systems, where we have seen that osmosis is an important feature of cell operation. The thermodynamic consequences of the distribution and transfer of charged species across cell membranes is explored further in Chapter 5. Consider a situation in which a high concentration of a salt such as NaCl is added to the solution on both sides of the membrane so that the number of cations that Pn− provides is insignificant in comparison with the number supplied by the additional salt. Apart from small imbalances of charge close to the membrane (which have important consequences, as we shall see in Chapter 5), electrical neutrality must be preserved in the bulk on both sides of the membrane: if an anion migrates, a cation must accompany it. For simplicity, we take the volumes of the solutions on each side of the membrane to be equal. On one side of the membrane—call it the ‘left-hand’ side—there are Pn−, Na+, and Cl− ions. In the ‘right-hand’ side there are Na+ and Cl− ions. The condition

119

120

3 PHASE EQUILIBRIA for equilibrium is that the chemical potentials of the Na+ and Cl− ions in solution are the same in both sides, so a net flow of Na+ and Cl− ions occurs until the chemical potentials are equalized. This equality occurs when m3(Na+) + m3(Cl−) + RT ln aL(Na+) + RT ln aL(Cl−) = m3(Na+) + m3(Cl−) + RT ln aR(Na+) + RT ln aR(Cl−) where the subscripts L and R refer to the left-hand and right-hand sides, respectively, separated by the membrane. It follows that RT ln aL(Na+)aL(Cl−) = RT ln aR(Na+)aR(Cl−) If we ignore activity coefficients and interpret [Na+]/c 3 and [Cl−]/c 3 as [Na+] and [Cl−], respectively, the two expressions are equal when [Na+]L[Cl−]L = [Na+]R[Cl−]R. As the Na+ ions are supplied by the polyelectrolyte as well as the added salt, the conditions for bulk electrical neutrality lead to the chargebalance equations [Na+]L = [Cl−]L + n[Pn−] and [Na+]R = [Cl−]R. We can now combine these three conditions to obtain expressions for the differences in ion concentrations across the membrane. For example, we write [Na+]L =

[Na+]R[Cl−]R [Na+]R2 = − + [Cl ]L [Na ]L − n[Pn−]

which rearranges to [Na+]L2 − [Na+]R2 = n[Pn−][Na+]L After applying the relation a2 − b2 = (a + b)(a − b) and rearranging, we obtain [Na+]L − [Na+]R =

n[Pn−][Na+]L [Na+]L + [Na+]R

It follows from the definition [Cl−] = 12 ([Cl−]L + [Cl−]R) and the charge–balance equations that [Na+]L + [Na+]R = [Cl−]L + [Cl−]R + n[Pn−] = 2[Cl−] + n[Pn−] Substitution of this result into the equation for [Na+]L − [Na+]R leads to [Na+]L − [Na+]R =

n[Pn−][Na+]L 2[Cl−] + n[Pn−]

(3.19a)

Similar manipulations lead to an equation for the difference in chloride ion concentration: [Cl−]L − [Cl−]R = −

n[Pn−][Cl−]L [Cl−]L + [Cl−]R

which becomes [Cl−]L − [Cl−]R = −

n[Pn−][Cl−]L 2[Cl−]

(3.19b)

Note that cations will dominate the anions in the compartment that contains the polyanion because the concentration difference is positive for Na+ and negative for Cl−. It also follows that from a measurement of the ion concentrations, it is possible to determine the net charge of the polyanion, which may be unknown.

3.8 IDEAL AND IDEAL–DILUTE SOLUTIONS

Example 3.3

Analyzing a Donnan equilibrium

Suppose that two equal volumes of 0.200 m NaCl(aq) solution are separated by a membrane and that the left-hand side of the experimental arrangement contains a polyelectrolyte Na6P at a concentration of 50 g dm−3. Assuming that the membrane is not permeable to the polyanion, which has a molar mass of 55 kg mol−1, calculate the molar concentrations of Na+ and Cl− in each compartment. Strategy We saw above that the sum of the equilibrium concentrations of Na+ in both compartments is

[Na+]L + [Na+]R = 2[Cl−] + n[Pn−] with [Cl−] = 0.200 mol dm−3, and [Pn−] being calculated from the mass concentration and the molar mass of the polyanion. At this point, we have one equation and two unknowns, [Na+]L and [Na+]R, so we use a second equation, eqn 3.19a, to solve for both Na+ ion concentrations. To calculate the Cl− ion concentrations, we use [Cl−]R = [Na+]R and [Cl−]L = [Na+]L − n[Pn−], with n = 6. Solution The molar concentration of the polyanion is [Pn−] = 9.1 × 10−4 mol

dm−3. It follows from eqn 3.19a that [Na+]L − [Na+]R =

6 × (9.1 × 10−4 mol dm−3) × [Na+ ]L 2 × (0.200 mol dm−3) + 6 × (9.1 × 10−4 mol dm−3)

The sum of Na+ concentrations is [Na+]L + [Na+]R = 2 × (0.200 mol dm−3) + 6 × (9.1 × 10−4 mol dm−3) = 0.405 mol dm−3 The solutions of these two equations are [Na+]L = 0.204 mol dm−3

[Na+]R = 0.201 mol dm−3

Then [Cl−]R = [Na+]R = 0.201 mol dm−3 [Cl−]L = [Na+]L − 6[Pn−] = 0.199 mol dm−3 Repeat the calculation for 0.300 m NaCl(aq), a polyelectrolyte Na10P of molar mass 33 kg mol−1 at a mass concentration of 50.0 g dm−3.

Self-test 3.7

Answer: [Na+]L = 0.31 mol dm−3, [Na+]R = 0.30 mol dm−3

(e) The thermodynamics of dissolving

We now have enough information to formulate a thermodynamic description of dissolving to form an ideal solution. As we see in the following Justification, when an amount nB of a solute B dissolves in an amount nA of a solvent A at a temperature T, DG = nRT{xA ln xA + xB ln xB} with n = nA + nB and the xJ the mole fractions in the mixture.

Gibbs energy of dissolving

(3.20)

121

122

3 PHASE EQUILIBRIA

Justification 3.8 The Gibbs energy of dissolving

The Gibbs energy of the two unmixed components is the sum of their individual Gibbs energies: Gi = nA mA* + nB m*B When B is dissolved in A to form an ideal solution, the Gibbs energy becomes Gf = nA mA + nB mB = nA{mA* + RT ln xA} + nB{m*B + RT ln xB} = nA mA* + nART ln xA + nB m*B + nBRT ln xB where the xJ are the mole fractions of the two components in the solution. The difference Gf − Gi is the change in Gibbs energy that accompanies dissolving. The pure chemical potentials cancel, so DG = RT{nA ln xA + nB ln xB} Because xJ = nJ/n, we can substitute nA = xAn and nB = xBn into the expression above and obtain The variation of the Gibbs energy of dissolving with composition for two components at constant temperature and pressure. Note that DG < 0 for all compositions, which indicates that two components mix spontaneously in all proportions.

Fig. 3.31

DG = nRT{xA ln xA + xB ln xB} which is eqn 3.20.

Equation 3.20 tells us the change in Gibbs energy when a solute dissolves to give an ideal solution (Fig. 3.31). The crucial feature is that because xA and xB are both less than 1, the two logarithms are negative (ln x < 0 if x < 1), so DG < 0 at all compositions. Therefore, dissolving to form an ideal solution is spontaneous in all proportions. Furthermore, if we compare eqn 3.20 with DG = DH − TDS, we can conclude that: DH = 0

Enthalpy of dissolving

(3.21a)

DS = −nR{xA ln xA + xB ln xB}

Entropy of dissolving

(3.21b)

The value of DH indicates that although there are interactions between the molecules, the solute–solute, solvent–solvent, and solute–solvent interactions are all the same, so the solute slips into solution without a change in enthalpy. There is an increase in entropy because the molecules are more dispersed in the solution than in the unmixed components. The entropy of the surroundings is unchanged because the enthalpy of the system is constant, so no energy escapes as heat into the surroundings. It follows that the increase in entropy of the system is the ‘driving force’ of the dissolving.

Colligative properties An ideal solute has no effect on the enthalpy of a solution in the sense that the enthalpy of mixing is zero. However, it does affect the entropy, and we found in eqn 3.21 that DS > 0 when a solute dissolves in a solvent to give an ideal solution. We can therefore expect a solute to modify the physical properties of the solution. Apart from lowering the vapor pressure of the solvent, which we have already considered, a nonvolatile solute has three main effects: it raises the boiling point of a solution, it lowers the freezing point, and it gives rise to an osmotic pressure.

3.9 THE MODIFICATION OF BOILING AND FREEZING POINTS

Table 3.4

Cryoscopic and ebullioscopic constants

Solvent

K f /(K kg mol−1)

K b /(K kg mol−1)

Acetic acid

3.90

3.07

Benzene

5.12

2.53

Camphor

40

Carbon disulfide

3.8

2.37

Naphthalene

6.94

5.8

Phenol

7.27

Tetrachloromethane Water

3.04

30

4.95

1.86

0.51

(The meaning of the last will be explained shortly.) These properties, which are called colligative properties, stem from a change in the dispersal of solvent molecules that depends on the number of solute particles present but is independent of the identity of the species we use to bring it about.3 Thus, a 0.01 mol kg−1 aqueous solution of any nonelectrolyte should have the same boiling point, freezing point, and osmotic pressure. 3.9 The modification of boiling and freezing points To understand the origins of the colligative properties and their effect on biological processes, it is useful to explore the modification of the boiling and freezing points of a solvent in a solution.

It is found empirically, and can be justified thermodynamically, that the elevation of boiling point, Tb, and the depression of freezing point, Tf, are both proportional to the molality, bB, of the solute: DTb = KbbB

Elevation of the boiling point

DTf = Kf bB

Depression of the freezing point

(3.22)

where Kb is the ebullioscopic constant and Kf is the cryoscopic constant of the solvent.4 The two constants can be estimated from other properties of the solvent, but both are best treated as empirical constants (Table 3.4). Self-test 3.8 Estimate the lowering of the freezing point of the solution made by dissolving 3.0 g (about one cube) of sucrose in 100 g of water.

Answer: 0.16 K

To understand the origin of these effects, we shall make two simplifying assumptions: 1) The solute is not volatile and therefore does not appear in the vapor phase. 2) The solute is insoluble in the solid solvent and therefore does not appear in the solid phase. 3 4

Hence, the name colligative, meaning ‘depending on the collection’. They are also called the ‘boiling-point constant’ and the ‘freezing-point constant’.

123

124

3 PHASE EQUILIBRIA

The chemical potentials of pure solid solvent and pure liquid solvent also decrease with temperature, and the point of intersection, where the chemical potential of the liquid rises above that of the solid, marks the freezing point of the pure solvent. A solute lowers the chemical potential of the solvent but leaves that of the solid unchanged. As a result, the intersection point lies farther to the left and the freezing point is therefore lowered.

Fig. 3.32

Mathematical toolkit 3.2

For example, a solution of sucrose in water consists of a solute (sucrose, C12H22O11) that is not volatile and therefore never appears in the vapor, which is therefore pure water vapor. The sucrose is also left behind in the liquid solvent when ice begins to form, so the ice remains pure. The origin of colligative properties is the lowering of chemical potential of the solvent by the presence of a solute, as expressed by eqn 3.12. We saw in Section 3.3 that the freezing and boiling points correspond to the temperatures at which the graph of the molar Gibbs energy of the liquid intersects the graphs of the molar Gibbs energy of the solid and vapor phases, respectively. Because we are now dealing with mixtures, we have to think about the partial molar Gibbs energy (the chemical potential) of the solvent. The presence of a solute lowers the chemical potential of the liquid, but because the vapor and solid remain pure, their chemical potentials remain unchanged. As a result, we see from Fig. 3.32 that the freezing point moves to lower values; likewise, from Fig. 3.33 we see that the boiling point moves to higher values. In other words, the freezing point is depressed, the boiling point is elevated, and the liquid phase exists over a wider range of temperatures. The elevation of boiling point is too small to have any practical significance. A practical consequence of the lowering of freezing point, and hence the lowering of the melting point of the pure solid, is its employment in organic chemistry to judge the purity of a sample, for any impurity lowers the melting point of a substance from its accepted value. The salt water of the oceans freezes at temperatures lower than that of fresh water, and salt is spread on highways to delay the onset of freezing. The addition of ‘antifreeze’ to car engines and, by natural processes, to arctic fish, is commonly held up as an example of the lowering of freezing point, but the concentrations are far too high for the arguments we have used here to be applicable. The 1,2-ethanediol (‘glycol’) used as antifreeze probably just interferes with bonding between water molecules. Likewise, the antifreeze proteins of arctic fish act by binding to small ice crystals and preventing larger crystals from forming.

Power series and expansions

A power series has the form c0 + c1(x − a) + c2(x − a) + · · · + cn(x − a) + · · · 2

=

n

∑ cn(x − a)n n=0

where cn and a are constants. It is often useful to express a function f(x) in the vicinity of x = a as a special power series called the Taylor series, or Taylor expansion, which has the form f(x) = f(a) +

1 A dn f D (x − a)n + · · · n! C dx n F a 1 A dn f D (x − a)n n! C dx n F a

···+ =

n=0

A df D 1 A d2 f D (x − a) + (x − a)2 + C dx F a 2! C dx 2 F a

where n! denotes a factorial given by n! = n(n − 1) (n − 2) . . . 1. The following Taylor expansions are often useful: (1 + x)−1 = 1 − x + x 2 − · · · ex = 1 + x + 12 x 2 + · · · ln x = (x − 1) − 12 (x − 1)2 + 13 (x − 1)3 − 14 (x − 1)4 + · · · ln(1 + x) = x − 12 x2 + 13 x 3 − · · · If x 0 and the slope of G is positive (up from left to right) when the mixture is rich in the products C and D because mC and mD are then high. At compositions corresponding to DrG < 0 the reaction tends to form more products; where D rG > 0, the reverse reaction is spontaneous, and the products tend to decompose into reactants. Where D rG = 0 (at the minimum of the graph where the slope is zero), the reaction has no tendency to form either products or reactants. In other words, the reaction is at equilibrium. That is, the criterion for chemical equilibrium at constant temperature and pressure is D rG = 0

Criterion of chemical equilibrium

(4.2)

4.2 The variation of D r G with composition The reactants and products in a biological cell are rarely at equilibrium, so we need to know how the reaction Gibbs energy depends on their concentrations.

Our starting point is the general expression for the composition dependence of the chemical potential derived in Section 3.8: mJ = mJ3 + RT ln aJ

Chemical potential of a species J

Fig. 4.2 The variation of Gibbs energy with progress of reaction showing how the reaction Gibbs energy, DrG, is related to the slope of the curve at a given composition. When DG and Dn are both infinitesimal, the slope is written dG/dn.

(4.3)

where aJ is the activity of the species J. When we are dealing with systems that may be treated as ideal, which will be the case in this chapter, we use the identifications given in Table 3.3: For solutes in an ideal solution, aJ = [J]/c 3, the molar concentration of J relative to the standard value c 3 = 1 mol dm−3. For perfect gases, aJ = pJ /p3, the partial pressure of J relative to the standard pressure p3 = 1 bar. For pure solids and liquids, aJ = 1. As in Chapter 3, to simplify the appearance of expressions in what follows, we shall not write c 3 and p3 explicitly. (a) The reaction quotient

Substitution of eqn 4.3 into eqn 4.1c gives DrG = {c(mC3 + RT ln aC ) + d(mD3 + RT ln aD)} − {a(mA3 + RT ln aA) + b(mB3 + RT ln aB)} = {(cmC3 + dmD3) − (amA3 + bmA3)} + RT{c ln aC + d ln aD − a ln aA − b ln aB} The first term on the right in the second equality is the standard reaction Gibbs energy, DrG 3: D rG 3 = {cmC3 + dmD3} − {amA3 + bmB3}

(4.4a)

Fig. 4.3 At the minimum of the curve, corresponding to equilibrium, DrG = 0. To the left of the minimum, DrG < 0, and the forward reaction is spontaneous. To the right of the minimum, DrG > 0, and the reverse reaction is spontaneous.

138

4 CHEMICAL EQUILIBRIUM

Because the standard states refer to the pure materials, the standard chemical potentials in this expression are the standard molar Gibbs energies of the (pure) species. Therefore, eqn 4.4a is the same as DrG 3 = {cGm3 (C) + dGm3 (D)} − {aGm3 (A) + bGm3 (B)}

Standard Gibbs energy of reaction

(4.4b)

We consider this important quantity in more detail shortly. At this stage, therefore, we know that D rG = DrG 3 + RT{c ln aC + d ln aD − a ln aA − b ln aB} and the expression for DrG is beginning to look much simpler. To make further progress, we rearrange the remaining terms on the right as follows: c ln aC + d ln aD − a ln aA − b ln aB = ln acC + ln adD − ln aaA − ln abB ac ad = ln acCadD − ln aaAabB = ln Ca Db a AaB At this point, we have deduced that DrG = DrG 3 + RT ln

acC adD aaAadB

(4.5)

To simplify the appearance of this expression still further, we introduce the (dimensionless) reaction quotient, Q, for reaction C: Q=

acC adD aaAabB

Definition of reaction quotient

(4.6)

Note that Q has the form of products divided by reactants, with the activity of each species raised to a power equal to its stoichiometric coefficient in the reaction; because activities are dimensionless quantities, Q is a dimensionless quantity. We can now write the overall expression for the reaction Gibbs energy at any composition of the reaction mixture as D rG = DrG 3 + RT ln Q

Reaction Gibbs energy

(4.7)

This simple but hugely important equation will occur several times in different disguises.

Example 4.1

Formulating a reaction quotient

Formulate the reaction quotients for reactions A (the isomerism of glucose-6phosphate) and B (the binding of oxygen to hemoglobin). Strategy Use Table 3.3 to express activities in terms of molar concentrations or pressures. Then use eqn 4.6 to write an expression for the reaction quotient Q. In reactions involving gases and solutes, the expression for Q will contain pressures and molar concentrations.

4.2 THE VARIATION OF D rG WITH COMPOSITION

Solution The reaction quotient for reaction A is

Q=

aF6P [F6P]/c 3 [F6P] = = aG6P [G6P]/c 3 [G6P]

For reaction B, the binding of oxygen to hemoglobin, the reaction quotient is Q=

aHb(O ) [Hb(O2)4]/c 3 = 4 aHbaO ([Hb]/c 3)(pO /p3)4 2 4

2

2

Because we are not writing the standard concentration and pressure explicitly, this expression simplifies to Q=

[Hb(O2)4] [Hb]pO4 2

with pJ the numerical value of the partial pressure of J in bar (so if pO = 2.0 bar, we just write pO = 2.0 when using this expression). 2

2

Self-test 4.1 Write the reaction quotient for the esterification reaction CH3COOH + C2H5OH → CH3COOC2H5 + H2O. (All four components are present in the reaction mixture as liquids: the mixture is not an aqueous solution.)

Answer: Q ≈ [CH3COOC2H5][H2O]/[CH3COOH][C2H5OH]

(b) Biological standard states

The thermodynamic definition of standard states of solutes takes them as being at unit activity (in elementary work, at c 3 = 1 mol dm−3). The conventional standard state of hydrogen ions (aH O = 1, corresponding to pH = 0, a strongly acidic solution) is not appropriate to normal biological conditions inside cells, where the pH is close to 7. Therefore, in biochemistry it is common to adopt the biological standard state, in which pH = 7, a neutral solution. When we adopt this convention we label the corresponding standard quantities as G ⊕, H ⊕, and S ⊕.1 Equation 4.8 allows us to relate the two standard Gibbs energies of formation. For a reaction of the form +

3

reactants + n H3O+(aq) → products the biological and thermodynamic standard states are related by DrG ⊕ = DrG 3 − RT ln(10−7)n = DrG 3 + 7nRT ln 10

Relation between standard values

(4.8)

where we have used the relations (x a)b = x ab and ln x ab = ab ln x. It follows that at 298.15 K:

DrG ⊕ = DrG 3 + n(39.96 kJ mol−1)

at 37°C (310 K, body temperature): DrG ⊕ = DrG 3 + n(41.5 kJ mol−1) There is no difference between thermodynamic and biological standard values if hydrogen ions are not involved in the reaction (n = 0).

1

Another convention to denote the biological standard state is to write X o′ or X3′.

139

140

4 CHEMICAL EQUILIBRIUM

Example 4.2

Converting between thermodynamic and biological standard states

The standard reaction Gibbs energy for the hydrolysis of ATP is +10 kJ mol−1 at 298 K. What is the biological standard state value? Strategy Because protons occur as products, lowering their concentration (from 1 mol dm−3 to 10−7 mol dm−3) suggests that the reaction will have a higher tendency to form products. Therefore, we expect a more negative value of the reaction Gibbs energy for the biological standard than for the thermodynamic standard. The two types of standard are related by eqn 4.8, with the activity of hydrogen ions 10−7 in place of 1. Solution The reaction quotient for the hydrolysis reaction + ATP4−(aq) + H2O(l) → ADP3−(aq) + HPO2− 4 (aq) + H3O (aq)

when all the species are in their standard states except the hydrogen ions, which are present at 10−7 mol dm−3, is Q=

aADP aHPO aH O 1 × 1 × 10−7 = = 10−7 aATP − aH O 1×1 3−

2−

3

4

+

4

2

The thermodynamic and biological standard values are therefore related by eqn 4.8 in the form DrG ⊕ = DrG 3 + RT ln(10−7) At 298 K DrG ⊕ = 10 kJ mol−1 + (8.3145 J K−1 mol−1) × (298 K) × ln(10−7) = 10 kJ mol−1 − 40 kJ mol−1 = −30 kJ mol−1 Note how the large change in pH changes the sign of the standard reaction Gibbs energy. The overall reaction for the glycolysis reaction (Case study 4.3) is C6H12O6(aq) + 2 NAD+(aq) + 2 ADP3−(aq) + 2 HPO42−(aq) + 2 H2O(l) → 2 CH3COCO2−(aq) + 2 NADH(aq) + 2 ATP4−(aq) + 2 H3O+(aq). For this reaction, DrG ⊕ = −80.6 kJ mol−1 at 298 K. What is the value of D rG 3?

Self-test 4.2

Answer: −0.7 kJ mol−1

4.3 Reactions at equilibrium We need to be able to identify the equilibrium composition of a reaction so that we can discuss the approach to equilibrium systematically.

At equilibrium, the reaction quotient has a certain (dimensionless) value called the equilibrium constant, K, of the reaction: K=

A aCc aDd D C aaAabB F equilibrium

Definition of equilibrium constant

(4.9)

We shall not normally write equilibrium; the context will always make it clear that Q refers to an arbitrary stage of the reaction, whereas K, the value of Q at

4.3 REACTIONS AT EQUILIBRIUM

141

equilibrium, is calculated from the equilibrium composition. It now follows from eqn 4.7 that at equilibrium 0 = D rG 3 + RT ln K and therefore that D rG 3 = −RT ln K

Expression for calculating an equilibrium constant

(4.10a)

This is one of the most important equations in the whole of chemical thermodynamics. Its principal use is to predict the value of the equilibrium constant of any reaction from tables of thermodynamic data, like those in the Resource section. Alternatively, we can use it to determine DrG 3 by measuring the equilibrium constant of a reaction.

A brief illustration

The first step in the metabolic breakdown of glucose is its phosphorylation to G6P: glucose(aq) + ATP(aq) → G6P(aq) + ADP(aq) + H+(aq) The standard reaction Gibbs energy for the reaction is −34 kJ mol−1 at 37°C, so it follows from eqn 4.10a that D G3 (−3.4 × 104 J mol−1) 3.4 × 104 ln K = − r = − = −1 −1 RT (8.3145 J K mol ) × (310 K) 8.3145 × 310 To calculate the equilibrium constant of the reaction, which (like the reaction quotient) is a dimensionless number, we use the relation eln x = x with x = K: K = e3.4×10 /8.3145×310 = 5.4 × 105 4

Self-test 4.3 Calculate the equilibrium constant of the reaction N2(g) + 3 H2(g) 7 2 NH3(g) at 25°C, given that DrG 3 = −32.90 kJ mol−1.

Answer: 5.8 × 105

If the biological standard state is used in place of the thermodynamic standard state, we write eqn 4.10a in the same way, D rG ⊕ = −RT ln K

Expression for calculating an equilibrium constant

(4.10b)

but interpret the concentration of any hydronium ions that occurs in K as relative to c ⊕ = 10−7 mol dm−3 rather than relative to c 3 = 1 mol dm−3. That is, we interpret the activity of hydronium ions in the expression for K as aH o = [H3O+]/c ⊕. 3

+

A brief illustration

The biological standard reaction Gibbs energy for the reaction in the preceding brief illustration is −75.5 kJ mol−1. The same calculation illustrated there but with this value gives K′ = 5.3 × 1012. This value is for

A note on good practice

The exponential function (ex) is very sensitive to the value of x, so evaluate it only at the end of a numerical calculation.

142

4 CHEMICAL EQUILIBRIUM

K′ =

[G6P][ADP]([H3O+]/10−7) = 5.3 × 1012 [glucose][ATP]

whereas the former value was for K=

[glucose][ADP][H3O+] = 5.4 × 105 [F6P][ATP]

Multiplication of both sides of the first equation of this pair by 10−7 gives a result in accord with the second equation.

(a) The significance of the equilibrium constant

Fig. 4.4 The relation between standard reaction Gibbs energy and the equilibrium constant of the reaction.

An important feature of eqn 4.10a is that it tells us that K > 1 if DrG 3< 0 (and correspondingly that K ′ > 1 if DrG ⊕ < 0). Broadly speaking, K > 1 implies that products are dominant at equilibrium, so we can conclude that a reaction is thermodynamically feasible if D rG 3 < 0 (Fig. 4.4). Conversely, because eqn 4.10a tells us that K < 1 if D rG 3 > 0, then we know that the reactants will be dominant in a reaction mixture at equilibrium if D rG 3 > 0. In other words, a reaction with DrG 3 > 0 is not thermodynamically feasible. Some care must be exercised with these rules, however, because the products will be significantly more abundant than reactants only if K >> 1 (more than about 103), and even a reaction with K < 1 may have a reasonable abundance of products at equilibrium. Table 4.1 summarizes the conditions under which DrG 3 < 0 and K > 1. Because D rG 3 = D r H 3 − TD r S 3, the standard reaction Gibbs energy is certainly negative if both D r H 3 < 0 (an exothermic reaction) and Dr S 3 > 0 (a reaction system that becomes more disorderly, such as by forming a gas). The standard reaction Gibbs energy is also negative if the reaction is endothermic (D r H 3 > 0) and TDr S 3 is sufficiently large and positive. Note that for an endothermic reaction to have D rG 3 < 0, its standard reaction entropy must be positive. Moreover, the temperature must be high enough for TD r S 3 to be greater than D r H 3 (Fig. 4.5). The switch of D rG 3 from positive to negative, corresponding to the switch from K < 1 (the reaction ‘does not go’) to K > 1 (the reaction ‘goes’), occurs at a temperature given by equating Dr H 3 − TDr S 3 to 0, which gives T=

Fig. 4.5 An endothermic reaction may have K > 1 provided the temperature is high enough for TDr S 3 to be large enough that, when subtracted from DrH 3, the result is negative.

Dr H 3 Dr S 3

Table 4.1

Thermodynamic criteria of spontaneity

1. If the reaction is exothermic (D r H 3 < 0) and D r S 3 > 0 D rG 3 < 0 and K > 1 at all temperatures 2. If the reaction is exothermic (D r H 3 < 0) and D r S 3 < 0 D rG 3 < 0 and K > 1 provided that T < D r H 3/D r S 3 3. If the reaction is endothermic (D r H 3 > 0) and D r S 3 > 0 D rG 3 < 0 and K > 1 provided that T > D r H 3/D r S 3 4. If the reaction is endothermic (D r H 3 > 0) and D r S 3 < 0 D rG 3 < 0 and K > 1 at no temperature

Temperature at which an endothermic reaction becomes spontaneous

(4.11)

4.3 REACTIONS AT EQUILIBRIUM

Suppose that the enthalpy change accompanying the dissociation of base pairs is of the order of +15 kJ per mole of base pairs and the corresponding entropy change is 45 J K−1 mol−1. At what temperature can you expect a DNA chain to denature spontaneously?

Self-test 4.4

Answer: 60°C

(b) The composition at equilibrium

An equilibrium constant expresses the composition of an equilibrium mixture as a ratio of products of activities. Even if we confine our attention to ideal systems, it is still necessary to do some work to extract the actual equilibrium concentrations or partial pressures of the reactants and products given their initial values (see, for example, Example 4.5).

Example 4.3

Calculating an equilibrium composition

Consider reaction A, for which DrG 3 = +1.7 kJ mol−1 at 25°C. Estimate the fraction f of F6P in equilibrium with G6P at 25°C, where f is defined as f=

[F6P] [F6P] + [G6P]

Strategy Express f in terms of K. To do so, recognize that if the numerator and

denominator in the expression for f are both divided by [G6P]; then the ratios [F6P]/[G6P] can be replaced by K. Calculate the value of K by using eqn 4.10a. Solution Division of the numerator and denominator by [G6P] gives

f=

[F6P]/[G6P] K = [F6P]/[G6P] + 1 K + 1

We find the equilibrium constant by rearranging eqn 4.10a into K = e−D G /RT r

3

with DrG 3 1.7 × 103 J mol−1 1.7 × 103 = = RT (8.3145 J K−1 mol−1) × (298 K) 8.3145 × 298 Therefore, K = e−1.7×10 /8.3145×298 = 0.50 3

and f=

0.50 = 0.33 0.50 + 1

That is, at equilibrium, 33 per cent of the solute is F6P and 67 per cent is G6P.

Estimate the composition of a solution in which two isomers A and B are in equilibrium (A 7 B) at 37°C and D rG 3 = −2.2 kJ mol−1.

Self-test 4.5

Answer: The fraction of B at equilibrium is feq = 0.70.

143

144

4 CHEMICAL EQUILIBRIUM

(c) The molecular origin of chemical equilibrium

Fig. 4.6 The Boltzmann distribution of populations over the energy levels of two species A and B with similar densities of energy levels; the reaction A → B is endothermic in this example. The bulk of the population is associated with the species A, so that species is dominant at equilibrium.

We can obtain a deeper insight into the origin and significance of the equilibrium constant by considering the Boltzmann distribution of molecules over the available states of a system composed of reactants and products. When atoms can exchange partners, as in a chemical reaction, the available states of the system include arrangements in which the atoms are present in the form of reactants and other atoms are present in the form of products. These arrangements have their characteristic sets of energy levels, but the Boltzmann distribution does not distinguish between their identities, only their energies. The atoms distribute themselves over both sets of energy levels in accord with the Boltzmann distribution (Fig. 4.6). At a given temperature, there is a specific distribution of populations and hence a specific composition of the reaction mixture. It can be appreciated from Fig. 4.6 that if the reactants and products both have similar arrays of molecular energy levels, then the dominant species in a reaction mixture at equilibrium will be the species with the lower set of energy levels. However, the fact that the equilibrium constant is related to the Gibbs energy (through ln K = −D rG 3/RT) is a signal that entropy plays a role as well as energy. Its role can be appreciated by referring to Fig. 4.7. We see that although the B energy levels lie higher than the A energy levels, in this instance they are much more closely spaced. As a result, their total population may be considerable and B could even dominate in the reaction mixture at equilibrium. Closely spaced energy levels correlate with a high entropy, so in this case we see that entropy effects dominate adverse energy effects. That is, a positive reaction enthalpy results in a lowering of the equilibrium constant (that is, an endothermic reaction can be expected to have an equilibrium composition that favors the reactants). However, if there is positive reaction entropy, then the equilibrium composition may favor products, despite the endothermic character of the reaction. Case study 4.1

Binding of oxygen to myoglobin and hemoglobin

Biochemical equilibria can be far more complex than those we have considered so far, but exactly the same principles apply. An example of a complex process is the binding of O2 by hemoglobin in blood, which is described only approximately by reaction B. The protein myoglobin (Mb, Atlas P10) stores O2 in muscle, and the protein hemoglobin (Hb, Atlas P7) transports O2 in blood. These two proteins are related, for hemoglobin is a tetramer of four myoglobin-like molecules. In each protein, the O2 molecule attaches to an iron ion in a heme group (Atlas R2) (Fig. 4.8). Fig. 4.7 Even though the reaction A → B is endothermic, the density of energy levels in B is so much greater than that in A, the population associated with B is greater than that associated with A; hence B is dominant at equilibrium.

First, consider the equilibrium between Mb and O2: Mb(aq) + O2(g) 7 MbO2(aq)

K=

[MbO2] [Mb]p

where p is the numerical value of the partial pressure of O2 gas in bar. It follows that the fractional saturation, s, the fraction of Mb molecules that are oxygenated, is s=

[MbO2] [MbO2] Kp = = [Mb]total [Mb] + [MbO2] 1 + Kp

The dependence of s on p is shown in Fig. 4.9.

Fractional saturation of myoglobin

(4.12)

4.3 REACTIONS AT EQUILIBRIUM

145

Now consider the equilibrium between Hb and O2: Hb(aq) + O2(g) 7 HbO2(aq)

K1 =

[HbO2] [Hb]p

HbO2(aq) + O2(g) 7 Hb(O2)2(aq)

K2 =

[Hb(O2)2] [HbO2]p

Hb(O2)2(aq) + O2(g) 7 Hb(O2)3(aq)

K3 =

[Hb(O2)3] [Hb(O2)2]p

Hb(O2)3(aq) + O2(g) 7 Hb(O2)4(aq)

[Hb(O2)4] K4 = [Hb(O2)3]p

To develop an expression for s, we express [Hb(O2)2] in terms of [HbO2] by using K2, then express [HbO2] in terms of [Hb] by using K1, and likewise for all the other concentrations of Hb(O2)3 and Hb(O2)4. It follows that [HbO2] = K1[Hb]p

[Hb(O2)2] = K1K2[Hb]p2

[Hb(O2)3] = K1 K2 K3[Hb]p3

[Hb(O2)4] = K1K2 K3 K4[Hb]p4

Fig. 4.8 One of the four polypeptide chains that make up the human hemoglobin molecule. The chains, which are similar to the oxygen storage protein myoglobin, consist of helical and sheet-like regions. The heme group is at the lower left.

The total concentration of bound O2 is [O2]bound = [HbO2] + 2[Hb(O2)2] + 3[Hb(O2)3] + 4[Hb(O2)4] = (1 + 2K2 p + 3K2 K3 p2 + 4K2 K3 K4 p3 )K1[Hb]p where we have used the fact that n O2 molecules are bound in Hb(O2)n, so the concentration of bound O2 in Hb(O2)2 is 2[Hb(O2)2], and so on. The total concentration of hemoglobin is [Hb]total = (1 + K1 p + K1K2 p2 + K1 K2 K3 p3 + K1 K2 K3 K4 p4)[Hb] Because each Hb molecule has four sites at which O2 can attach, the fractional saturation is s= =

[O2]bound 4[Hb]total (1 + 2K2 p + 3K2 K3 p2 + 4K2 K3 K4 p3)K1 p 4(1 + K1 p + K1K2 p2 + K1 K2 K3 p3 + K1 K2 K3 K4 p4)

Fractional saturation of hemoglobin

(4.13)

A reasonable fit of the experimental data can be obtained with K1 = 0.01, K2 = 0.02, K3 = 0.04, and K4 = 0.08 when p is expressed in torr. The binding of O2 to hemoglobin is an example of cooperative binding, in which the binding of a ligand (in this case O2) to a biopolymer (in this case Hb) becomes more favorable thermodynamically (that is, the equilibrium constant increases) as the number of bound ligands increases up to the maximum number of binding sites. We see the effect of cooperativity in Fig. 4.9. Unlike the myoglobin saturation curve, the hemoglobin saturation curve is sigmoidal (S shaped): the fractional saturation is small at low ligand concentrations, increases sharply at intermediate ligand concentrations, and then levels off at high ligand concentrations. Cooperative binding of O2 by hemoglobin is explained by an allosteric effect, in which an adjustment of the conformation of a molecule when one substrate binds affects the ease with which a subsequent substrate molecule binds. The details of the allosteric effect in hemoglobin will be explored in Case study 10.4.

Fig. 4.9 The variation of the fractional saturation of myoglobin and hemoglobin molecules with the partial pressure of oxygen. The different shapes of the curves account for the different biological functions of the two proteins.

146

4 CHEMICAL EQUILIBRIUM

The differing shapes of the saturation curves for myoglobin and hemoglobin have important consequences for the way O2 is made available in the body: in particular, the greater sharpness of the Hb saturation curve means that Hb can load O2 more fully in the lungs and unload it more fully in different regions of the organism. In the lungs, where p ≈ 105 Torr (14 kPa), s ≈ 0.98, representing almost complete saturation. In resting muscular tissue, p is equivalent to about 38 Torr (5 kPa), corresponding to s ≈ 0.75, implying that sufficient O2 is still available should a sudden surge of activity take place. If the local partial pressure falls to 22 Torr (3 kPa), s falls to about 0.1. Note that the steepest part of the curve falls in the range of typical tissue oxygen partial pressure. Myoglobin, on the other hand, begins to release O2 only when p has fallen below about 22 Torr, so it acts as a reserve to be drawn on only when the Hb oxygen has been used up. 4.4 The standard reaction Gibbs energy The standard reaction Gibbs energy is central to the discussion of chemical equilibria and the calculation of equilibrium constants. It is also a useful indicator of the energy available from catabolism to drive anabolic processes, such as the synthesis of proteins.

We have seen that standard reaction Gibbs energy, DrG 3, is defined as the difference in standard molar Gibbs energies of the products and the reactants weighted by the stoichiometric coefficients, n, in the chemical equation D rG 3 =

∑ nGm3 − Reactants ∑ nGm3 Products

Definition of standard Gibbs energy of reaction

(4.14)

For example, the standard reaction Gibbs energy for reaction A is the difference between the molar Gibbs energies of fructose-6-phosphate and glucose-6phosphate in solution at 1 mol dm−3 and 1 bar. We cannot calculate DrG 3 from the standard molar Gibbs energies themselves because these quantities are not known. One practical approach is to calculate the standard reaction enthalpy from standard enthalpies of formation (Section 1.11), the standard reaction entropy from Third-Law entropies (Section 2.5), and then to combine the two quantities by using D rG 3 = D r H 3 − TDr S3

Example 4.4

Construction of D rG 3

(4.15)

Calculating the standard reaction Gibbs energy of an enzymecatalyzed reaction

Evaluate the standard reaction Gibbs energy at 25°C for the reaction CO2(g) + H2O(l) → H2CO3(aq) catalyzed by the enzyme carbonic anhydrase in red blood cells. Strategy Obtain the relevant standard enthalpies of formation and standard

entropies from the Resource section. Then calculate the standard reaction enthalpy and the standard reaction entropy from Dr H 3 =

nDf H 3 −

nSm3 −

Products

DrS 3 =

Products

nDf H 3

Reactants

nSm3

Reactants

and the standard reaction Gibbs energy from eqn 4.14.

4.4 THE STANDARD REACTION GIBBS ENERGY

Solution The standard reaction enthalpy is

D r H 3 = D f H 3(H2CO3,aq) − {D f H 3(CO2,g) + D f H 3(H2O,l)} = −699.65 kJ mol−1 − {(−393.51 kJ mol−1) + (−285.83 kJ mol−1)} = −20.31 kJ mol−1 The standard reaction entropy was calculated in the brief illustration in Section 2.5: DrS 3 = −96.3 J K−1 mol−1 which, because 96.3 J is the same as 9.63 × 10−2 kJ, corresponds to −9.63 × 10−2 kJ K−1 mol−1. Therefore, from eqn 4.15, DrG 3 = (−20.31 kJ mol−1) − (298.15 K) × (−9.63 × 10−2 kJ K−1 mol−1) = +8.40 kJ mol−1 Use the information in the Resource section to determine the standard reaction Gibbs energy for 3 O2(g) → 2 O3(g) from standard enthalpies of formation and standard entropies. Self-test 4.6

Answer: +326.4 kJ mol−1

(a) Standard Gibbs energies of formation

We saw in Section 1.11 how to use standard enthalpies of formation of substances to calculate standard reaction enthalpies. We can use the same technique for standard reaction Gibbs energies. To do so, we list the standard Gibbs energy of formation, D f G 3, of a substance, which is the standard reaction Gibbs energy (per mole of the species) for its formation from the elements in their reference states. The concept of reference state was introduced in Section 1.11 (a reminder: it is the most stable form of the element under the prevailing conditions; do not confuse ‘reference state’ with ‘standard state’, but be aware that a reference state of an element will also be in its standard state if the pressure is 1 bar); the temperature is arbitrary, but we shall almost always take it to be 25°C (298 K). For example, the standard Gibbs energy of formation of liquid water, D f G 3(H2O,l), is the standard reaction Gibbs energy for H2(g) + 12 O2(g) → H2O(l) and is −237 kJ mol−1 at 298 K. Some standard Gibbs energies of formation are listed in Table 4.2 and more can be found in the Resource section. It follows from the definition that the standard Gibbs energy of formation of an element in its reference state is zero because reactions such as C(s, graphite) → C(s, graphite) are null (that is, nothing happens). The standard Gibbs energy of formation of an element in a phase different from its reference state is nonzero: C(s, graphite) → C(s, diamond)

D f G 3(C, diamond) = +2.90 kJ mol−1

Many of the values in the tables have been compiled by combining the standard enthalpy of formation of the species with the standard entropies of the compound and the elements, as illustrated in Example 4.4, but there are other sources of data and we encounter some of them later. Standard Gibbs energies of formation can be combined to obtain the standard Gibbs energy of almost any reaction. We use the now familiar expression

147

148

4 CHEMICAL EQUILIBRIUM

Table 4.2

Standard Gibbs energies of formation at 298.15 K* D f G 9/(kJ mol−1)

Substance Gases Carbon dioxide, CO2

−394.36

Methane, CH4

−50.72

Nitrogen oxide, NO

+86.55

Water, H2O

−228.57

Liquids Ethanol, CH3CH2OH

−174.78

Hydrogen peroxide, H2O2

−120.35

Water, H2O

−237.13

Solids α-d-Glucose C6H12O6 Glycine, CH2(NH2)COOH Sucrose, C12H22O11 Urea, CO(NH2)2

−917.2 −532.9 −1543 −197.33

Solutes in aqueous solution Carbon dioxide, CO2

−385.98

Carbonic acid, H2CO3

−623.08

Phosphoric acid, H3PO4

−1018.7

*Additional values are given in the Data section.

DrG 3 =

∑ nD f G 3 − Reactants ∑ nD f G 3 Products

Calculation of standard Gibbs energy of reaction

(4.16)

If we need the biological standard reaction Gibbs energy, we convert DrG 3 to DrG⊕ by using eqn 4.8. A brief illustration

To determine the standard reaction Gibbs energy for the complete oxidation of solid sucrose, C12H22O11(s), by oxygen gas to carbon dioxide gas and liquid water, C12H22O11(s) + 12 O2(g) → 12 CO2(g) + 11 H2O(l) we carry out the following calculation: DrG 3 = {12D f G 3(CO2,g) + 11D f G 3(H2O,l)} − {Df G 3(C12H22O11,s) + 12D f G 3(O2,g)} = {12(−394 kJ mol−1) + 11(−237 kJ mol−1)} − {−1543 kJ mol−1 + 0} = −5.79 × 103 kJ mol−1 Calculate the standard reaction Gibbs energy of the oxidation of ammonia to nitric oxide according to the equation 4 NH3(g) + 5 O2(g) → 4 NO(g) + 6 H2O(g). Self-test 4.7

Answer: −959.42 kJ mol−1

4.4 THE STANDARD REACTION GIBBS ENERGY

149

(b) Stability and instability

Standard Gibbs energies of formation of compounds have their own significance as well as being useful in calculations of K. They are a measure of the ‘thermodynamic altitude’ of a compound above or below a ‘sea level’ of stability represented by the elements in their reference states (Fig. 4.10). If the standard Gibbs energy of formation is positive and the compound lies above ‘sea level’, then the compound has a spontaneous tendency to sink toward thermodynamic sea level and decompose into the elements. That is, K < 1 for their formation reaction. We say that a compound with D f G 3 > 0 is thermodynamically unstable with respect to its elements or that it is endergonic. Thus, the endergonic substance ozone, for which D f G 3 = +163 kJ mol−1, has a spontaneous tendency to decompose into oxygen under standard conditions at 25°C. More precisely, the equilibrium constant for the reaction 32 O2(g) 7 O3(g) is less than 1 (much less, in fact: K = 2.7 × 10−29). However, although ozone is thermodynamically unstable, it can survive if the reactions that convert it into oxygen are slow. That is the case in the upper atmosphere, and the O3 molecules in the ozone layer survive for long periods. Benzene (D f G 3 = +124 kJ mol−1) is also thermodynamically unstable with respect to its elements (K = 1.8 × 10−22). However, the fact that bottles of benzene are everyday laboratory commodities also reminds us of the point made at the start of the chapter, that spontaneity is a thermodynamic tendency that might not be realized at a significant rate in practice. Another useful point that can be made about standard Gibbs energies of formation is that there is no point in searching for direct syntheses of a thermodynamically unstable compound from its elements (under standard conditions, at the temperature to which the data apply) because the reaction does not occur in the required direction: the reverse reaction, decomposition, is spontaneous. Endergonic compounds must be synthesized by alternative routes or under conditions for which their Gibbs energy of formation is negative and they lie beneath thermodynamic sea level. Compounds with D f G 3 < 0 (corresponding to K > 1 for their formation reactions) are said to be thermodynamically stable with respect to their elements or exergonic. Exergonic compounds lie below the thermodynamic sea level of the elements (under standard conditions). An example is the exergonic compound ethane, with D f G 3 = −33 kJ mol−1: the negative sign shows that the formation of ethane gas from its elements is spontaneous in the sense that K > 1 (in fact, K = 7.1 × 105 at 25°C).

The response of equilibria to the conditions In introductory chemistry, we meet the empirical rule of thumb known as Le Chatelier’s principle: When a system at equilibrium is subjected to a disturbance, the composition of the system adjusts so as to tend to minimize the effect of the disturbance. Le Chatelier’s principle is only a rule of thumb, and to understand why reactions respond as they do and to calculate the new equilibrium composition, we need to use thermodynamics. We need to keep in mind that some changes in conditions affect the value of D rG 3 and therefore of K (temperature is the only instance), whereas others change the consequences of K having a particular fixed value without changing the value of K (the pressure, for instance).

Fig. 4.10 The standard Gibbs energy of formation of a compound is like a measure of the compound’s altitude above or below sea level: compounds that lie above sea level have a spontaneous tendency to decompose into the elements (and to revert to sea level). Compounds that lie below sea level are stable with respect to decomposition into the elements.

150

4 CHEMICAL EQUILIBRIUM

4.5 The presence of a catalyst Enzymes are biological versions of catalysts and are so ubiquitous that we need to know how their action affects chemical equilibria.

We study the action of catalysts (a substance that accelerates a reaction without itself appearing in the overall chemical equation), especially enzymes, in Chapter 8 and at this stage do not need to know in detail how they work other than that they provide an alternative, faster route from reactants to products. Although the new route from reactants to products is faster, the initial reactants and the final products are the same. The quantity DrG 3 is defined as the difference of the standard molar Gibbs energies of the reactants and products, so it is independent of the path linking the two. It follows that an alternative pathway between reactants and products leaves DrG 3 and therefore K unchanged. That is, the presence of a catalyst does not change the equilibrium constant of a reaction. 4.6 The effect of temperature In organisms, biochemical reactions occur over a very narrow range of temperatures, and changes by only a few degrees can have serious consequences, including death. Therefore, it is important to know how changes in temperature, such as those brought about by infections, affect biological processes.

According to Le Chatelier’s principle, we can expect a reaction to respond to a lowering of temperature by releasing heat and to respond to an increase of temperature by absorbing heat. That is: When the temperature is raised, the equilibrium composition of an exothermic reaction will tend to shift toward reactants; the equilibrium composition of an endothermic reaction will tend to shift toward products. In each case, the response tends to minimize the effect of raising the temperature. But why do reactions at equilibrium respond in this way? Le Chatelier’s principle is only a rule of thumb and gives no clue to the reason for this behavior. As we shall now see, the origin of the effect is the dependence of DrG 3, and therefore of K, on the temperature. First, we consider the effect of temperature on D rG 3. We use the relation D rG 3 = D r H 3 − TD r S 3 and make the assumption that neither the reaction enthalpy nor the reaction entropy varies much with temperature (over small ranges, at least). It follows that change in D rG 3 = −(change in T) × Dr S 3

(4.17)

This expression is easy to apply when there is a consumption or formation of gas because, as we have seen (Section 2.5), gas formation dominates the sign of the reaction entropy. Now consider the effect of temperature on K itself. At first, this problem looks troublesome because both T and DrG 3 appear in the expression for K. However, as we show in the following Justification, the effect of temperature can be expressed very simply as the van ’t Hoff equation.2 ln K2 = ln K1 +

Dr H 3 A 1 1 D − R C T1 T2 F

van ’t Hoff equation

(4.18)

2 There are several ‘van ’t Hoff equations’. To distinguish them, this one is sometimes called the van ’t Hoff isochore.

4.6 THE EFFECT OF TEMPERATURE

151

where K1 is the equilibrium constant at the temperature T1 and K2 is its value when the temperature is T2. All we need to know to calculate the temperature dependence of an equilibrium constant, therefore, is the standard reaction enthalpy. Justification 4.1 The van ’t Hoff equation

As before, we use the approximation that the standard reaction enthalpy and entropy are independent of temperature over the range of interest, so the entire temperature dependence of D rG 3 stems from the T in DrG 3 = D r H 3 − TD r S 3. At a temperature T1, ln K1 = −

DrG 3 D H3 D S3 =− r + r RT1 RT1 R

At another temperature T2, when DrG 3′ = Dr H 3 − T2Dr S 3 and the equilibrium constant is K2, a similar expression holds: ln K2 = −

Dr H 3 Dr S 3 + RT2 R

The difference between the two is eqn 4.18. Let’s explore the information in the van ’t Hoff equation. Consider the case when T2 > T1. Then the term in parentheses in eqn 4.18 is positive. If D r H 3 > 0, corresponding to an endothermic reaction, the entire term on the right is positive. In this case, therefore, ln K2 > ln K1. That being so, we conclude that K2 > K1 for an endothermic reaction. In general, the equilibrium constant of an endothermic reaction increases with temperature. The opposite is true when Dr H 3 < 0, so we can conclude that the equilibrium constant of an exothermic reaction decreases with an increase in temperature. Statistical principles also give us insight into the temperature dependence of the equilibrium constant. The typical arrangement of energy levels for an endothermic reaction is shown in Fig. 4.11a. When the temperature is increased, the Boltzmann distribution adjusts and the populations change as shown. The change corresponds to an increased population of the higher energy states at the expense of the population of the lower-energy states. We see that the states that arise from the B molecules become more populated at the expense of the A molecules. Therefore, the total population of B states increases, and B becomes more abundant in the equilibrium mixture. Conversely, if the reaction is exothermic (Fig. 4.11b), then an increase in temperature increases the population of the A states (which start at higher energy) at the expense of the B states, so the reactants become more abundant.

Coupled reactions in bioenergetics We remarked in the introduction to this chapter that thermodynamics enables us to determine whether one reaction can drive another forward. We now have enough information to take this step. A simple mechanical analogy is a pair of weights joined by a string (Fig. 4.12): the lighter of the pair of weights will be pulled up as the heavier weight falls down. Although the lighter weight has a

Fig. 4.11 The effect of temperature on a chemical equilibrium can be interpreted in terms of the change in the Boltzmann distribution with temperature and the effect of that change in the population of the species. (a) In an endothermic reaction, the population of B increases at the expense of A as the temperature is raised. (b) In an exothermic reaction, the opposite happens.

If two weights are coupled as shown here, then the heavier weight will move the lighter weight in its nonspontaneous direction: overall, the process is still spontaneous. The weights are the analogs of two chemical reactions: a reaction with a large negative DrG can force another reaction with a smaller DrG to run in its nonspontaneous direction. Fig. 4.12

152

4 CHEMICAL EQUILIBRIUM

natural tendency to move downward, its coupling to the heavier weight results in it being raised. The thermodynamic analog is an endergonic reaction, a reaction with a positive Gibbs energy, DrG (the analog of the lighter weight moving up), being forced to occur by coupling it to an exergonic reaction, a reaction with a negative Gibbs energy, DrG′ (the analog of the heavier weight falling down). The overall reaction is spontaneous because the sum DrG + D rG′ is negative. The whole of life’s activities depend on couplings of this kind, for the oxidation reactions of food act as the heavy weights that drive other reactions forward and result in the formation of proteins from amino acids, the actions of muscles for propulsion, and even the activities of the brain for reflection, learning, and imagination.

Case study 4.2

ATP and the biosynthesis of proteins

The function of adenosine triphosphate, ATP4− (Atlas N3) or (more succinctly) ATP, is to store the energy made available when food is oxidized and then to supply it on demand to a wide variety of processes, including muscular contraction, reproduction, and vision. We saw in Case study 2.2 that the essence of ATP’s action is its ability to lose its terminal phosphate group by hydrolysis and to form adenosine diphosphate, ADP3− (Atlas N2): ATP4−(aq) + H2O(l) → ADP3−(aq) + HPO 42−(aq) + H3O+(aq) This reaction is exergonic under the conditions prevailing in cells and can drive an endergonic reaction forward if suitable enzymes are available to couple the reactions. One reason why ATP is so potent is that its concentration in cells is high, so its chemical potential is also high. The biological standard values for the hydrolysis of ATP at 37°C are D rG ⊕ = −31 kJ mol−1

Dr H ⊕ = −20 kJ mol−1

D r S ⊕ = +34 J K−1 mol−1

The hydrolysis is therefore exergonic (D rG < 0) under these conditions, and 31 kJ mol−1 is available for driving other reactions. On account of its exergonic character, the ADP–phosphate bond has been called a ‘high-energy phosphate bond’. The name is intended to signify a high tendency to undergo reaction and should not be confused with ‘strong’ bond in its normal chemical sense (that of a high bond enthalpy). In fact, even in the biological sense it is not of very ‘high energy’. The action of ATP depends on the bond being intermediate in strength. Thus ATP acts as a phosphate donor to a number of acceptors (such as glucose) but is recharged with a new phosphate group by more powerful phosphate donors in the phosphorylation steps in the respiration cycle. In the cell, each ATP molecule can be used to drive an endergonic reaction for which DrG 3 does not exceed 31 kJ mol−1. For example, the biosynthesis of sucrose from glucose and fructose can be driven by enzyme-catalyzed processes in plants because the reaction is endergonic to the extent DrG ⊕ = +23 kJ mol−1. The biosynthesis of proteins is strongly endergonic, not only on account of the enthalpy change but also on account of the large decrease in entropy that occurs when many amino acid residues are assembled into a precisely determined sequence. For instance, the formation of a peptide link is endergonic, with DrG ⊕ = +17 kJ mol−1, but the biosynthesis occurs indirectly and is equivalent to the consumption of three ATP molecules for each link.

4.6 THE EFFECT OF TEMPERATURE

In a moderately small protein such as myoglobin, with about 150 peptide links, the construction alone requires 450 ATP molecules and therefore about 12 mol of glucose molecules for 1 mol of protein molecules. Fats yield almost twice as much energy per gram as carbohydrates. What mass of fat would need to be metabolized to synthesize 1.0 mol of myoglobin molecules? Self-test 4.8

Answer: 1.1 kg

Adenosine triphosphate is not the only phosphate species capable of driving other less exergonic reactions. For instance, creatine phosphate (3) can release its phosphate group in a hydrolysis reaction, and D rG ⊕ = −43 kJ mol−1. These different exergonicities give rise to the concept of transfer potential, which is the negative of the value of DrG ⊕ for the hydrolysis reaction. Thus, the transfer potential of creatine phosphate is +43 kJ mol−1. Just as one exergonic reaction can drive a less exergonic reaction, so the hydrolysis of a species with a high transfer potential can drive the phosphorylation of a species with a lower transfer potential (Table 4.3).

Table 4.3

Transfer potentials at 298.15 K

Substance

Transfer potential, −D rG 9/(kJ mol−1)

AMP

14

ATP, ADP

31

1,3-Bis(phospho)glycerate

49

Creatine phosphate

43

Glucose-6-phosphate

14

Glycerol-1-phosphate

10

Phosphoenolpyruvate Pyrophosphate, HP2O

Case study 4.3

62 3− 7

33

The oxidation of glucose

The breakdown of glucose in the cell begins with glycolysis, a partial oxidation of glucose by nicotinamide adenine dinucleotide (NAD+, Atlas N4) to pyruvate ion, CH3COCO2− (4). Metabolism continues in the form of the citric acid cycle, in which pyruvate ions are oxidized to CO2, and ends with oxidative phosphorylation, in which O2 is reduced to H2O. Glycolysis is the main source of energy during anaerobic metabolism, a form of metabolism in which inhaled O2 does not play a role. The citric acid cycle and oxidative phosphorylation are the main mechanisms for the extraction of energy from carbohydrates during aerobic metabolism, a form of metabolism in which inhaled O2 does play a role. Glycolysis occurs in the cytosol, the aqueous material encapsulated by the cell membrane, and consists of 10 enzyme-catalyzed reactions (Fig. 4.13).

153

154

4 CHEMICAL EQUILIBRIUM

The reactions of glycolysis, in which glucose is partially oxidized by nicotinamide adenine dinucleotide (NAD+, Atlas N4) to pyruvate ion.

Fig. 4.13

A brief comment

From now on, we shall represent biochemical reactions with chemical equations written with a shorthand method, in which some substances are given ‘nicknames’ and charges are not always given explicitly. For example, H2PO 2− 4 is written as Pi, ATP4− as ATP, and so on. We need to show hydrogen ions explicitly (because they account for differences between thermodynamic and biological standard states and for the role of pH); in such cases charges will not seem to be balanced.

The process needs to be initiated by consumption of two molecules of ATP per molecule of glucose. The first ATP molecule is used to drive the phosphorylation of glucose to glucose-6-phosphate (G6P): glucose(aq) + ATP(aq) → G6P(aq) + ADP(aq) + H+(aq) DrG ⊕ = −17 kJ mol−1 (Note that ATP, ADP, and G6P denote charged species, so charges are in fact balanced in this and similar equations, but charge balance is not displayed explicitly.) As we saw in Section 4.1, the next step is the isomerization of G6P to fructose-6-phosphate (F6P). The second ATP molecule consumed during glycolysis drives the phosphorylation of F6P to fructose-1,6-diphosphate (FDP): F6P(aq) + ATP(aq) → FDP(aq) + ADP(aq) + H+(aq) DrG ⊕ = −14 kJ mol−1 In the next step, FDP is broken into two three-carbon units, dihydroxyacetone phosphate (1,3-dihydroxypropanone phosphate, CH2OHCOCH2OPO32−, 5) and glyceraldehyde-3-phosphate (6), which exist in mutual equilibrium. Only the glyceraldehyde-3-phosphate is oxidized by NAD+ to pyruvate ion, with formation of two ATP molecules. As glycolysis proceeds, all the dihydroxyacetone phosphate is converted to glyceraldehyde-3-phosphate, so the result is the consumption of two NAD+ molecules and the formation of four ATP molecules per molecule of glucose.

4.6 THE EFFECT OF TEMPERATURE

155

The oxidation of glucose by NAD+ to pyruvate ions has DrG ⊕ = −147 kJ mol−1 at blood temperature. In glycolysis, the oxidation of one glucose molecule is coupled to the net conversion of two ADP molecules to two ATP molecules (‘net’ because two ATP molecules are consumed and four are formed), so the net reaction of glycolysis is glucose(aq) + 2 NAD+(aq) + 2 ADP(aq) + 2 Pi(aq) + 2 H2O(l) → 2 CH3COCO2−(aq) + 2 NADH(aq) + 2 ATP(aq) + 2 H3O+(aq) The biological standard reaction Gibbs energy is (−147) −2(−31) kJ mol−1 = −85 kJ mol−1. The reaction is exergonic and therefore spontaneous under biological standard conditions: the oxidation of glucose is used to ‘recharge’ the ATP. In cells that are deprived of O2, pyruvate ion is reduced to lactate ion, CH3CH(OH)CO2− (7) by NADH.3 Very strenuous exercise, such as bicycle racing, can decrease sharply the concentration of O2 in muscle cells, and the condition known as muscle fatigue results from increased concentrations of lactate ion. The standard Gibbs energy of combustion of glucose is −2880 kJ mol−1, so terminating its oxidation at pyruvate is a poor use of resources, akin to the partial combustion of hydrocarbon fuels in a badly tuned engine. In the presence of O2, pyruvate is oxidized further during the citric acid cycle and oxidative phosphorylation, which occur in the mitochondria of cells. The further oxidation of carbon derived from glucose begins with a reaction between pyruvate ion, NAD+, and coenzyme A (CoA, Atlas N6) to give acetyl CoA, NADH, and CO2. Acetyl CoA is then oxidized by NAD+ and flavin adenine dinucleotide (FAD, Atlas N7) in the citric acid cycle (Fig. 4.14), which

The reactions of the citric acid cycle, in which acetyl CoA is oxidized by NAD+ and FAD, resulting in the synthesis of GTP (shown) or ATP, depending on the type of cell. The GTP molecules are eventually converted to ATP.

Fig. 4.14

3

In yeast, the terminal products are ethanol and CO2.

156

4 CHEMICAL EQUILIBRIUM

requires eight enzymes and results in the synthesis of guanosine triphosphate (GTP, Atlas N8) from guanosine diphosphate, GDP, or of ATP from ADP: Acetyl CoA(aq) + 3 NAD+(aq) + FAD(aq) + GDP(aq) + Pi(aq) + 2 H2O(l) → 2 CO2(g) + 3 NADH(aq) + 2 H3O+(aq) + FADH2(aq) + GTP(aq) + CoA(aq) D rG ⊕ = −57 kJ mol−1 In cells that produce GTP, the enzyme nucleoside diphosphate kinase catalyzes the transfer of a phosphate group to ADP to form ATP: GTP(aq) + ADP(aq) → GDP(aq) + ATP(aq) For this reaction, DrG ⊕ = 0 because the phosphate group transfer potentials for GTP and ATP are essentially identical. Overall, we write the oxidation of glucose as a result of glycolysis and the citric acid cycle as glucose(aq) + 10 NAD+(aq) + 2 FAD(aq) + 4 ADP(aq) + 4 Pi(aq) + 2 H2O(l) → 6 CO2(g) + 10 NADH(aq) + 6 H3O+(aq) + 2 FADH2(aq) + 4 ATP(aq) The NADH and FADH2 go on to reduce O2 during oxidative phosphorylation (Section 5.10b), which also produces ATP. The citric acid cycle and oxidative phosphorylation generate as many as 38 ATP molecules for each glucose molecule consumed. Each mole of ATP molecules extracts 31 kJ from the 2880 kJ supplied by 1 mol C6H12O6 (180 g of glucose), so 1178 kJ is stored for later use. Therefore, aerobic oxidation of glucose is much more efficient than glycolysis.

Proton transfer equilibria An enormously important biological aspect of chemical equilibrium is that involving the transfer of protons (hydrogen ions, H+) between species in aqueous environments, such as living cells. Even small drifts in the equilibrium concentration of hydrogen ions can result in disease, cell damage, and death. In this section we see how the general principles outlined earlier in the chapter are applied to proton transfer equilibria. 4.7 Brønsted–Lowry theory Cells have elaborate procedures for using proton transfer equilibria, and this function cannot be understood without knowing which species provide protons and which accept them and how to express the concentration of hydrogen ions in solution.

According to the Brønsted–Lowry theory of acids and bases, an acid is a proton donor and a base is a proton acceptor. The proton, which in this context means a hydrogen ion, H+, is highly mobile and acids and bases in water are always in equilibrium with their deprotonated and protonated counterparts and hydronium ions (H3O+, 8). Thus, an acid HA, such as HCN, immediately establishes the equilibrium HA(aq) + H2O(l) 7 H3O+(aq) + A−(aq)

K=

aH O+ aA aHAaH O

3

2

(4.19a)

4.8 PROTONATION AND DEPROTONATION

A base B, such as NH3, immediately establishes the equilibrium B(aq) + H2O(l) 7 HB+(aq) + OH−(aq)

K=

aHB aOH aBaH O +

(4.19b)

2

In these equilibria, A− is the conjugate base of the acid HA, and BH+ is the conjugate acid of the base B. Even in the absence of added acids and bases, proton transfer occurs between water molecules, and the autoprotolysis equilibrium4 2 H2O(l) 7 H3O+(aq) + OH−(aq) K =

aH O aOH aH2 O 3

+

Autoprotolysis equilibrium

(4.20)

2

is always present. As will be familiar from introductory chemistry, the hydronium ion concentration is commonly expressed in terms of the pH, which is defined formally as pH = −log aH O 3

Definition of pH

+

(4.21)

where the logarithm is to base 10. In elementary work, the hydronium ion activity is replaced by the numerical value of its molar concentration, [H3O+], which is equivalent to setting the activity coefficient g equal to 1. A brief illustration

If the molar concentration of H3O+ is 2.0 mmol dm−3 (where 1 mmol = 10−3 mol), then pH ≈ − log(2.0 × 10−3) = 2.70 If the molar concentration were 10 times less, at 0.20 mmol dm−3, then the pH would be 3.70.

Notice that the higher the pH, the lower the concentration of hydronium ions in the solution and that a change in pH by 1 unit corresponds to a 10-fold change in their molar concentration. However, it should never be forgotten that the replacement of activities by molar concentration is invariably hazardous. Because ions interact over long distances, the replacement is unreliable for all but the most dilute solutions. Death is likely if the pH of human blood plasma changes by more than ±0.4 from its normal value of 7.4. What is the approximate range of molar concentrations of hydrogen ions for which life can be sustained?

Self-test 4.9

Answer: 16 nmol dm−3 to 100 nmol dm−3 (1 nmol = 10 −9 mol)

4.8 Protonation and deprotonation The protonation and deprotonation of molecules are key steps in many biochemical reactions, and we need to be able to describe procedures for treating protonation and deprotonation processes quantitatively.

All the solutions we consider are so dilute that we can regard the water present as being a nearly pure liquid and therefore as having unit activity (see Table 3.3). 4

Autoprotolysis is also called autoionization.

157

158

4 CHEMICAL EQUILIBRIUM

This feature leads to convenient expressions for quantities that measure the strengths of acids and bases and the extent of protonation of bases and deprotonation of acids. (a) The strengths of acids and bases

When we set aH O = 1 for all the solutions we consider, the resulting equilibrium constant is called the acidity constant, Ka, of the acid HA:5 2

HA(aq) + H2O(l) 7 H3O+(aq) + A−(aq) a a Ka = H O A aHA +

3

Definition of acidity constant

(4.22a)

In elementary applications, the activities are replaced by the numerical values of the molar concentrations, and we write Ka =

[H3O+][A−] [HA]

(4.22b)

Data are widely reported in terms of the negative common (base 10) logarithm of this quantity: pKa = −log Ka

Definition of pKa

(4.23)

It follows from eqn 4.10a (D rG = −RT ln K) that pK a is proportional to DrG 3 for the proton transfer reaction. More explicitly, pK a = D rG 3/(RT ln 10), with ln 10 = 2.303. . . . Therefore, manipulations of pK a and related quantities are actually manipulations of standard reaction Gibbs energies in disguise. 3

Self-test 4.10 Show that pK a = D rG 3/(RT ln 10). Hint: ln x = ln 10 × log x.

The value of the acidity constant indicates the extent to which proton transfer occurs at equilibrium in aqueous solution. The smaller the value of Ka (for instance 10−8 compared with 10−6) and therefore the larger the value of pK a (for instance, 8 compared with 6), the lower is the concentration of deprotonated molecules. Most acids have K a < 1 (and usually much less than 1), with pK a > 0, indicating only a small extent of deprotonation in water. These acids are classified as weak acids. A few acids, most notably, in aqueous solution, HCl, HBr, HI, HNO3, H2SO4, and HClO4, are classified as strong acids and are commonly regarded as being completely deprotonated in aqueous solution.6 The corresponding expression for a base is called the basicity constant, K b : B(aq) + H2O(l) 7 HB+(aq) + OH−(aq) a a K b = HB OH aB +

Definition of basicity constant

(4.24a)

and the corresponding value of pK b = −log K b. As for acids, in elementary applications the activities are replaced by the numerical values of the molar concentrations and we use 5 Acidity constants are also called acid ionization constants and, less appropriately, dissociation constants. 6 Sulfuric acid, H2SO4, is strong with respect only to its first deprotonation; HSO −4 is weak.

4.8 PROTONATION AND DEPROTONATION

Kb =

[HB+][OH−] [B]

(4.24b)

A strong base is fully protonated in solution in the sense that K b > 1. One example is the oxide ion, O2−, which cannot survive in water but is immediately and fully converted into its conjugate acid OH−. A weak base is not fully protonated in water, in the sense that K b < 1 (and usually much less than 1). Ammonia, NH3, and its organic derivatives the amines are all weak bases in water, and only a small proportion of their molecules exist as the conjugate acid (NH 4+ or RNH 3+). The autoprotolysis constant for water, K w, is obtained in a similar way by setting the activity of water in eqn 4.19 to its ‘pure’ value: Definition of the autoprotolysis constant of water

K w = aH O aOH ≈ [H3O+][OH−] 3

+

(4.25)

At 25°C, K w = 1.0 × 10−14 and pK w = −log K w = 14.00. As may be confirmed by multiplying the two constants together, the basicity constant of a base B and the acidity constant of its conjugate acid, HB+, B(aq) + H2O(l) 7 HB+(aq) + OH−(aq)

Kb =

aHB aOH aB

HB+(aq) + H2O(l) 7 H3O+(aq) + B(aq)

Ka =

aH O aB aHB

+

3

+

+

are related by Ka K b =

aH O aB aHB aOH × = aH O aOH = K w aHB aB 3

+

+

+

3

+

Relation between Ka and Kb

(4.26a)

The implication of this relation is that K a increases as K b decreases to maintain a product equal to the constant K w. That is, as the strength of a base decreases, the strength of its conjugate acid increases and vice versa. On taking the negative common logarithm of both sides of eqn 4.26a, we obtain pKa + pK b = pKw

Relation between pKa and pKb

(4.26b)

The great advantage of this relation is that the pKb values of bases may be expressed as the pK a of their conjugate acids, so the strengths of all weak acids and bases may be listed in a single table (Table 4.4). A brief illustration

If the acidity constant of the conjugate acid (CH3NH3+) of the base methylamine (CH3NH2) is reported as pKa = 10.56, we can infer that the basicity constant of methylamine itself is pK b = pK w − pKa = 14.00 − 10.56 = 3.44

Another useful relation is obtained by taking the negative common logarithm of both sides of the definition of Kw in eqn 4.24, which gives pH + pOH = pK w

Relation between pH and pOH

(4.27)

159

160

4 CHEMICAL EQUILIBRIUM

Table 4.4

Acidity and basicity constants* at 298.15 K

Acid/base

Kb

pK b

Ka

pK a

Trichloroacetic acid, CCl3COOH

3.3 × 10−14

13.48

3.0 × 10−1

0.52

Benzenesulfonic acid, C6H5SO3H

5.0 × 10−14

13.30

2 × 10−1

0.70

Iodic acid, HIO3

5.9 × 10

13.23

Sulfurous acid, H2SO3

−13

6.3 × 10

Chlorous acid, HClO2

1.0 × 10−12

Strongest weak acids

1.7 × 10

−1

0.77

12.19

1.6 × 10

−2

1.81

12.00

1.0 × 10−2

2.00

Phosphoric acid, H3PO4

−12

1.3 × 10

11.88

7.6 × 10

−3

2.12

Chloroacetic acid, CH2ClCOOH

7.1 × 10−12

11.15

1.4 × 10−3

2.85

Lactic acid, CH3CH(OH)COOH

1.2 × 10−11

10.92

8.4 × 10−4

3.08

Nitrous acid, HNO2

−11

2.3 × 10

10.63

4.3 × 10

−4

3.37

Hydrofluoric acid, HF

2.9 × 10−11

10.55

3.5 × 10−4

3.45

Formic acid, HCOOH

5.6 × 10−11

10.25

1.8 × 10−4

3.75

−4

4.19

−14

Benzoic acid, C6H5COOH

−10

1.5 × 10

9.81

6.5 × 10

Acetic acid, CH3COOH

5.6 × 10−10

9.25

1.8 × 10−4

4.75

Carbonic acid, H2CO3

2.3 × 10−8

7.63

4.3 × 10−7

6.37

Hypochlorous acid, HClO

−7

3.3 × 10

6.47

3.0 × 10

−8

7.53

Hypobromous acid, HBrO

5.0 × 10−6

5.31

2.0 × 10−9

8.69

Boric acid, B(OH)3H†

1.4 × 10−5

4.86

7.2 × 10−10

9.14

Hydrocyanic acid, HCN

−5

2.0 × 10

4.69

4.9 × 10

9.31

Phenol, C6H5OH

7.7 × 10−5

4.11

1.3 × 10−10

Hypoiodous acid, HIO

4.3 × 10

3.36

2.3 × 10

−4

−10

−11

9.89 10.64

Weakest weak bases 7.7 × 10−1

0.10

9.37

2.3 × 10

−5

4.63

8.75

5.6 × 10−6

5.35

7.97

9.1 × 10−7

6.03

Nicotine, C10H11N2

−6

1.0 × 10

5.98

1.0 × 10

−8

8.02

Morphine, C17H19O3N

1.6 × 10−6

5.79

6.3 × 10−9

8.21

Hydrazine, NH2NH2

1.7 × 10−6

5.77

5.9 × 10−9

8.23

Ammonia, NH3

−5

1.8 × 10

4.75

5.6 × 10

9.25

Trimethylamine, (CH3)3N

6.5 × 10−5

4.19

1.5 × 10−10

Methylamine, CH3NH2

−4

3.6 × 10

3.44

2.8 × 10

−11

10.56

Dimethylamine, (CH3)2NH

5.4 × 10−4

3.27

1.9 × 10−11

10.73

Ethylamine, C2H5NH2

6.5 × 10−4

3.19

1.5 × 10−11

10.81

Triethylamine, (C2H5)3N

1.0 × 10−3

2.99

1.0 × 10−11

11.01

Urea, CO(NH2)2

1.3 × 10−14

Aniline, C6H5NH2

4.3 × 10

Pyridine, C5H5N

1.8 × 10−9

Hydroxylamine, NH2OH

1.1 × 10−8

−10

13.90

−10

Strongest weak bases *Values for polyprotic acids—those capable of donating more than one proton—refer to the first deprotonation. † The proton transfer equilibrium is B(OH)3(aq) + 2 H2O(l) 7 H3O+(aq) + B(OH)−4 (aq).

9.81

4.8 PROTONATION AND DEPROTONATION where pOH = −log aOH . This enormously important relation means that the activities (in elementary work, the molar concentrations) of hydronium and hydroxide ions of a given solution are related by a seesaw relation: as one goes up, the other goes down to preserve the value of pKw. −

Self-test 4.11 The molar concentration of OH− ions in a certain solution

is 0.010 mmol dm−3. What is the pH of the solution? Answer: 9.00

(b) The pH of a solution of a weak acid

The most reliable way to estimate the pH of a solution of a weak acid is to consider the contributions from deprotonation of the acid and autoprotolysis of water to the total concentration of hydronium ion in solution (see Further information 4.1). Autoprotolysis may be ignored if the weak acid is the main contributor of hydronium ions, a condition that is satisfied if the acid is not very weak and is present at not too low a concentration. Then we can estimate the pH of a solution of a weak acid and calculate either of these fractions by using the following strategy: Organize the necessary work into a table with columns headed by the species present in the mixture (ignoring H2O) and, in successive rows write: 1. The initial molar concentrations of the species, ignoring any contributions to the concentration of H3O+ or OH− from the autoprotolysis of water. 2. The changes in these quantities that must take place for the system to reach equilibrium. 3. The resulting equilibrium values. In most cases, we do not know the change that must occur for the system to reach equilibrium, so the change in the concentration of H3O+ is written as x and the reaction stoichiometry is used to write the corresponding changes in the other species. When the values at equilibrium (the last row of the table) are substituted into the expression for the acidity constant, we obtain an equation for x in terms of Ka. This equation can be solved for x. In general, solution of the equation for x results in several mathematically possible values of x. We select the chemically acceptable solution by considering the signs of the predicted concentrations: they must be positive. Example 4.5

Estimating the pH of a solution of a weak acid

Acetic acid lends a sour taste to vinegar and is produced by aerobic oxidation of ethanol by bacteria in fermented beverages, such as wine and cider: CH3CH2OH(aq) + O2(g) → CH3COOH(aq) + H2O(l). Estimate the pH of (a) 0.15 m CH3COOH(aq) and (b) 1.5 × 10−4 m CH3COOH(aq). Strategy Proceed as outlined above. Solution We draw up the following equilibrium table based on the proton

transfer equilibrium CH3COOH(aq) + H2O(l) 7 H3O+(aq) + CH3CO 2−(aq). Species Initial concentration/(mol dm−3) Change to reach equilibrium/(mol dm−3) Equilibrium concentration/(mol dm−3)

CH3COOH 0.15 −x 0.15 − x

H3O+ 0 +x x

CH3CO 2− 0 +x x

161

162

4 CHEMICAL EQUILIBRIUM

(a) The value of x is found by inserting the equilibrium concentrations into the expression for the acidity constant: Ka =

A note on good practice

When an approximation has been made, verify at the end of the calculation that the approximation is consistent with the result obtained. In this case, we assumed that x Ka1), where S is the numerical

Distinguish between amphiprotic, which means that a species can both accept and donate protons, and amphoteric, which means that a substance can react with both an acid and a base. Aluminum is amphoteric but not amphiprotic.

170

4 CHEMICAL EQUILIBRIUM

Fig. 4.17 The fractional composition of the protonated and deprotonated forms of an amino acid NH2CHRCOOH (with arbitrarily chosen values of pK), in which the group R does not participate in proton transfer reactions.

value of the molar concentration of the salt providing the anion), the pH of such a solution is given by The pH of the solution of an amphiprotic anion

pH = 12 (pKa1 + pKa2)

(4.34)

A brief illustration

Using values from Table 4.5, we can immediately conclude that the pH of the solution of sodium hydrogencarbonate of any concentration (subject to the conditions just quoted) is pH = 12 (6.37 + 10.25) = 8.31 The solution is basic. We can treat a solution of potassium hydrogenphosphate in the same way, taking into account only the second and third acidity constants of H3PO4 because protonation as far as H3PO4 is negligible (see Table 4.5): pH = 12 (7.21 + 12.67) = 9.94

4.11 Buffer solutions Cells cease to function and may be damaged irreparably if the pH changes significantly, so we need to understand how the pH is stabilized by a buffer.

Suppose that we make an aqueous solution by dissolving known amounts of a weak acid (which provides the species HA) and its conjugate base (which provides the species A−). To calculate the pH of this solution, we make use of the expression for K a of the weak acid, eqn 4.22, with [HA] = [acid] and [A−] = [base], and write Ka =

aH O abase aH O [base] ≈ aacid [acid] 3

+

3

+

which rearranges first to aH O ≈ 3

+

Ka[acid] [base]

4.11 BUFFER SOLUTIONS

and then, by taking negative common logarithms, to the Henderson–Hasselbalch equation: pH = pKa − log

[acid] [base]

Henderson–Hasselbalch equation:

(4.35)

When the concentrations of the conjugate acid and base are equal, the second term on the right of eqn 4.35 is log 1 = 0, so under these conditions pH = pK a. Although the equation has been derived without making any assumptions about [acid] and [base], it is common to suppose that, because the acid is weak, [acid] and [base] are unchanged from the values used to make up the solution; that is, we disregard the small amount of deprotonation of the added acid and the small amount of protonation of the added base.

A brief illustration

To calculate the pH of a solution formed from equal amounts of CH3COOH(aq) and NaCH3CO2(aq), we note that the latter dissociates (in the sense that the ions separate) fully in water, yielding Na+(aq) and CH3CO2−(aq), the conjugate base of CH3COOH(aq). Because [CH3COOH] = [CH3CO2−] (that is, [acid] = [base]), for this solution provided we disregard protonation and deprotonation, pH ≈ pK a. Because the pKa of CH3COOH(aq) is 4.75 (Table 4.4), it follows that pH = 4.8 (more realistically, pH = 5). Self-test 4.18 Calculate the pH of an aqueous solution that contains equal

amounts of NH3 and NH4Cl. Answer: 9.25; more realistically: 9

It is observed that solutions containing known amounts of an acid and that acid’s conjugate base show buffer action, the ability of a solution to oppose changes in pH when small amounts of strong acids and bases are added. An acid buffer solution, one that stabilizes the solution at a pH below 7, is typically prepared by making a solution of a weak acid (such as acetic acid) and a salt that supplies its conjugate base (such as sodium acetate). A base buffer, one that stabilizes a solution at a pH above 7, is prepared by making a solution of a weak base (such as ammonia) and a salt that supplies its conjugate acid (such as ammonium chloride). Physiological buffers are responsible for maintaining the pH of blood within a narrow range of 7.37 to 7.43, thereby stabilizing the active conformations of biological macromolecules and optimizing the rates of biochemical reactions. An acid buffer stabilizes the pH of a solution because the abundant supply of A− ions (from the salt) can remove any H3O+ ions brought by additional strong acid; furthermore, the abundant supply of HA molecules (from the acid component of the buffer) can provide H3O+ ions to react with any strong base that is added. Similarly, in a base buffer the weak base B can accept protons when a strong acid is added and its conjugate acid BH+ can supply protons if a strong base is added. The following example explores the quantitative basis of buffer action.

171

172

4 CHEMICAL EQUILIBRIUM

Example 4.8

Assessing buffer action

Estimate the effect of addition of 0.020 mol of hydronium ions (from a solution of a strong acid, such as hydrochloric acid) on the pH of 1.0 dm3 of (a) 0.15 m CH3COOH(aq) and (b) a buffer solution containing 0.15 m CH3COOH(aq) and 0.15 m NaCH3CO2(aq). Strategy Before addition of hydronium ions, the pH of solutions (a) and (b) is 2.8 (Example 4.5) and 4.8 (see the preceding brief illustration). After addition to solution (a) the initial molar concentration of CH3COOH(aq) is 0.15 m and that of H3O+(aq) is (0.020 mol)/(1.0 dm3) = 0.020 m. After addition to solution (b), the initial molar concentrations of CH3COOH(aq), CH3CO2−(aq), and H3O+(aq) are 0.15 m, 0.15 m, and 0.020 m, respectively. The weak base already present in solution, CH3CO2−(aq), reacts immediately with the added hydronium ions:

CH3CO2−(aq) + H3O+(aq) → CH3COOH(aq) + H2O(l) We use the adjusted concentrations of CH3COOH(aq) and CH3CO2−(aq) and eqn 4.45 to calculate a new value of the pH of the buffer solution. Solution For addition of a strong acid to solution (a), we draw up the following

equilibrium table to show the effect of the addition of hydronium ions: Species Initial concentration/(mol dm−3) Change to reach equilibrium/(mol dm−3) Equilibrium concentration/(mol dm−3)

CH3COOH 0.15 −x 0.15 − x

H3O+ 0.020 +x 0.020 + x

CH3CO2− 0 +x x

The value of x is found by inserting the equilibrium concentrations into the expression for the acidity constant: Ka =

[H3O+][CH3CO2−] (0.020 + x)x = [CH3COOH] 0.15 − x

As in Example 4.5, we assume that x is very small; in this case x pKa.

6. The equilibrium constant of a reaction is independent of the presence of a catalyst.

13. The pH of a buffer solution containing equal concentrations of a weak acid and its conjugate base is pH = pKa.

7. The equilibrium constant K increases with temperature if D r H 3 > 0 (an endothermic reaction) and decreases if Dr H 3 < 0 (an exothermic reaction).

FURTHER INFORMATION

175

Checklist of key equations Property

Equation

Comment

Reaction quotient

Q = acC aDd /aAa abB

Dimensionless

Equilibrium constant

K = (a a /a a )

Dimensionless

Reaction Gibbs energy

D rG = DrG 3 + RT ln Q

Standard reaction Gibbs energy

D rG 3 =

c d C D

a b A B equilibrium

nD f G 3 −

Products

nD f G 3

Procedure for calculation

Reactants

Relation to K

D rG 3 = −RT ln K

van ’t Hoff equation

ln K2 = ln K1 + (Dr H 3/R)(1/T1 − 1/T2)

Relation between standard states

D rG = DrG + 7nRT ln 10

pH

pH = − log aH O

Relation between pH and pOH

pH + pOH = pKw

Acidity constant

K a = aH O aA /aHA

Dilute solutions (aH O = 1)

Basicity constant

Kb = aHB aOH /aB

Dilute solutions (aH O = 1)

Autoprotolysis constant

Kw = aH O aOH

Assumes Dr H 3 constant over range

3

3

+

3

+

3

Definition

+

2

+

2

Relation between pK a and pK b

pK a + pKb = pKw

pH of amphiprotic anion solution

pH = 12 (pKa1 + pK a2)

Henderson–Hasselbalch equation

pH = pKa − log[acid]/[base]

K a2 > Kw /K a2, and S >> K a1

Further information Further information 4.1 The contribution of autoprotolysis to pH

Some acids are so weak and undergo so little deprotonation that the autoprotolysis of water can contribute significantly to the pH. We must also take autoprotolysis into account when we find by using the procedures in Example 4.5 that the pH of a solution of a weak acid is greater than 6. We begin the calculation by noting that, apart from water, there are four species in solution: HA, A−, H3O+, and OH−. Because there are four unknown quantities, we need four equations to solve the problem. Two of the equations are the expressions for Ka and Kw (eqns 4.20 and 4.24), written here in terms of molar concentrations: [H O+][A−] Ka = 3 [HA]

Kw = [H3O+][OH−]

(4.36)

A third equation takes charge balance, the requirement that the solution be electrically neutral, into account. That is, the sum of the concentrations of the cations must be equal to the sum of the concentrations of the anions. In our case, the charge balance equation is [H3O+] = [OH−] + [A−]

(4.37)

We also know that the total concentration of A groups in all forms in which they occur, which we denote as A, must be equal

to the initial concentration of the weak acid. This condition, known as material balance, gives our final equation: A = [HA] + [A−]

(4.38)

Now we are ready to proceed with a calculation of the hydronium ion concentration in the solution. First, we combine eqns 4.36 and 4.37 and write [A−] = [H3O+] −

Kw [H3O+]

(4.39)

We continue by substituting this expression into eqn 4.39 and solving for [HA]: [HA] = A − [H3O+] +

Kw [H3O+]

(4.40)

On substituting the expressions for [A−] (eqn 4.39) and [HA] (eqn 4.40) into the first of eqn 4.36, we obtain

Ka =

A Kw D [H3O+] [H3O+] − C [H3O+]F A − [H3O+] +

Kw [H3O+]

(4.41)

Rearrangement of this expression gives [H3O+]3 + Ka[H3O+]2 − (Kw + Ka A)[H3O+] − Ka Kw = 0 (4.42)

176

4 CHEMICAL EQUILIBRIUM with [M+] = S. The difference of these two equations can be expressed as [H3O+] = [OH−] + [A2−] − [H2 A] =

Kw [HA−]Ka2 [HA−][H3O+] + − + [H3O ] [H3O+] Ka1

(4.46)

Multiplication through by [H3O+]Ka1 turns this expression into Ka1[H3O+]2 = Kw Ka1 + [HA−]K a1Ka2 − [H3O+]2[HA−] The orange lines are the exact solutions of eqn 4.41 for a series of values of A (expressed as 10−x, with −x displayed). The dotted lines are the corresponding approximate solutions obtained from eqn 4.42b. Fig. 4.18

and we see that [H3O+] is determined by solving this cubic equation, a task that is best accomplished with a calculator or mathematical software. Figure 4.18 summarizes the outcome. There are several experimental conditions that allow us to simplify eqn 4.42. For example, when K a A >> Kw and A[H3O+] >> K w it becomes [H3O+]2 + Ka[H3O+] − Ka A = 0

(4.43a)

which can be solved for [H3O+]. If the extent of deprotonation is very small, we let [H3O+] 0, the reverse reaction is spontaneous and Ecell < 0. At equilibrium D rG = 0 and therefore Ecell = 0 too. Equation 5.12 provides an electrical method for measuring a reaction Gibbs energy at any composition of the reaction mixture: we simply measure the zerocurrent cell potential and convert it to D rG. Conversely, if we know the value of D rG at a particular composition, then we can predict the cell potential. A brief illustration

Suppose D rG ≈ −1 × 102 kJ mol−1 and n = 1; then Ecell =

−D rG −(−1 × 105 J mol−1) = =1V nF 1 × (9.6485 × 105 C mol−1)

Most electrochemical cells bought commercially are indeed rated at between 1 and 2 V.

4 We saw in Chapter 1 that the criterion of thermodynamic reversibility is the reversal of a process by an infinitesimal change in the external conditions. 5 This quantity was called the electromotive force, emf, of the electrochemical cell, but that name is deprecated by IUPAC because a potential difference is not a force.

5.7 STANDARD POTENTIALS

Our next step is to see how Ecell varies with composition by combining eqn 5.12 and eqn 4.6, showing how the reaction Gibbs energy varies with composition: DrG = DrG3 + RT ln Q In this expression, D rG3 is the standard reaction Gibbs energy and Q is the reaction quotient for the cell reaction. When we substitute this relation into eqn 5.12 written as Ecell = −D rG/nF, we obtain the Nernst equation: 3 Ecell = E cell −

RT lnQ nF

Nernst equation

(5.13)

The standard cell potential

(5.14)

3 is the standard cell potential: E cell 3 E cell =−

D rG 3 nF

The standard cell potential is often interpreted as the cell potential when all the reactants and products are in their standard states (unit activity for all solutes, pure gases, and solids, a pressure of 1 bar). However, because such an electro3 chemical cell is not in general attainable, it is better to regard E cell simply as the standard Gibbs energy of the reaction expressed as a potential. Note that if all the stoichiometric coefficients in the equation for a cell reaction are multiplied by a factor, then D rG3 is increased by the same factor, but so too is n, so the standard cell potential is unchanged. Likewise, Q is raised to a power equal to the factor (so if the factor is 2, Q is replaced by Q2), and because ln Q2 = 2 ln Q, and likewise for other factors, the second term on the right-hand side of the Nernst equation is also unchanged. That is, Ecell is independent of how we write the balanced equation for the cell reaction.

A brief illustration

At 25.00°C, RT (8.31447 J K−1) × (298.15 K) = = 2.5693 × 10−2 J C−1 F 9.6485 × 104 C mol−1 Because 1 J = 1 V C, 1 J C −1 = 1 V, and 10−3 V = 1 mV, we can write this result as RT = 25.693 mV F or approximately 25.7 mV. It follows from the Nernst equation that for a reaction in which n = 1, if Q is decreased by a factor of 10, then the potential of the electrochemical cell becomes more positive by (25.7 mV) × ln 10 = 59.2 mV. The reaction has a greater tendency to form products. If Q is increased by a factor of 10, then the cell potential falls by 59.2 mV and the reaction has a lower tendency to form products.

5.7 Standard potentials To discuss the thermodynamics of biological processes, we need to be able to predict the standard reaction Gibbs energies of biological electron transfer reactions and their variation with pH.

197

198

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

Each electrode in a galvanic cell makes a characteristic contribution to the overall cell potential. Although it is not possible to measure the contribution of a single electrode, one electrode can be assigned a value zero and the others assigned relative values on that basis. The specially selected electrode is the standard hydrogen electrode (SHE): Pt(s) | H2(g) | H+(aq)

E 3 = 0 at all temperatures

(a) Thermodynamic standard potentials

A brief comment

Standard potentials are also called standard electrode potentials and standard reduction potentials. If in an older source of data you come across a ‘standard oxidation potential,’ reverse its sign and use it as a standard reduction potential.

The (thermodynamic) standard potential, E3(Ox/Red), of a couple Ox/Red is measured by constructing an electrochemical cell in which the couple of interest forms the right-hand electrode and the standard hydrogen electrode is on the left. For example, the standard potential of the Ag+/Ag couple is the standard potential of the cell Pt(s) | H2(g) | H+(aq) | | Ag+(aq) | Ag(s) and is +0.80 V. Table 5.1 lists a selection of standard potentials; a longer list will be found in the Resource section. We saw in Section 4.2 that in biochemical work it is common to adopt the biological standard state (pH = 7, corresponding to neutral solution), rather than the thermodynamic standard state (pH = 0). To convert standard potentials to biological standard potentials, E ⊕, we must first consider the variation of potential with pH. The two potentials differ when hydrogen ions are involved in the half-reaction, as in the fumaric acid/succinic acid couple fum/suc with fum = HOOCCH=CHCOOH and suc = HOOCCH2CH2COOH, which plays a role in the citric acid cycle (Case study 4.3): HOOCCH=CHCOOH(aq) + 2 H+(aq) + 2 e− → HOOCCH2CH2COOH(aq) When hydrogen ions occur as reactants, as in this example, an increase in pH, corresponding to a decrease in hydrogen ion activity, favors the formation of reactants, so the fumaric acid has a lower thermodynamic tendency to become reduced. We expect, therefore, the potential of the fumaric/succinic acid couple to decrease as the pH is increased. (b) Variation of potential with pH

To establish the quantitative variation of reduction potential with pH for a reaction as a first step in determining the effect of changing from pH = 0 to pH = 7 we use the Nernst equation. Thus, for fixed fumaric acid and succinic acid concentrations, the potential of the fumaric/succinic redox couple is E′

1 4 2 4 3

E = E3 −

RT a RT asuc RT ln suc2 = E 3 − ln + ln aH 2F afumaH 2F afum T

+

+

We then use a result from Mathematical toolkit 5.1 to write ln aH = ln 10 × log aH = −ln 10 × pH +

+

and obtain E = E′ −

RT ln 10 × pH F

(5.15a)

5.7 STANDARD POTENTIALS

Table 5.1

Standard potentials at 25°C E 9/V

Reduction half-reaction Oxidizing agent

Reducing agent

Strongly oxidizing + 2 e−

→ 2 F−

+2.87

S2O

+2e

→ 2 SO42−

+2.05

Au+

+ e−

→ Au

+1.69

Pb4+

+ 2 e−

→ Pb2+

+1.67

Ce4+

+ e−

→ Ce3+

+1.61

MnO−4 + 8 H+

+ 5 e−

→ Mn2+ + 4 H2O

+1.51

→ 2 Cl

+1.36

F2 2− 8

Cl2

+2e

Cr2O72− + 14 H+

+ 6 e−

→ 2 Cr 3+ + 7 H2O

+1.33

O2 + 4 H+

+ 4 e−

→ 2 H2O

+1.23,

Br2

+ 2 e−

→ 2 Br−

+1.09

Ag+

+ e−

→ Ag

+0.80

Hg 22+

+ 2 e−

→ 2 Hg

+0.79

Fe

+ e−

→ Fe2+

+0.77

I2

+ e−

→ 2 I−

+0.54

O2 + 2 H2O

+ 4 e−

→ 4 OH−

+0.40,

Cu2+

+ 2 e−

→ Cu

+0.34

AgCl

+ e−

→ Ag + Cl−

+0.22

+0.81 at pH = 7

3+

0.81 at pH = 7

2H

+2e

→ H2

0, by definition

Fe3+

+ 3 e−

→ Fe

−0.04

O2 + H2O

+ 2 e−

→ HO−2 + OH−

−0.08

Pb2+

+ 2 e−

→ Pb

−0.13

Sn2+

+ 2 e−

→ Sn

−0.14

Fe

+2e

→ Fe

−0.44

Zn2+

+ 2 e−

→ Zn

−0.76

2 H2O

+ 2 e−

→ H2 + 2 OH−

−0.83,

Al3+

+ 3 e−

→ Al

−1.66

Mg2+

+ 2 e−

→ Mg

−2.36

Na

+e

→ Na

−2.71

Ca2+

+ 2 e−

→ Ca

−2.87

K+

+ e−

→K

−2.93

Li+

+ e−

→ Li

−3.05

+

2+

−0.42 at pH = 7

+

Strongly reducing For a more extensive table, see the Resource section.

Note that this result is valid only for a half-reaction in which ne = 2 and the stoichiometric coefficient of H+(aq) is 2 and appears as a reactant. In general: E = E′ −

nH RT ln 10 × pH ne F +

Variation of potential with pH

(5.15b)

199

200

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

A brief illustration

At 25°C and when nH = ne , +

E = E′ − (59.2 mV) × pH We see that an increase of 1 unit in pH decreases the potential by 59.2 mV, which is in agreement with the remark above, that the reduction of fumaric acid is discouraged by an increase in pH.

For a hydrogen electrode half-reaction, 12 H2(g) + e− → H+(aq), with E 3 = 0, the same calculation gives E=−

RT ln 10 × pH F

This expression is the basis of a method for measuring the pKa of an acid electrically. As we saw in Section 4.11, the pH of a solution containing equal amounts of the acid and its conjugate base is pH = pKa. Therefore, by measuring the potential of the cell SHE | | solution | HE, where SHE is a standard hydrogen electrode and HE is a hydrogen electrode dipping into the solution, we can determine the latter’s pH and therefore the pKa of the acid. It is in fact unwieldy to use an actual hydrogen electrode, and a far more convenient approach is developed later (In the laboratory 5.1). (c) The biological standard potential

We can now use eqn 5.15 to convert standard potentials to biological standard potentials. If the hydrogen ions appear as reactants in the reduction half-reaction, then the potential is decreased below its standard value (for the fumaric/succinic couple, by 7 × 59.2 mV = 414 mV, or about 0.4 V). If the hydrogen ions appear as products, then the biological standard potential is higher than the thermodynamic standard potential. The precise change depends on the number of electrons and protons participating in the half-reaction, as expressed by eqn 5.15b and illustrated in Example 5.4. Biological standard potentials are important in the discussion of the electron transfer reactions of oxidative phosphorylation (Section 5.10). Table 5.2 is a partial list of biological standard potentials for redox couples that participate in important biochemical electron transfer reactions.

Example 5.4

Converting a standard potential to a biological standard value

Calculate the biological standard potential of the NAD/NADH couple at 25°C (Example 5.2) from its thermodynamic value. The reduction half-reaction is NAD+(aq) + H+(aq) + 2 e− → NADH(aq)

E3 = −0.11 V

Strategy Write the Nernst equation for the potential, and express the reaction quotient in terms of the activities of the species. All species except H+ are in their standard states, so their activities are all equal to 1. The remaining task is to express the hydrogen ion activity in terms of the pH, exactly as was done in the text, and set pH = 7. Solution The Nernst equation for the half-reaction, with ne = 2, is

5.7 STANDARD POTENTIALS

Table 5.2

201

Biological standard potentials at 25°C E ⊕/V

Reduction half-reaction Oxidizing agent

Reducing agent

Strongly oxidizing O 2 + 4 H+

+ 4 e−

→ 2 H2O

+0.81 +0.36

Fe (Cyt f )

+e

→ Fe (Cyt f )

O2 + 2 H2O

+ 4 e−

→ 2 H2O2

+0.30

Fe3+ (Cyt c)

+ e−

→ Fe2+ (Cyt c)

+0.25

→ Fe2+ (Cyt b)

+0.08

3+

2+

+ e−

Fe3+ (Cyt b) Dehydroascorbic acid + 2 H

+2e

Coenzyme Q + 2 H+

→ Ascorbic acid

+0.08

+ 2 e−

→ Coenzyme QH2

+0.04

Oxaloacetate2− + 2 H+

+ 2 e−

→ Malate2−

−0.17

Pyruvate + 2 H

+2e

→ Lactate−

−0.18

FAD + 2 H+

+ 2 e−

→ FADH2

−0.22

Glutathione (ox) + 2 H

+2e

→ Glutathione (red)

−0.23

Lipoic acid (ox) + 2 H+

+ 2 e−

→ Lipoic acid (red)

−0.29

NAD+ + H+

+ 2 e−

→ NADH

−0.32

→ H2 + 2 OH

+

+

+

2 H2O

+2e

Ferredoxin (ox)

+ e−

→ Ferredoxin (red)

−0.43

O2

+ e−

→ O2−

−0.45

−0.42

Strongly reducing For a more extensive table, see the Resource section. 1

8 #

E = E3 −

RT a RT ln NADH = E 3 + ln aH 2F aH aNAD 2F +

A note on good practice +

+

# $ 1

We rearrange this expression to E = E3 +

RT RT ln 10 ln aH = E 3 − × pH = E 3 − (29.58 mV) × pH 2F 2F +

The biological standard potential (at pH = 7) is therefore E ⊕ = (−0.11 V) − (29.58 × 10−3 V) × 7 = −0.32 V Calculate the biological standard potential of the half-reaction O2(g) + 4 H (aq) + 4 e− → 2 H2O(l) at 25°C given its value +1.23 V under thermodynamic standard conditions. Self-test 5.6

+

Answer: +0.82 V

In the laboratory 5.1

Ion-selective electrodes

Special electrodes can be constructed to measure concentrations of ionic species, such as Na+, K+, Ca2+, and hydronium ions, which are important in biochemical reactions. The potential of a hydrogen electrode is directly

Whenever possible, avoid replacing activities by concentrations, especially when aiming to relate the electrode potential to pH, for the latter is defined in terms of the activity of hydrogen ions.

202

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

proportional to the pH of the solution. However, in practice, indirect methods are much more convenient to use than one based on the standard hydrogen electrode, and the hydrogen electrode is replaced by a glass electrode (Fig. 5.11). A glass electrode is an example of an ion-selective electrode, an electrode that generates a potential in response to the presence of a solution of specific ions. The glass of a glass electrode is based on lithium silicate doped with heavymetal oxides; it is filled with a phosphate buffer solution containing Cl− ions. Conveniently, the electrode has Ecell ≈ 0 when the external medium is at pH = 7. The electrode is calibrated using solutions of known pH (for example, one of the buffer solutions described in Section 4.11). What range should a voltmeter have (in volts) to display changes of pH from 1 to 14 at 25°C if it is arranged to give a reading of zero when pH = 7? Self-test 5.7

Answer: From −0.42 V to +0.35 V, a range of 0.77 V A glass electrode has a potential that varies with the hydrogen ion concentration in the medium in which it is immersed. It consists of a thin glass membrane containing an electrolyte and a silver chloride electrode, Ag(s) | AgCl(s) | Cl−(aq). The electrode is used in conjunction with a reference electrode, such as a calomel electrode, Hg(l) | Hg2Cl2(s) | Cl−(aq), that makes contact with the test solution through a salt bridge.

Fig. 5.11

Glass electrodes can be made responsive to Na+, K+, and NH4+ ions by using glasses doped with Al2O3 and B2O3. More sophisticated devices can extend the range of ions that can be detected in a test solution. For example, a porous hydrocarbon-attracting membrane can be attached to a small reservoir of a hydrophobic liquid, such as dioctylphenylphosphonate, that saturates it (Fig. 5.12). The liquid contains a compound, such as (RO)2PO2− with R a C8 to C18 chain, which binds to the ion. The bound ions traverse the membrane and give rise to a transmembrane potential, which is detected by an electrode in the assembly. Electrodes of this construction can be designed to be sensitive to a variety of ionic species, including Ca2+ ions.

Applications of standard potentials The measurement of the potential of an electrochemical cell is a convenient source of thermodynamic information on reactions. In practice the standard values (and the biological standard values) of these quantities are the ones normally determined. 5.8 The determination of thermodynamic functions Calorimetry is not always practicable, especially for biochemically important reactions, but in some cases their thermodynamic properties can be measured electrochemically.

The structure of an ion-selective electrode. Ions bound to a compound, the chelating agent, in the hydrophilic liquid are able to migrate through the lipophilic membrane.

Fig. 5.12

We have seen that the standard potential of an electrochemical cell is related to 3 the standard reaction Gibbs energy by eqn 5.14 (D rG 3 = −nFE cell ), therefore, by measuring the standard potential of a cell driven by the reaction of interest, we can obtain the standard reaction Gibbs energy. If we were interested in the biological standard state, then we would use the same expression but with the standard potential at pH = 7 (D rG ⊕ = −nFE ⊕cell). From the standard reaction Gibbs

5.8 THE DETERMINATION OF THERMODYNAMIC FUNCTIONS

energy, the equilibrium constant, standard entropy, and standard enthalpy can be calculated. (a) Calculation of the equilibrium constant

A special case of the Nernst equation has great importance in chemistry. Suppose the reaction has reached equilibrium; then Q = K, where K is the equilibrium constant of the cell reaction. However, because a chemical reaction at equilibrium cannot do work, it generates zero potential difference between the electrodes. Setting Q = K and Ecell = 0 in the Nernst equation gives ln K =

3 nFE cell RT

The equilibrium constant in terms of the standard cell potential

(5.16)

This very important equation—which is simply eqn 4.10 expressed electrochemically—lets us predict equilibrium constants from the standard potential of an electrochemical cell. Note that 3 • If E cell > 0, then K > 1 and at equilibrium the cell reaction lies in favor of products. 3 • If E cell < 0, then K < 1 and at equilibrium the cell reaction lies in favor of reactants.

A brief illustration

Because the standard potential of the Daniell cell is +1.10 V, the equilibrium constant for the cell reaction (reaction A) is ln K = =

2 × (9.6485 × 104 C mol−1) × (1.10 V) (8.3145 J K−1 mol−1) × (298.15 K) 2 × 9.6485 × 1.10 × 104 8.3145 × 298.15

(where we have used 1 C V = 1 J to cancel units) and therefore K = 1.5 × 1037. Hence, the displacement of copper by zinc goes virtually to completion in the sense that the ratio of concentrations of Zn2+ ions to Cu2+ ions at equilibrium is about 1037. This value is far too large to be measured by classical analytical techniques, but its electrochemical measurement is straightforward. Note that a standard cell potential of +1 V corresponds to a very large equilibrium constant (and −1 V would correspond to a very small one).

It is also possible to use the data in Tables 5.1 and 5.2 to calculate the standard potential of an electrochemical cell formed from any pair of electrodes and, from the standard cell potential, the equilibrium constant of the cell reaction. To calculate the standard potential of an electrochemical cell, we take the difference of the standard potentials of the appropriate electrodes: 3 E cell = E 3R − E3L

The standard cell potential from standard potentials

(5.17a)

where E 3R is the standard potential of the right-hand electrode and E3L is that of the left. The analogous expression for the biological standard state is

203

204

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

The biological standard cell potential from biological standard potentials

E ⊕cell = E ⊕R − E⊕L

(5.17b)

When dealing with biological systems, the focus is not necessarily on reactions occurring at electrodes but on electron transfer processes in the cytosol or membranes of biological cells. We can still estimate the standard reaction Gibbs energy (and hence the equilibrium constant) of biological electron transfer reactions by 3 using eqn 5.14 (written as DrG3 = −nFE cell ) if we express the chemical equation for the redox reaction as the difference of two reduction half-reactions with 3 or E ⊕cell from eqn 5.17 and use known standard potentials. We then find E cell eqn 5.14 for the calculation of the standard reaction Gibbs energy or eqn 5.16 for the calculation of the equilibrium constant. The approach is illustrated in the following example. Example 5.5

Calculating the equilibrium constant of a biological electron transfer reaction

The reduced and oxidized forms of riboflavin form a couple with E ⊕ = −0.21 V and the acetate/acetaldehyde couple has E⊕ = −0.60 V under the same conditions. What is the equilibrium constant for the reduction of riboflavin (Rib) by acetaldehyde (ethanal) in neutral solution at 25°C? The reaction is RibO(aq) + CH3CHO(aq) 7 Rib(aq) + CH3COOH(aq) where RibO is the oxidized form of riboflavin and Rib is the reduced form. Strategy The aim is to find the values of E ⊕cell and n corresponding to the reac-

tion, for then we can use a modified form of eqn 5.16 to calculate the value of K in neutral solution from E ⊕cell. To do so, we express the equation as the difference of two reduction half-reactions. The stoichiometric number of the electron in these matching half-reactions is the value of n we require. We then look up the biological standard potentials for the couples corresponding to the half-reactions and calculate their difference to find E ⊕cell. Solution The two reduction half-reactions are

right: RibO(aq) + 2 H+(aq) + 2 e− → Rib(aq) + H2O(l)

E ⊕ = −0.21 V

left: CH3COOH(aq) + 2 H+(aq) + 2 e− → CH3CHO(aq) + H2O(l) E⊕ = −0.60 V and their difference is the redox reaction required. Note that n = 2. The corresponding standard cell potential is E ⊕cell = (−0.21 V) − (−0.60 V) = +0.39 V It follows that ln K =

2FE ⊕cell 2 × (9.6485 × 104 C mol−1) × (0.39 V) = RT (8.3145 J K−1 mol−1) × (298.15 K) =

2 × 9.6485 × 0.39 × 104 8.3145 × 298.15

Therefore, because K = eln K, K = e(2×9.6485×0.39×10 )/(8.3145×298.15) = 1.5 × 1013 4

We conclude that riboflavin can be reduced by acetaldehyde in neutral solution. However, there may be mechanistic reasons—the energy required to

5.8 THE DETERMINATION OF THERMODYNAMIC FUNCTIONS

break covalent bonds, for instance—that make the reduction too slow to be feasible in practice. Note that, because hydrogen ions do not appear in the chemical equation, the equilibrium constant is independent of pH. Self-test 5.8 What is the equilibrium constant for the reduction of riboflavin with rubredoxin, a bacterial iron–sulfur protein, in the reaction

riboflavin(ox) + rubredoxin(red) 7 riboflavin(red) + rubredoxin(ox) given that the biological standard potential of the rubredoxin couple is −0.06 V? Answer: 8.5 × 10−6; the reactants are favored

(b) Calculation of standard potentials

The relation between the standard cell potential and the standard reaction Gibbs energy is a convenient route for the calculation of the standard potential of a couple from two other standard potentials. We make use of the fact that G is a state function and that the Gibbs energy of an overall reaction is the sum of the Gibbs energies of the reactions into which it can be divided. In general, we cannot combine the E values directly because they depend on the value of n, which may be different for the two couples.

Example 5.6

Calculating a standard potential from two other standard potentials

The superoxide ion (O−2 ) is an undesirable by-product of some enzymecatalyzed reactions. It is metabolized by the enzyme superoxide dismutase (SOD) in a disproportionation (or dismutation), a reaction that both oxidizes and reduces a species. The reaction catalyzed by SOD is 2 O−2 (aq) + 2 H+(aq) → H2O2(aq) + O2(g) where O−2 is oxidized to O2 and reduced to O22 (in H2O2). Hydrogen peroxide, H2O2, is also produced by other biochemical reactions. It is a toxic substance that is metabolized by catalases and peroxidases. The disproportionation catalyzed by catalase is −

2 H2O2(aq) → 2 H2O(l) + O2(g) Given the standard potentials E⊕(O2,O−2 ) = −0.45 V and E ⊕(O2,H2O2) = +0.30 V, calculate E ⊕(O −2 ,H2O2), the biological standard potential for the SOD-catalyzed reduction of O−2 to H2O2. Strategy We need to convert the two E ⊕ to D rG ⊕ by using eqn 5.14 modified

for the biological standard state, add them appropriately, and then convert the overall D rG⊕ so obtained to the required E ⊕ by using eqn 5.14 again. Because the Fs cancel at the end of the calculation, carry them through. Solution The electrode reactions are as follows:

(a) O2(g) + e− → O−2 (aq) E⊕ = −0.45 V

DrG ⊕(a) = −F × (−0.45 V) = (+0.45 V) × F

(b) O2(g) + 2 H+(aq) + 2 e− → H2O2(aq) E⊕ = +0.30 V

DrG ⊕(b) = −2F × (0.30 V) = (−0.60 V) × F

205

206

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

The required reaction is (c) O−2 (aq) + 2 H+(aq) + e− → H2O2(aq)

DrG ⊕(c) = −FE⊕

Because (c) = (b) − (a), it follows that A note on good practice

Whenever combining standard potentials to obtain the standard potential of a third couple, always work via the Gibbs energies because they are additive, whereas in general standard potentials are not.

DrG ⊕(c) = DrG⊕(b) − DrG ⊕(a) Therefore, from eqn 5.14, FE ⊕(c) = −{(−0.60 V)F − (+0.45 V)F} The Fs cancel, and we are left with E⊕(c) = +1.05 V. Given the standard potentials E3(Fe3+,Fe) = −0.04 V and E (Fe ,Fe) = −0.44 V, calculate E3(Fe3+,Fe2+). Self-test 5.9 3

2+

Answer: +0.76 V

(c) Calculation of the standard reaction entropy and enthalpy

Once DrG 3 has been measured, we can use thermodynamic relations to determine other properties. For instance, the entropy of the cell reaction can be obtained from the change in the potential with temperature: DrS 3 = nF

3 dE cell dT

The standard reaction entropy from the standard cell potential

(5.18)

A brief comment

Infinitesimally small quantities may be treated like any other quantity in algebraic manipulations. Thus, the expression dy = adx may be rewritten as dy/dx = a, dx/dy = 1/a, and so on.

Justification 5.4 The reaction entropy from the electrochemical cell potential

In Section 3.3 we used the fact that, at constant pressure, when the temperature changes by dT, the Gibbs energy changes by dG = −SdT. Because this equation applies to the reactants and the products, it follows that d(DrG 3) = −D rS 3 × dT 3 Substitution of D rG 3 = −nFE cell then gives 3 nF × dE cell = D rS 3 × dT

which rearranges into eqn 5.18.

We see from eqn 5.18 that the standard cell potential increases with temperature if the standard reaction entropy is positive and that the slope of a plot of potential against temperature is proportional to the reaction entropy (Fig. 5.13). An implication is that if the cell reaction produces a lot of gas (corresponding to a positive reaction entropy), then its potential will increase with temperature. The opposite is true for a reaction that consumes gas. Finally, we can combine the results obtained so far by using G = H − TS in the form H = G + TS to obtain the standard reaction enthalpy:

The variation of the standard potential of a cell with temperature depends on the standard entropy of the cell reaction.

Fig. 5.13

Dr H 3 = DrG 3 + TDr S 3

The standard reaction enthalpy from the standard reaction Gibbs energy and entropy

(5.19)

with D rG 3 determined from the cell potential and D r S3 from its temperature variation. Thus, we now have a noncalorimetric method of measuring a reaction enthalpy.

5.10 THE RESPIRATORY CHAIN

5.9 The electrochemical series Some organic co-factors and metal centers in proteins act as electron transfer agents in a number of biological processes; we need to be able to predict which species is reduced or oxidized in a redox reaction. 3 3 We have seen that a cell reaction has K > 1 if E cell > 0 and that E cell > 0 corresponds 3 to reduction at the right-hand electrode. We have also seen that E cell may be written as the difference of the standard potentials of the redox couples in the 3 right and left electrodes (eqn 5.17, E cell = E 3R − E3L ). A reaction corresponding to reduction at the right-hand electrode therefore has K > 1 if E3L < E3R, and we can conclude that

A couple with a low standard potential has a thermodynamic tendency to reduce a couple with a high standard potential. More briefly: low reduces high and, equivalently, high oxidizes low. The same arguments apply to the biological standard values of the potentials. A brief illustration

Consider the iron-containing protein ferredoxin, which participates in plant photosynthesis (Section 5.11), and cytochrome c, which participates in the last steps of respiration (Section 5.10). It follows from Table 5.2 that E⊕(ferredoxinox,ferredoxinred) = −0.43 V < E ⊕(cyt cox,Cyt cred) = +0.25 V and ferredoxin has a thermodynamic tendency to reduce cytochrome c at pH = 7. Hence, K > 1 for the reaction Cyt cox(aq) + ferredoxinred(aq) 7 Cyt cred(aq) + ferredoxinox(aq)

Self-test 5.10 Does NAD+ have a thermodynamic tendency to oxidize the

pyruvate ion at pH = 7? Answer: No

Electron transfer in bioenergetics Electron transfer between protein-bound cofactors or between proteins plays a role in a number of biological processes, such as the oxidative breakdown of foods, photosynthesis, nitrogen fixation, the reduction of atmospheric N2 to NH3 by certain microorganisms, and the mechanisms of action of oxidoreductases, which are enzymes that catalyze redox reactions. Here, we examine the redox reactions associated with photosynthesis and the aerobic oxidation of glucose. These processes are related by the reactions aerobic oxidation C6H12O6(s) + 6 O2(g) fffffg bccccc 6 CO2(g) + 6 H2O(l) photosynthesis

5.10 The respiratory chain The centrally important processes of biochemistry include the electrochemical reactions between proteins in the mitochondrion of the cell, for they are responsible for delivering the electrons extracted from glucose to water.

207

208

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

The half-reactions for the oxidation of glucose and the reduction of O2 are C6H12O6(s) + 6 H2O(l) → 6 CO2(g) + 24 H+(aq) + 24 e− 6 O2(g) + 24 H+(aq) + 24 e− → 12 H2O(l) We see that the exergonic oxidation of one C6H12O6 molecule requires the transfer of 24 electrons to six O2 molecules. However, the electrons do not flow directly from glucose to O2. In biological cells, glucose is oxidized to CO2 by NAD+ and FAD during glycolysis and the citric acid cycle (Case study 4.3): C6H12O6(s) + 10 NAD+ + 2 FAD + 4 ADP + 4 Pi + 2 H2O → 6 CO2 + 10 NADH + 2 FADH2 + 4 ATP + 6 H+ In the respiratory chain, electrons from the powerful reducing agents NADH and FADH2 pass through four membrane-bound protein complexes and two mobile electron carriers before reducing O2 to H2O. We shall see that the electron transfer reactions drive the synthesis of ATP at three of the membrane protein complexes. (a) Electron transfer reactions

The respiratory chain begins in complex I (NADH-Q oxidoreductase), where NADH is oxidized by coenzyme Q (Q, Atlas M5) in a two-electron reaction: H+ + NADH + Q fffg NAD+ + QH2 complex I

E⊕cell = +0.42 V, D rG⊕ = −81 kJ mol−1 where the reduction of Q to Q2− is accompanied by uptake of two H+ ions to yield QH2. Additional Q molecules are reduced by FADH2 in complex II (succinate-Q reductase): FADH2 + Q fffg FAD + QH2 complex II

E⊕cell = +0.32 V, DrG⊕ = −62 kJ mol−1 Reduced Q migrates to complex III (Q-cytochrome c oxidoreductase), which catalyzes the reduction of the protein cytochrome c (Cyt c). Cytochrome c contains the heme c group, the central iron ion of which can exist in oxidation states +3 and +2. The net reaction catalyzed by complex III is QH2 + 2 Fe3+(Cyt c) fffg Q + 2 Fe2+(Cyt c) + 2 H+ complex III

E⊕cell = +0.15 V, D rG⊕ = −29 kJ mol−1 Reduced cytochrome c carries electrons from complex III to complex IV (cytochrome c oxidase), where O2 is reduced to H2O: 2 Fe2+(Cyt c) + 2 H+ + 12 O2 fffg 2 Fe3+(Cyt c) + H2O complex IV

E⊕cell = +0.56 V, D rG⊕ = −108 kJ mol−1 (b) Oxidative phosphorylation

The reactions that occur in complexes I, II, III, and IV are exergonic and together could drive the synthesis of ATP: ADP + Pi + H+ → ATP

D rG⊕ = +31 kJ mol−1

We saw in Case study 4.2 that the phosphorylation of ADP to ATP can be coupled to the exergonic dephosphorylation of other molecules. Indeed, this is

5.11 PLANT PHOTOSYNTHESIS

the mechanism by which ATP is synthesized during glycolysis and the citric acid cycle (Case study 4.3). However, the process of oxidative phosphorylation taking place in mitochondria operates by a different mechanism. The structure of a mitochondrion is shown in Fig 5.14. The protein complexes associated with the electron transport chain span the inner membrane, and phosphorylation takes place in the intermembrane space. The Gibbs energy of the reactions in complexes I, III, and IV is first used to do the work of moving protons across the mitochondrial membrane. The complexes are oriented asymmetrically in the inner membrane so that the protons abstracted from one side of the membrane can be deposited on the other side. For example, the oxidation of NADH by Q in complex I is coupled to the transfer of four protons across the membrane. The coupling of electron transfer and proton pumping in complexes III and IV contribute further to a gradient of proton concentration across the membrane. Then the enzyme H+-ATPase uses the energy stored in the proton gradient to phosphorylate ADP to ATP. Experiments show that 11 molecules of ATP are made for every three molecules of NADH and one molecule of FADH2 that are oxidized by the respiratory chain. The ATP is then hydrolyzed on demand to perform useful biochemical work throughout the cell. Complex II does not contribute to oxidative phosphorylation because it does not have a proton pump. The chemiosmotic theory proposed by Peter Mitchell explains how H+-ATPases use the energy stored in a transmembrane proton gradient to synthesize ATP from ADP. It follows from eqn 5.8 that we can estimate the Gibbs energy available for phosphorylation by writing DGm = RT ln

[H+]in + FDf [H+]out

The Gibbs energy available for phosphorylation according to the chemiosmotic theory

(5.20)

where Df = fin − fout is the membrane potential difference and we have used z = +1. After using ln [H+] = (ln 10) log [H+] and substituting DpH = pHin − pHout = −log [H+]in + log [H+]out, it follows that DGm = FDf − (RT ln 10)DpH

(5.21)

A brief illustration

In the mitochondrion, DpH ≈ −1.4 and Df ≈ 0.14 V, so it follows from eqn 5.20 that DGm ≈ +21.5 kJ mol−1. Because 31 kJ mol−1 is needed for phosphorylation (Case study 4.2), we conclude that at least 2 mol H+ (and probably more) must flow through the membrane for the phosphorylation of 1 mol ADP.

5.11 Plant photosynthesis We need to appreciate that the mechanism of formation of glucose from carbon dioxide and water in photosynthetic organisms is distinctly different from the mechanism of glucose breakdown.

In plant photosynthesis, solar energy drives the endergonic reduction of CO2 to glucose, with concomitant oxidation of water to O2 (DrG⊕ = +2880 kJ mol−1). The process takes place in the chloroplast, a special organelle of the plant cell. Electrons flow from reductant to oxidant via a series of electrochemical reactions that are coupled to the synthesis of ATP. First, the leaf absorbs solar energy and transfers

209

Fig. 5.14 The general structure of a mitochondrion.

210

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

it to membrane protein complexes known as photosystem I and photosystem II.6 The absorption of energy from light decreases the reduction potential of special dimers of chlorophyll a molecules (Atlas R3) known as P700 (in photosystem I) and P680 (in photosystem II). In their high-energy or excited states, P680 and P700 initiate electron transfer reactions that culminate in the oxidation of water to O2 and the reduction of NADP+ (Atlas N5) to NADPH: 2 NADP+ + 2 H2O fg O2 + 2 NADPH + 2 H+ light

It is clear that energy from light is required to drive this reaction because, in the dark, E⊕cell = −1.135 V and DrG⊕ = +438.0 kJ mol−1. Working together, photosystem I and the enzyme ferredoxin:NADP+ oxidoreductase catalyze the light-induced reduction of NADP+ to NADPH. The electrons required for this process come initially from P700 in its excited state. The resulting P700 is then reduced by the mobile carrier plastocyanin (Pc), a protein in which the bound copper ion can exist in oxidation states +2 and +1. The net reaction is NADP+ + 2 Cu+(Pc) + H+ ffffffg NADPH + 2 Cu2+(Pc) light, photosystem I

Oxidized plastocyanin accepts electrons from reduced plastoquinone (PQ). The process is catalyzed by the cytochrome b6 f complex, a membrane protein complex that resembles complex III of mitochondria: PQH2 + 2 Cu2+(Pc) ffffg PQ + 2 H+ + 2 Cu+(Pc) Cyt b6 f complex

E⊕cell = +0.370 V, D rG⊕ = −71.4 kJ mol−1 Plastoquinone is reduced by water in a process catalyzed by light and photosystem II. The electrons required for the reduction of plastoquinone come initially from P680 in its excited state. The resulting P680 is then reduced ultimately by water. The net reaction is H2O + PQ ffffffg 12 O2 + PQH2 light, photosystem II

Electron transfer reactions are coupled to the movement of protons across membranes. Photophosphorylation uses the energy stored in the transmembrane proton gradient to phosphorylate ADP to ATP in H+-ATPases (Fig. 5.15). We see that plant photosynthesis uses an abundant source of electrons (water) and of energy (the Sun) to drive the endergonic reduction of NADP+, with concomitant synthesis of ATP. Experiments show that for each molecule of NADPH formed in the chloroplast of green plants, one molecule of ATP is synthesized. The ATP and NADPH molecules formed by the light-induced electron transfer reactions of plant photosynthesis participate directly in the reduction of CO2 to glucose in the chloroplast: 6 CO2 + 12 NADPH + 12 ATP + 12 H+ → C6H12O6 + 12 NADP+ + 12 ADP + 12 Pi + 6 H2O In summary, electrochemical reactions mediated by membrane protein complexes harness energy in the form of ATP. Plant photosynthesis uses solar energy to transfer electrons from a poor reductant (water) to carbon dioxide. In the process, high-energy molecules (carbohydrates, such as glucose) are synthesized in the cell. Animals feed on the carbohydrates derived from photosynthesis. 6

See Chapter 13 for details of the energy transfer process.

CHECKLIST OF KEY CONCEPTS

211

In plant photosynthesis, light-induced electron transfer processes lead to the oxidation of water to O2 and the reduction of NADP+ to NADPH, with concomitant production of ATP. The energy stored in ATP and NADPH is used to reduce CO2 to carbohydrate in a separate set of reactions. The scheme summarizes the general patterns of electron flow and does not show all the intermediate electron carriers in photosystems I and II, the cytochrome b6 f complex, and ferredoxin:NADP+ oxidoreductase. Fig. 5.15

During aerobic metabolism, the O2 released by photosynthesis as a waste product is used to oxidize carbohydrates to CO2, driving biological processes such as biosynthesis, muscle contraction, cell division, and nerve conduction. Hence, the sustenance of life on Earth depends on a tightly regulated carbon–oxygen cycle that is driven by solar energy.

Checklist of key concepts 1. Deviations from ideal behavior in ionic solutions are ascribed to the interaction of an ion with its ionic atmosphere. 2. According to the Debye–Hückel limiting law, the mean activity coefficient of ions in a solution is related to the ionic strength of the solution. 3. The Gibbs energy of transfer of an ion across a cell membrane is determined by an activity gradient and a membrane potential difference that arises from differences in Coulomb repulsions on each side of the bilayer.

8. The cell potential is the potential difference that the cell produces when operating reversibly. 9. The standard potential of a couple is the standard potential of a cell in which it forms the right-hand electrode and a hydrogen electrode is on the left. 10. Biological standard potentials are measured in neutral solution (pH = 7). 11. A couple with a low standard potential has a thermodynamic tendency (in the sense K > 1) to reduce a couple with a high standard potential.

4. A galvanic cell is an electrochemical cell in which a spontaneous chemical reaction produces a potential difference.

12. In the respiratory chain, electrons from NADH and FADH2 pass through four membrane-bound protein complexes and two mobile electron carriers before reducing O2 to H2O.

5. An electrolytic cell is an electrochemical cell in which an external source of current is used to drive a non-spontaneous chemical reaction.

13. The chemiosmotic theory explains how H+-ATPases use the energy stored in a transmembrane proton gradient to synthesize ATP from ADP.

6. A redox reaction is expressed as the difference of two reduction half-reactions.

14. Plant photosynthesis uses solar energy to transfer electrons from a poor reductant (water) to carbon dioxide.

7. In an electrochemical cell, a cathode is the site of reduction; an anode is the site of oxidation.

212

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

Checklist of key equations Property

Equation

Comment

Mean activity coefficient

g± = (g+g−)1/2

MX salt

g± = (g g )

p q 1/s + −

s=p+q

Debye–Hückel limiting law

log g± = −A | z+ z− | I 1/2

Ionic strength

I = 12 ∑ z 2i bi /b3 i ±

MpXq salt Valid as I → 0 Definition

Extended Debye–Hückel law

log g = −{A | z+ z− | I /(1 + BI )} + CI

Gibbs energy of transfer of an ion across a biological membrane

DGm = RT ln([A]in /[A]out) + zF Df + D rG ATP

Maximum non-expansion work

w′max = DG

Relation between the cell potential and the reaction Gibbs energy

−nFEcell = D rG

Nernst equation

3 − (RT/nF) ln Q Ecell = E cell

Standard cell potential

3 E cell = −DrG 3/nF

Equilibrium constant in terms of the standard cell potential

3 ln K = nFE cell /RT

Standard reaction entropy from the standard cell potential

3 /dT) Dr S 3 = nF(dEcell

Standard reaction enthalpy

Dr H3 = DrG 3 + TDrS 3

Gibbs energy available for phosphorylation

DGm = RT ln ([H+]in/[H+]out) + FDf

1/2

2

Constant pressure and temperature

Chemiosmotic theory

Discussion questions 5.1 Describe the general features of the Debye–Hückel theory of electrolyte solutions. 5.2 Describe the mechanism of proton conduction in water. 5.3 Distinguish between galvanic, electrolytic, and fuel cells. 5.4 Explain why some reactions that are not redox reactions may be

5.6 Review the concepts in Chapters 1 through 5 and prepare a summary of the experimental and calculational methods that can be used to measure or estimate the Gibbs energies of phase transitions and chemical reactions. 5.7 Review the concepts in Chapters 1 through 5 and discuss how ATP is formed during the metabolism of glucose.

used to generate an electric current. 5.5 Describe a method for the determination of the standard potential of an electrochemical cell.

Exercises 5.8 Relate the ionic strengths of (a) KCl, (b) FeCl3, and (c) CuSO4 solutions to their molalities, b.

5.12 Estimate the mean ionic activity coefficient and activity of a solution that is 0.010 mol kg−1 CaCl2(aq) and 0.030 mol kg−1 NaF(aq).

5.9 Calculate the ionic strength of a solution that is 0.10 mol kg−l in

5.13 The mean activity coefficients of HBr in three dilute aqueous solutions at 25°C are 0.930 (at 5.0 mmol kg−1), 0.907 (at 10.0 mmol kg−1), and 0.879 (at 20.0 mmol kg−1). Estimate the value of B in the extended Debye–Hückel law, with C = 0.

KCl(aq) and 0.20 mol kg−1 in CuSO4(aq). 5.10 Calculate the masses of (a) Ca(NO3)2 and, separately, (b) NaCl to add to a 0.150 mol kg−1 solution of KNO3(aq) containing 500 g of solvent to raise its ionic strength to 0.250. 5.11 Express the mean activity coefficient of the ions in a solution of

CaCl2 in terms of the activity coefficients of the individual ions.

5.14 The addition of a small amount of a salt, such as (NH4)2SO4, to

a solution containing a charged protein increases the solubility of the protein in water. This observation is called the salting-in effect. However, the addition of large amounts of salt can decrease the

EXERCISES

solubility of the protein to such an extent that the protein precipitates from solution. This observation is called the salting-out effect and is used widely by biochemists to isolate and purify proteins. Consider the equilibrium PXn(s) 7 Pn+ (aq) + nX−(aq), where Pn+ is a polycationic protein of charge +n and X− is its counter-ion. Use Le Chatelier’s principle and the physical principles behind the Debye– Hückel theory to provide a molecular interpretation for the salting-in and salting-out effects. 5.15 The overall reaction for the active transport of Na+ and K+ ions +

+

by the Na /K pump is 3 Na+(aq,inside) + 2 K+(aq,outside) + ATP → ADP + Pi + 3 Na+(aq,outside) + 2 K+(aq,inside)

213

process, which occurs in an acidic environment. (b) Estimate the values of E ⊕cell, D rG⊕, and K for the reaction at 25°C. 5.24 A fuel cell develops an electric potential from the chemical

reaction between reagents supplied from an outside source. What is the cell potential of a cell fuelled by (a) hydrogen and oxygen, each at 1 bar and 298 K, and (b) the combustion of butane at 1.0 bar and 298 K? 5.25 Consider a hydrogen electrode in HBr(aq) at 25°C operating

at 1.45 bar. Estimate the change in the electrode potential when the solution is changed from 5.0 mmol dm−3 to 25.0 mmol dm−3. 5.26 A hydrogen electrode can, in principle, be used to monitor

At 310 K, DrG⊕ for the hydrolysis of ATP is −31.3 kJ mol−1. Given that the [ATP]/[ADP] ratio is of the order of 100, is the hydrolysis of 1 mol ATP sufficient to provide the energy for the transport of Na+ and K+ according to the equation above? Take [Pi] = 1.0 mol dm−3.

changes in the molar concentrations of weak acids in biologically active solutions. Consider a hydrogen electrode in a solution of lactic acid as part of an overall galvanic cell at 25°C and 1 bar. Estimate the change in the electrode potential when the concentration of lactic acid in the solution is changed from 5.0 mmol dm−3 to 25.0 mmol dm−3.

5.16 Vision begins with the absorption of light by special cells in

5.27 Write the cell reactions and electrode half-reactions for the

the retina. Ultimately, the energy is used to close ligand-gated ion channels, causing sizable changes in the transmembrane potential. The pulse of electric potential travels through the optical nerve and into the optical cortex, where it is interpreted as a signal and incorporated into the web of events we call visual perception (see Chapter 13). Taking the resting potential as −30 mV, the temperature as 310 K, permeabilities of the K+ and Cl− ions as PK+ = 1.0 and PCl− = 0.45, respectively, and the concentrations as [K+]in = 100 mmol dm−3, [Na+]in = 10 mmol dm−3, [Cl−]in = 10 mmol dm−3, [K+]out = 5 mmol dm−3, [Na+]out = 140 mmol dm−3, and [Cl−]out = 100 mmol dm−3, calculate the relative permeability (Case study 5.1) of the Na+ ion.

following cells:

5.17 Is the conversion of pyruvate ion to lactate ion

CH3COCO2−(aq) + NADH(aq) + H+(aq) → CH3CH2(OH)CO2−(aq) + NAD+(aq)

(a) Pt(s) | H2(g, pL) | HCl(aq) | H2(g, pR) | Pt(s) (b) Pt(s) | Cl2(g) | HCl(aq) | | HBr(aq) | Br2(l) | Pt(s) (c) Pt(s) | NAD+(aq), H+(aq),NADH(aq) | | oxaloacetate2−(aq), H+(aq),malate2−(aq) | Pt(s) (d) Fe(s) | Fe2+(aq) | | Mn2+(aq), H+(aq) | MnO2(s) | Pt(s) 5.28 Write the Nernst equations for the cells in the preceding exercise. 5.29 Devise cells to study the following biochemically important

reactions. In each case state the value for n to use in the Nernst equation. (a) CH3CH2OH(aq) + NAD+(aq) → CH3CHO(aq) + NADH(aq) + H+(aq)

a redox reaction?

(b) ATP4−(aq) + Mg2+(aq) → MgATP2−(aq)

5.18 Express the reaction in Exercise 5.17 as the difference of two

(c) 2 cyt-c(red, aq) + CH3COCO2−(aq) + 2 H+(aq) → 2 cyt-c(ox, aq) + CH3CH(OH)CO2−(aq)

half-reactions. 5.19 Express the reaction in which ethanol is converted to +

acetaldehyde (propanal) by NAD in the presence of alcohol dehydrogenase as the difference of two half-reactions and write the corresponding reaction quotients for each half-reaction and the overall reaction. 5.20 Express the oxidation of cysteine, HSCH2CH(NH2)COOH, to cystine, HOOCCH(NH2)CH2SSCH2CH(NH2)COOH, as the difference of two half-reactions, one of which is O2(g) + 4 H+(aq) + 4 e− → 2 H2O(l). 5.21 One of the steps in photosynthesis is the reduction of NADP+ by

ferredoxin (fd) in the presence of ferredoxin:NADP oxidoreductase: 2 fdred(aq) + NADP+(aq) + 2 H+(aq) → 2 fdox(aq) + NADPH(aq). Express this reaction as the difference of two half-reactions. How many electrons are transferred in the reaction event? + 5.22 From the biological standard half-cell potentials E ⊕ cell(O2,H ,H2O)

= +0.82 V and E ⊕cell(NAD,H+,NADH) = −0.32 V, calculate the standard potential arising from the reaction in which NADH is oxidized to NAD+ and the corresponding biological standard reaction Gibbs energy. 5.23 Cytochrome c oxidase receives electrons from reduced

cytochrome c (cyt-cred ) and transmits them to molecular oxygen, with the formation of water. (a) Write a chemical equation for this

5.30 Use the standard potentials of the electrodes to calculate the

standard potentials of the cells devised in Exercise 5.29. 5.31 The permanganate ion is a common oxidizing agent. What is the standard potential of the MnO4−,H+/Mn2+ couple at (a) pH = 6.00 and (b) general pH? 5.32 State what you would expect to happen to the cell potential

when the following changes are made to the corresponding cells in Exercise 5.27. Confirm your prediction by using the Nernst equation in each case. (a) The pressure of hydrogen in the left-hand compartment is increased. (b) The concentration of HCl is increased. (c) Acid is added to both compartments. (d) Acid is added to the right-hand compartment. 5.33 State what you would expect to happen to the cell potential when

the following changes are made to the corresponding cells devised in Exercise 5.29. Confirm your prediction by using the Nernst equation in each case. (a) The pH of the solution is raised. (b) A solution of Epsom salts (magnesium sulfate) is added. (c) Sodium lactate is added to the solution. 5.34 (a) Calculate the standard potential of the cell

Hg(l) | HgCl2(aq) | | TlNO3(aq) | Tl(s) at 25°C. (b) Calculate the cell potential when the molar concentration of the Hg2+ ion is 0.150 mol dm−3 and that of the Tl+ ion is 0.93 mol dm−3.

214

5 THERMODYNAMICS OF ION AND ELECTRON TRANSPORT

5.35 Calculate the biological standard Gibbs energies of reactions of

the following reactions and half-reactions: (a) 2 NADH(aq) + O2(g) + 2 H+(aq) → 2 NAD+(aq) + 2 H2O(l) E ⊕cell = +1.14 V (b) Malate2−(aq) + NAD+(aq) → oxaloacetate2−(aq) + NADH(aq) + H+(aq) (c) O2(g) + 4 H+(aq) + 4 e− → 2 H2O(l)

5.44 One ecologically important equilibrium is that between

E ⊕cell = −0.154 V E ⊕cell = +0.81 V

5.36 The silver–silver chloride electrode, Ag(s) | AgCl(s) | Cl−(aq),

consists of metallic silver coated with a layer of silver chloride (which does not dissolve in water) in contact with a solution containing chloride ions. (a) Write the half-reaction for the silver–silver chloride half-electrode. (b) Estimate the potential of the cell Ag(s) | AgCl(s) | KCl(aq, 0.025 mol kg−1) | | AgNO3(aq, 0.010 mol kg−1) | Ag(s) at 25°C. 5.37 (a) Calculate the standard potential of the cell Pt(s) | cysteine(aq),

cystine(aq) | | H+(aq) | O2(g) | Pt(s) and the standard Gibbs energy and enthalpy of the cell reaction at 25°C. (b) Estimate the value of D rG 3 at 35°C. Use E 3 = −0.34 V for the cystine/cysteine couple. 5.38 The biological standard potential of the couple pyruvic acid/

lactic acid is −0.19 V. What is the thermodynamic standard potential of the couple? Pyruvic acid is CH3COCOOH and lactic acid is CH3CH(OH)COOH. 5.39 Calculate the biological standard values of the potentials (the

two potentials and the cell potential) for the system in Exercise 5.37 at 310 K. 5.40 (a) Does FADH2 have a thermodynamic tendency to reduce

coenzyme Q at pH = 7? (b) Does oxidized cytochrome b have a thermodynamic tendency to oxidize reduced cytochrome f at pH = 7?

5.41 Radicals, very reactive species containing one or more unpaired

electrons, are among the by-products of metabolism. Evidence is accumulating that radicals are involved in the mechanism of aging and in the development of a number of conditions, ranging from cardiovascular disease to cancer. Antioxidants are substances that reduce radicals readily. Which of the following known antioxidants is the most efficient (from a thermodynamic point of view): ascorbic acid (vitamin C), reduced glutathione, reduced lipoic acid, or reduced coenzyme Q? 5.42 The biological standard potential of the redox couple pyruvic

acid/lactic acid is −0.19 V and that of the fumaric acid/succinic acid couple is +0.03 V at 298 K. What is the equilibrium constant at pH = 7 for the reaction pyruvic acid + succinic acid 7 lactic acid + fumaric acid 5.43 Tabulated thermodynamic data can be used to predict the

standard potential of a cell even if it cannot be measured directly. The presence of glyoxylate ion produced by the action of the enzyme glycolate oxidase on glycolate ion can be monitored by the following redox reaction: 2 cyt-c(ox,aq) + glycolate−(aq) 7 2 cyt-c(red,aq) + glyoxylate−(aq) + 2 H+(aq)

The equilibrium constant for the reaction above is 2.14 × 1011 at pH = 7.0 and 298 K. (a) Calculate the biological standard potential of the corresponding galvanic cell and (b) the biological standard potential of the glyoxylate−/glycolate− couple. carbonate and hydrogencarbonate (bicarbonate) ions in natural water. (a) The standard Gibbs energies of formation of CO32−(aq) and HCO3−(aq) are −527.81 kJ mol−1 and −586.77 kJ mol−1, respectively. What is the standard potential of the HCO3−/CO2− 3 ,H2 couple? (b) Calculate the standard potential of a cell in which the cell reaction is Na2CO3(aq) + H2O(l) → NaHCO3(aq) + NaOH(aq). (c) Write the Nernst equation for the cell, (d) Predict and calculate the change in potential when the pH is changed to 7.0. (e) Calculate the value of pKa for HCO3−(aq). 5.45 The dichromate ion in acidic solution is a common oxidizing agent for organic compounds. Derive an expression for the potential of an electrode for which the half-reaction is the reduction of Cr2O72− ions to Cr3+ ions in acidic solution. 5.46 The potential of the cell Pt(s) | H2(g) | HCl(aq) | AgCl(s) | Ag(s) is +0.312 V at 25°C. What is the pH of the electrolyte solution? 5.47 The standard potential of the AgCl/Ag,Cl− couple fits the

expression E3/V = 0.23659 − 4.8564 × 10−4(q/°C) − 3.4205 × 10−6 (q/°C)2 + 5.869 × 10−9(q/°C)3 Calculate the standard Gibbs energy and enthalpy of formation of Cl−(aq) and its entropy (relative to H+) at 298 K. 5.48 If the mitochondrial electric potential between the matrix

and the intermembrane space were 70 mV, as is common for other membranes, how much ATP could be synthesized from the transport of 4 mol H+, assuming the pH difference remains the same? 5.49 Under certain stress conditions, such as viral infection

or hypoxia, plants have been shown to have an intercellular pH increase of about 0.1 pH. Suppose this pH change also occurs in the mitochondrial intermembrane space. How much ATP can now be synthesized for the transport of 2 mol H+, assuming no other changes occur? 5.50 In anaerobic bacteria, the source of carbon may be a molecule

other than glucose and the final electron acceptor some molecule other than O2. Could a bacterium evolve to use the ethanol/nitrate pair instead of the glucose/O2 pair as a source of metabolic energy? 5.51 The following reaction occurs in the cytochrome b6 f complex,

a component of the electron transport chain of plant photosynthesis: cyt-b(red) + cyt-f(ox) 7 cyt-b(ox) + cyt-f(red) (a) Calculate the biological standard Gibbs energy of this reaction. (b) The Gibbs energy for hydrolysis of ATP under conditions found in the chloroplast is −50 kJ mol−1 and the synthesis of ATP by ATPase requires the transfer of four protons across the membrane. How many electrons must pass through the cytochrome b6 f complex to lead to the generation of a transmembrane proton gradient that is large enough to drive ATP synthesis in the chloroplast?

PROJECT

215

Project 5.52 The standard potentials of proteins are not commonly measured by the methods described in this chapter because proteins often lose their native structure and their function when they react on the surfaces of electrodes. In an alternative method, the oxidized protein is allowed to react with an appropriate electron donor in solution. The standard potential of the protein is then determined from the Nernst equation, the equilibrium concentrations of all species in solution, and the known standard potential of the electron donor. We shall illustrate this method with the protein cytochrome c.

(a) The one-electron reaction between cytochrome c, cyt-c, and 2,6-dichloroindophenol, D, can be written as cyt-cox + Dred 7 cyt-cred + Dox 3 Consider Ecyt and E D3 to be the standard potentials of cytochrome c and D, respectively. Show that, at equilibrium (eq), a plot of ln([Dox]eq/ [Dred]eq) against ln([cyt-cox]eq/[cyt-cred]eq) is linear with a slope of 1 and

3 y-intercept F(Ecyt − E D3)/RT, where equilibrium activities are replaced by the numerical values of equilibrium molar concentrations.

(b) The following data were obtained for the reaction between oxidized cytochrome c and reduced D at pH 6.5 buffer and 298 K. The ratios [Dox]eq/[Dred]eq and [cyt-cox]eq/[cyt-cred]eq were adjusted by adding known volumes of a solution of sodium ascorbate, a reducing agent, to a solution containing oxidized cytochrome c and reduced D. From the data and the standard potential of D of +0.237 V, determine the standard potential of cytochrome c at pH = 6.5 and 298 K. [Dox]eq / [Dred ]eq

0.002 79 0.008 43 0.0257 0.0497 0.0748 0.238 0.534

[cyt-cox ]eq / 0.0106 [cyt-cred]eq

0.0230

0.0894 0.197

0.335

0.809 1.39

This page intentionally left blank

PART 2 The kinetics of

life processes

The branch of physical chemistry called chemical kinetics is concerned with the rates of chemical reactions. Chemical kinetics deals with how rapidly reactants are consumed and products formed, how reaction rates respond to changes in the conditions or the presence of a catalyst, and the identification of the steps by which a reaction takes place. One reason for studying the rates of reactions is the practical importance of being able to predict how quickly a reaction mixture approaches equilibrium. The rate might depend on variables under our control, such as the temperature and the presence of a catalyst, and we might be able to optimize it by the appropriate choice of conditions. Another reason is that the study of reaction rates leads to an understanding of the mechanism of a reaction, its analysis into a sequence of elementary steps. For example, by analyzing the rates of biochemical reactions, we may discover how they take place in an organism and contribute to the activity of a cell. Enzyme kinetics, the study of the effect of enzymes on the rates of reactions, is also an important window on how these macromolecules work and is treated in Chapter 8 using the concepts developed in Chapters 6 and 7.

This page intentionally left blank

6

The rates of reactions When dealing with physical and chemical changes, we need to cope with a wide variety of different rates. Even a process that appears to be slow may be the outcome of many faster steps. This is particularly true in the chemical reactions that underlie life. Some of the earlier steps in photosynthesis may take place in about 1–100 ps. The binding of a neurotransmitter can have an effect after about 1 s. Once a gene has been activated, a protein may emerge in about 100 s, but even that timescale incorporates many others, including the wriggling of a newly formed polypeptide chain into its working conformation, each step of which may take about 1 ps. On a grander view, some of the equations of chemical kinetics are applicable to the behavior of whole populations of organisms: such societies change on timescales of 107–109 s.

Reaction rates In the laboratory 6.1

Experimental techniques

221

Rate laws and rate constants

223

6.3

Reaction order

224

6.4

The determination of the rate law

225

6.2

6.5

In the laboratory 6.1

Experimental techniques

Some of the more common methods for investigations of reaction kinetics are listed in Table 6.1. (a) The determination of concentration

Spectrophotometry, the measurement of the absorption of light by a material, is used widely to monitor concentration. The technique is based on Beer’s law (see Chapter 12), which states that the incident and transmitted intensities

219

The definition of reaction rate

6.1

Reaction rates The raw data from experiments to measure reaction rates are the concentrations or (for gases) partial pressures of reactants and products at a series of times after the reaction is initiated. Ideally, information on any intermediates should also be obtained, but often intermediates cannot be studied because their existence is too fleeting or their concentration too low. More information about the reaction can be extracted if data are obtained at a series of different temperatures. The first step in the investigation of the rate and mechanism of a reaction is the determination of the overall stoichiometry of the reaction and the identification of any side reactions. The next step is to determine how the concentrations of the reactants and products change with time after the reaction has been initiated. Because the rates of chemical reactions are sensitive to temperature, the temperature of the reaction mixture must be held constant throughout the course of the reaction, for otherwise the observed rate would be a meaningless average of the rates for different temperatures. The next few sections look at these observations in more detail.

219

Integrated rate laws 228

Case study 6.1

Pharmaco*kinetics

234

The temperature dependence of reaction rates

235

6.6

6.7

The Arrhenius equation

235

Preliminary interpretation of the Arrhenius parameters

237

Checklist of key concepts 239 Checklist of key equations 239 Discussion questions

239

Exercises

240

Project

242

220

Table 6.1

6 THE RATES OF REACTIONS

Kinetic techniques

Technique

Range of timescales/s

Flash photolysis

10−15

Fluorescence decaya

10−10–10 −6

Ultrasonic absorption

10−10–10 −4

EPR

10 −9–10 −4

b

Electric field jumpc Temperature jump

c

10−7–1 −6

10 –1

Phosphorescence decay a

10 −6–10

NMRb

10 −5–1

Pressure jumpc

>10 −5

Stopped flow

>10 −3

a

Fluorescence and phosphorescence are modes of emission of radiation from a material; see Chapter 12. b EPR is electron paramagnetic resonance (or electron spin resonance, ESR); NMR is nuclear magnetic resonance; see Chapter 13. c These techniques are discussed in In the laboratory 7.1.

The arrangement used in the flow technique for studying reaction rates. The reactants are squirted into the mixing chamber at a steady rate from the syringes or by using peristaltic pumps (pumps that squeeze the fluid through flexible tubes, like in our intestines). The location of the spectrometer (acting as a detector) corresponds to different times after initiation.

Fig. 6.1

I and I0, respectively, of light passing through a sample of length L are related to the molar concentration [J] of the absorbing species J by I = I010−ε[J]L

Beer’s law

(6.1a)

The molar absorption coefficient ε (epsilon) depends on the wavelength of the radiation. Once the value of ε has been measured (in a separate experiment) for an absorbing species taking part in a reaction, either as a reactant or a product, its concentration may be monitored by using eqn 6.1a in the form [J] =

1 I log 0 I εL

(6.1b)

Note that log denotes a logarithm to the base 10. Reactions that change the concentration of hydrogen ions can be studied by monitoring the pH of the solution with a glass electrode. Other methods of monitoring the composition include the detection of light emission, microscopy, mass spectrometry, gas chromatography, and magnetic resonance (both EPR and NMR; Chapter 13). Polarimetry and circular dichroism (Chapter 12), which monitor the optical activity of a reaction mixture, are occasionally applicable. (b) Monitoring the time dependence

In a real-time analysis, the composition of a system is analyzed while the reaction is in progress by direct spectrophotometric observation of the reaction mixture. In the flow method, the reactants are mixed as they flow together in a chamber (Fig. 6.1). The reaction continues as the thoroughly mixed solutions flow through a capillary outlet tube at about 10 m s−1, and different points along the tube correspond to different times after the start of the reaction. Spectrophotometric determination of the composition at different positions along the tube is equivalent to the determination of the composition of the reaction mixture at different times after mixing. This technique was originally developed in connection with the study of the rate at which oxygen combines with hemoglobin (Case study 4.1). Its disadvantage is that a large volume of reactant solution is necessary because the mixture must flow continuously through the apparatus. This disadvantage is particularly important for reactions that take place very rapidly, because the flow must be rapid if it is to spread the reaction over an appreciable length of tube. The stopped-flow technique avoids this disadvantage (Fig. 6.2). The two solutions are mixed very rapidly (in less than 1 ms) by injecting them into a mixing chamber designed to ensure that the flow is turbulent and that complete mixing occurs very quickly. Behind the reaction chamber there is an observation cell fitted with a plunger that moves back as the liquids flood in, but that comes up against a stop after a certain volume has been admitted. The filling of that chamber corresponds to the sudden creation of an initial sample of the reaction mixture. The reaction then continues in the thoroughly mixed solution and is monitored spectrophotometrically. Because only a small, single charge of the reaction chamber is prepared, the technique is much more economical than the flow method. Modern techniques of monitoring composition spectrophotometrically can span repetitively a wavelength range of 300 nm at 1 ms intervals. The suitability of the stopped-flow technique to the study of small samples means that it is appropriate for biochemical reactions, and it has been widely used to study the kinetics of protein folding and unfolding. In a typical

6.1 THE DEFINITION OF REACTION RATE

221

experiment, a sample of the protein with a high concentration of a chemical denaturant, such as urea or guanidinium hydrochloride, is mixed with a solution containing a much lower concentration of the same denaturant. On entering the mixing chamber, the denaturant is diluted and the protein re-folds. Unfolding is observed by mixing a sample of folded protein with a solution containing a high concentration of denaturant. These experiments probe conformational changes that occur on a millisecond timescale, such as the formation of contacts between helical segments in a large protein. Very fast reactions can be studied by flash photolysis, in which the sample is exposed to a brief flash of light that initiates the reaction and then the contents of the reaction chamber are monitored spectrophotometrically. Biological processes that depend on the absorption of light, such as photosynthesis and vision, can be studied in this way. Lasers can be used to generate nanosecond flashes routinely, picosecond flashes quite readily, and flashes as brief as a few femtoseconds in special arrangements. Spectra are recorded at a series of times following the flash, using instrumentation described in Chapter 12. In a relaxation technique the reaction mixture is initially at equilibrium but is then disturbed by a rapid change in conditions, such as a sudden increase in temperature. The equilibrium composition before the application of the perturbation becomes the initial state for the return of the system to its equilibrium composition at the new temperature, and the return to equilibrium— the ‘relaxation’ of the system—is monitored spectroscopically. Relaxation techniques are described in more detail in In the laboratory 7.1. In contrast to real-time analysis, quenching methods are based on stopping, or quenching, the reaction after it has been allowed to proceed for a certain time and the composition is analyzed at leisure. In the chemical quench flow method, the reactants are mixed in much the same way as in the flow method, but the reaction is quenched by another reagent, such as a solution of acid or base, after the mixture has traveled along a fixed length of the outlet tube. Different reaction times can be selected by varying the flow rate along the outlet tube. An advantage of the chemical quench flow method over the stopped-flow method is that spectroscopic fingerprints are not needed in order to measure the concentration of reactants and products. Once the reaction has been quenched, the solution may be examined by rather ‘slow’ techniques, such as gel electrophoresis, mass spectrometry, and chromatography. In the freeze quench method, the reaction is quenched by cooling the mixture within milliseconds, and the concentrations of reactants, intermediates, and products are measured spectroscopically.

6.1 The definition of reaction rate The concepts introduced here for the description of reaction rates are used whenever we explore such biological processes as enzymatic transformations, electron transfer reactions in metabolism, and the transport of molecules and ions across membranes.

The average rate of a reaction is defined in terms of the rate of change of the concentration of a designated species: average rate =

| D[J] | Dt

Definition of average rate

(6.2a)

Fig. 6.2 In the stopped-flow technique the reagents are driven quickly into the mixing chamber and then the time dependence of the concentrations is monitored.

222

6 THE RATES OF REACTIONS

The rate of a chemical reaction is the slope (without the sign) of the tangent to the curve showing the variation of concentration of a species with time. This graph is a plot of the concentration of a reactant, which is consumed as the reaction progresses. The rate of consumption decreases in the course of the reaction as the concentration of reactant decreases.

Fig. 6.3

where D[J] is the change in the molar concentration of the species J that occurs during the time interval Dt. We have put the change in concentration between modulus signs (that is, the instruction to disregard the sign of the change) to ensure that all rates are positive: if J is a reactant, its concentration will decrease and D[J] itself is negative but | D[J] | is positive. With the concentration measured in moles per cubic decimeter (moles per liter) and the time in seconds, the average rate is reported in moles per cubic decimeter per second (mol dm−3 s−1). Because the rates at which reactants are consumed and products are formed typically change in the course of a reaction, it is necessary to consider the instantaneous rate, v, of the reaction, its rate at a specific instant. The instantaneous rate of consumption of a reactant is the slope of a graph of its molar concentration plotted against the time, with the slope evaluated as the tangent to the graph at the instant of interest (Fig. 6.3) and reported as a positive quantity. The instantaneous rate of formation of a product is also the slope of the tangent to the graph of its molar concentration plotted and also reported as a positive quantity. More formally: instantaneous rate =

| d[J] | dt

Definition of instantaneous rate

(6.2b)

In general, the various reactants in a given reaction are consumed at different rates, and the various products are also formed at different rates. However, these rates are related by the stoichiometry of the reaction. For example, in the decomposition of urea, (NH2)2CO, in acidic solution (NH2)2CO(aq) + 2 H2O(l) → 2 NH 4+(aq) + CO32−(aq) provided any intermediates are not present in significant quantities, the rate of formation of NH 4+ is twice the rate of disappearance of (NH2)2CO because for 1 mol (NH2)2CO consumed, 2 mol NH 4+ is formed: rate of formation of NH 4+ = 2 × rate of consumption of (NH2)2CO or, in terms of derivatives, v=

d[NH4+] d[(NH2)2CO] = −2 dt dt

One consequence of this kind of relation is that we have to be careful to specify exactly what species we mean when we report a reaction rate. Self-test 6.1 The rate of formation of NH3 in the reaction N2(g) + 3 H2(g) → 2 NH3(g) was reported as 1.2 mmol dm−3 s−1 under a certain set of conditions. What is the rate of consumption of H2?

Answer: 1.8 mmol dm−3 s−1

The problem of having a variety of different rates for the same reaction is avoided by bringing the stoichiometric coefficients into the definition of the rate. Thus, for a reaction of the type aA+bB→cC+dD we write the unique reaction rate as any of the four following quantities: v=

1 d[D] 1 d[C] 1 d[A] 1 d[B] = =− =− d dt c dt a dt b dt

Now there is a single rate for the reaction.

Definition of unique rate

(6.3)

6.2 RATE LAWS AND RATE CONSTANTS

A brief illustration

For the decomposition of urea in the reaction specified above (with a = 1, c = 2, and d = 1 (the water is in excess and its abundance would not be measured)), the unique reaction rate would be calculated from any of the following three quantities: v=

1 d[NH 4+] d[CO32−] d[(NH2)2CO] = =− 2 dt dt dt

6.2 Rate laws and rate constants The observed dependence of rate on the composition of the reaction mixture is often exploited for the purpose of slowing down some processes and speeding up others; it is also a window on the underlying mechanism of the reaction.

An empirical observation of the greatest importance is that the rate of reaction is often found to be proportional to the molar concentrations of the reactants raised to a simple power. For example, it may be found that the rate is directly proportional to the concentrations of the reactants A and B, so v = kr[A][B]

(6.4)

The coefficient kr, which is characteristic of the reaction being studied, is called the rate constant. The rate constant is independent of the concentrations of the species taking part in the reaction but depends on the temperature. An experimentally determined equation of this kind is called the ‘rate law’ of the reaction. More formally: A rate law is an equation that expresses the rate of reaction in terms of the molar concentrations (or partial pressures) of the species in the overall reaction (including, possibly, the products). The units of kr are always such as to convert the product of concentrations into a rate expressed as a change in concentration divided by time. For example, if the rate law is the one shown above, with concentrations expressed in moles per cubic decimeter (mol dm−3), then the units of kr will be cubic decimeters per mole per second (dm3 mol−1 s−1) because v

1 4 2 4 3

[B]

1 2 3

[A]

1 2 3

1 4 2 4 3

kr

dm mol s × mol dm × mol dm = mol dm−3 s−1 3

−1 −1

−3

−3

In gas-phase studies, such as those used to study reactions in planetary atmospheres, concentrations are commonly expressed in molecules per cubic centimeter (molecules cm−3), so the rate constant for the reaction above would be expressed in cm3 molecule−1 s−1. We can use the same approach to determine the units of the rate constant from rate laws of any form. For example, the rate constant for a reaction with a rate law of the form kr[A] is commonly expressed in s−1.

A reaction has a rate law of the form k r[A]2[B]. What are the units of the rate constant kr if the reaction rate is measured in mol dm−3 s−1? Self-test 6.2

Answer: dm6 mol−2 s−1

223

224

6 THE RATES OF REACTIONS

Once the rate law and the rate constant of the reaction have been determined, we can predict the rate of the reaction for any given composition of the reaction mixture. We shall also see that we can use the observed rate law to predict the concentrations of the reactants and products at any time after the start of the reaction. Furthermore, a rate law is an important guide to the mechanism of the reaction, the individual molecular steps by which it takes place, for any proposed mechanism must be consistent with the observed rate law. 6.3 Reaction order Once a reaction has been classified according to its rate law, we can use the same expressions to predict the composition of the reaction mixture at any stage of the reaction: specifically, many enzyme-catalyzed reactions and biological electron transfer reactions are kinetically similar.

Many reactions can be classified on the basis of their order, the power to which the concentration of a species is raised in the rate law: first order in A:

v = kr[A]

(6.5a)

first order in A, first order in B:

v = kr[A][B]

(6.5b)

second order in A:

v = kr[A]2

(6.5c)

The overall order of a reaction is the sum of the orders of all the components. The rate laws in eqns 6.5b and 6.5c both correspond to reactions that are second order overall.

A brief illustration

The re-formation of a DNA double helix after the double helix has been separated into two strands by raising the temperature or the pH: strand + complementary strand → double helix is found to obey the rate law v = kr[strand][complementary strand] This reaction is first order in each strand and second order overall. The reduction of nitrogen dioxide by carbon monoxide, NO2(g) + CO(g) → NO(g) + CO2(g) is found to obey the rate law v = kr[NO2]2 which is second order in NO2 and, because no other species occurs in the rate law, second order overall. The rate of the latter reaction is independent of the concentration of CO provided that some CO is present. It is therefore zero order in CO because a concentration raised to the power zero is 1 ([CO]0 = 1, just as x 0 = 1 in algebra).

A reaction need not have an integral order, and many gas-phase reactions do not. For example, if a reaction is found to have the rate law

6.4 THE DETERMINATION OF THE RATE LAW v = kr[A]1/2[B]

(6.6)

then it is half order in A, first order in B, and three-halves order overall. If a rate law is not of the form [A]x[B]y[C]z . . . , then the reaction does not have an overall order. For example, a typical rate law for the action of an enzyme E on a substrate S is (see Chapter 8) v=

kr[E][S] [S] + KM

(6.7a)

where KM is a constant (not a rate constant). This rate law is first order in the enzyme but does not have a specific order with respect to the substrate. Under certain circ*mstances a complicated rate law without an overall order may simplify into a law with a definite order. For example, if the substrate concentration in the enzyme-catalyzed reaction is so low that [S] KM, the rate law in eqn 6.7a becomes v = kr[E], [S] >> KM

(6.7c)

and the reaction is first order in E and zeroth order in S. It is very important to note that a rate law is established experimentally and cannot in general be inferred from the chemical equation for the reaction. The reaction of an enzyme with a substrate, for example, has a very simple stoichiometry, but its rate law (eqn 6.7a) is moderately complicated. In some cases, however, the rate law does happen to reflect the reaction stoichiometry. This is the case with the re-naturation of DNA in the brief illustration. 6.4 The determination of the rate law Because reaction order is such an important concept for the classification and investigation of biochemical reactions, we need to know how it is determined experimentally.

At the simplest level, a quick comparison of rates with two different concentrations of a reactant can indicate the order of the reaction. Thus, if the rate doubles when the concentration is doubled, we can infer that the reaction is first order in that reactant, and if it quadruples, then it is second order. However, to assess the data more fully, we need to be more systematic. There are two principal approaches. In one, we use rate measurements directly; in the other, we use concentration measurements, not rates. In this section we prepare the ground for the first approach and describe its implementation. The second approach needs more preparation and is described in Section 6.5. (a) Isolation and pseudo-order reactions

The determination of a rate law is simplified by the isolation method, in which all the reactants except one are present in large excess. The dependence of the rate on each of the reactants can be found by isolating each of them in turn—focusing on a single species by having all the others present in large excess—and piecing together a picture of the overall rate law.

225

226

6 THE RATES OF REACTIONS

If a reactant B is in large excess it is a good approximation to take its concentration as constant throughout the reaction. Then, although the true rate law might be v = kr[A][B]2, we can approximate [B] by its initial value [B]0 (from which it hardly changes in the course of the reaction) and write A pseudo-first-order reaction, B in excess

v = k r′[A] with k r′ = kr[B]02

(6.8a)

Because the true rate law has been forced into first-order form by assuming a constant B concentration, the effective rate law is classified as pseudo-first order and k ′r is called the effective rate constant for a given, fixed concentration of B. If, instead, the concentration of A were in large excess, and hence effectively constant, then the rate law v = k r[A][B]2 would simplify to A pseudo-second-order reaction, A in excess

v = k r″[B]2 with k r″ = k r[A]0

(6.8b)

This pseudo-second-order rate law is also much easier to analyze and identify than the complete law. In a similar manner, a reaction may appear to be zeroth order. For instance, the oxidation of ethanol to acetaldehyde (ethanal) by NAD+ in the liver in the presence of the enzyme liver alcohol dehydrogenase, CH3CH2OH(aq) + NAD+(aq) + H2O(l) → CH3CHO(aq) + NADH(aq) + H3O+(aq) is zeroth order overall as the ethanol is in excess and the concentration of the NAD+ is maintained at a constant level by normal metabolic processes. Many reactions in aqueous solution that are reported as first or second order are actually pseudo-first or pseudo-second order: the solvent water participates in the reaction, but it is in such large excess that its concentration remains constant. (b) The method of initial rates

In the method of initial rates, which is often used in conjunction with the isolation method, the instantaneous rate is measured at the beginning of the reaction for several different initial concentrations of reactants. If the initial rate doubles when the initial concentration of the isolated reactant is doubled, then the reaction is first order in that reactant, and so on. To use the data more fully we suppose that the rate law for a reaction with A isolated is v = k r′[A]a, then the initial rate of the reaction, v0, is given by the initial concentration of A: v0 = k r′[A]0a

Initial rate of an ath-order reaction

(6.9)

1

Taking logarithms gives log v0 = log k r′ + a log [A]0

(6.10)

This equation has the form of the equation for a straight line: y = intercept + slope × x with y = log v0 and x = log [A]0. It follows that, for a series of initial concentrations, a plot of the logarithms of the initial rates against the logarithms of the initial concentrations of A should be a straight line and that the slope of the graph will be 1

For a review of logarithms, see Mathematical toolkit 5.1.

6.4 THE DETERMINATION OF THE RATE LAW

227

a, the order of the reaction with respect to the species A, and log k r′ is given by the intercept at log [A]0 = 0 (Fig. 6.4). An important point to note is that the method of initial rates might not reveal the entire rate law, for in a complex reaction we may not be able to specify an order with respect to a reactant (see eqn 6.7a) or the products themselves might affect the rate.

Example 6.1

Using the method of initial rates

The following data were obtained on the initial rate of binding of glucose to the enzyme hexokinase: [glucose]0 /(mmol dm−3) V0 /(mol dm−3 s−1)

1.00 5.0 7.0 21.0

(a) (b) (c)

1.54 7.6 11.0 34.0

3.12 15.5 23.0 70.0

4.02 20.0 31.0 96.0

The plot of log v0 (shifted by log k r′) against log [A]0 gives straight lines with slopes equal to the order of the reaction.

Fig. 6.4

The enzyme concentrations are (a) 1.34 mmol dm−3, (b) 3.00 mmol dm−3, and (c) 10.0 mmol dm−3. Find the orders of reaction with respect to glucose and hexokinase and the rate constant. Strategy We assume that the initial rate law has the form

v0 = kr[glucose]0a[hexokinase]b0 For constant [hexokinase]0, the initial rate law has the form v0 = k r′[glucose]0a, with k r′ = kr[hexokinase]b0, so log v0 = log k r′ + a log [glucose]0 We need to make a plot of log v0 against log [glucose]0 for a given [hexokinase]0 and find the reaction order a from the slope and the value of k r′ from the intercept at log [glucose]0 = 0. Then, because log k r′ = log kr + b log [hexokinase]0 plot log k r′ against log [hexokinase]0 to find log kr from the intercept and b from the slope. Solution The data give the following points for the graph:

log ([glucose]0 /mol dm−3) log (V0 /mol dm−3 s−1)

(a) (b) (c)

−3.00 0.699 0.844 1.32

−2.81 0.881 1.04 1.53

−2.51 1.19 1.36 1.85

−2.40 1.30 1.49 1.98

The graph of the data is shown in Fig. 6.5. The slopes of the lines are 1, so a = 1, and the effective rate constants kr are as follows: [hexokinase]0 /(mol dm−3) log ([hexokinase]0 /mol dm−3) log (k r′/dm3 mol−1 s−1)

1.34 × 10−3 −2.87 3.69

3.00 × 10−3 −2.52 4.04

1.00 × 10−2 −2.00 4.56

Figure 6.6 is the plot of log k r′ against log [hexokinase]0. The slope is 1, so b = 1. The intercept at log [hexokinase]0 = 0 is log kr = 6.56, so kr = 3.6 × 106 dm3 mol−1 s−1. The overall (initial) rate law is v0 = kr[glucose]0[hexokinase]0

Fig. 6.5 The plots of the data in Example 6.1 for finding the order with respect to glucose.

A note on good practice

When taking the logarithm of a number of the form x.xx × 10n, there are four significant figures in the answer: the figure before the decimal point is simply the power of 10. Strictly, the logarithms are of the quantity divided by its units.

228

6 THE RATES OF REACTIONS

The initial rate of a certain reaction depended on concentration of a substance J as follows: Self-test 6.3

[J]0 /(mmol dm−3) V0 /(10−7 mol dm−3 s−1)

5.0 3.6

10.2 9.6

17 41

30 130

Find the order of the reaction with respect to J and the rate constant. Answer: 2; 1.6 × 10−2 dm3 mol−1 s−1

6.5 Integrated rate laws The rate laws summarize useful information about the progress of a reaction and allow us to predict the composition of a reaction mixture at any time, including the concentrations of biochemically significant intermediates. Fig. 6.6 The plots of the data in Example 6.1 for finding the order with respect to hexokinase.

A rate law tells us the rate of the reaction at a given instant (when the reaction mixture has a particular composition). This is rather like being given the speed of a car at each point of its journey. For a car journey, we may want to know the distance that a car has traveled at a certain time given its varying speed. Similarly, for a chemical reaction, we may want to know the composition of the reaction mixture at a given time given the varying rate of the reaction. An integrated rate law is an expression that gives the concentration of a species as a function of the time. Integrated rate laws have two principal uses. One is to predict the concentration of a species at any time after the start of the reaction. Another is to help find the rate constant and order of the reaction. Indeed, although we have introduced rate laws through a discussion of the determination of reaction rates, these rates are rarely measured directly because slopes are so difficult to determine accurately. Almost all experimental work in chemical kinetics deals with integrated rate laws; their great advantage being that they are expressed in terms of the experimental observables of concentration and time. Computers can be used to find numerical solutions of even the most complex rate laws. However, we now see that in a number of simple cases, solutions can be expressed as relatively simple functions and prove to be very useful. (a) Zeroth-order reactions

For a chemical reaction and zeroth-order rate law of the form A → products, v = −

d[A] = kr dt

Zeroth-order rate law

(6.11a)

the concentration of A falls linearly until all A has been consumed: [A] = [A]0 − krt for krt ≤ [A]0 [A] = 0 for krt > [A]0

Zeroth-order integrated rate law

(6.11b)

(b) First-order reactions

For a chemical reaction with a first-order rate law of the form A → products, v = −

d[A] = kr[A] dt

First-order rate law

(6.12a)

6.5 INTEGRATED RATE LAWS

229

we show in the following Justification that the integrated rate law is ln

[A] = −krt [A]0

First-order integrated rate law

(6.12b)

where [A]0 is the initial concentration of A. Two alternative forms of this expression are ln [A] = ln [A]0 − krt

(6.12c)

[A] = [A]0e−k t

(6.12d)

r

Equation 6.12d has the form of an exponential decay (Fig. 6.7). A common feature of all first-order reactions, therefore, is that the concentration of the reactant decays exponentially with time. Justification 6.1 First-order integrated rate laws

A first-order rate equation has the form −

d[A] = kr[A] dt

Fig. 6.7 The exponential decay of the reactant in a first-order reaction. The greater the rate constant, the more rapid is the decay.

and is an example of a ‘first-order differential equation’. Because the terms d[A] and dt may be manipulated like any algebraic quantity, we rearrange the differential equation into d[A] = −krdt [A] and then integrate both sides. Integration from t = 0, when the concentration of A is [A]0, to the time of interest, t, when the molar concentration of A is [A], is written as

[A]

[A]0

d[A] = −kr [A]

冮 dt t

We now use the standard integral

冮 dxx = ln x + constant and obtain the expression ln [A] − ln [A]0 = −krt which rearranges into eqn 6.12c.

Equation 6.12d lets us predict the concentration of A at any time after the start of the reaction. However, we can also use the result to confirm that a reaction is first order and to determine the rate constant: eqn 6.12c shows that if we plot ln [A] against t, then we will get a straight line if the reaction is first order (Fig. 6.8). If the experimental data do not give a straight line when plotted in this way, then the reaction is not first order. If the line is straight, then it follows from the same equation that its slope is −kr, so we can also determine the rate constant from the graph.

Fig. 6.8 The determination of the rate constant of a first-order reaction. A straight line is obtained when ln c is plotted against t; the slope is −kr. The data are from Case study 6.1. The structure shown is that of propranolol, one of the first b-blockers.

230

6 THE RATES OF REACTIONS

Mathematical toolbox 6.1 Differential equations

An ordinary differential equation is a relation between derivatives of a function of one variable and the function itself, as in a

d2y dy + b + cy + d = 0 dx 2 dx

The coefficients a, b, etc., may be functions of x. The order of the equation is the order of the highest derivative that occurs in it, so eqn 6.12a is a first-order equation and the expression above is a second-order equation. ‘Solving’ a differential equation is the process of determining the function, in this case y(x), that satisfies it.

In many cases it is found that various constants appear in the solution, such as y(x) + constant. These constants are determined by imposing various boundary conditions on the solutions, values that the solution must have at specified points. A second-order differential equation requires two boundary conditions, a first-order equation requires one. For timedependent solutions, the boundary condition is termed an initial condition, and is typically the value that the solution must have at t = 0.

A useful indication of the rate of a first-order chemical reaction is the half-life, t1/2, of a reactant, which is the time it takes for the concentration of the species to fall to half its initial value. We can find the half-life of a species A that decays in a first-order reaction (eqn 6.12a) by substituting [A] = 12[A]0 and t = t1/2 into eqn 6.12b: 1 [A] krt1/2 = −ln 2 0 = −ln 12 = ln 2 [A]0

it follows that t1/2 =

ln 2 kr

Half-life of a first-order reaction

(6.13)

A brief illustration

Because the rate constant for the first-order denaturation of hemoglobin is 2.00 × 10−4 s−1 at 60°C, the half-life of properly folded hemoglobin is 57.7 min. Hence, the concentration of folded hemoglobin falls to half its initial value in 57.7 min, and then to half that concentration again in a further 57.7 min, and so on (Fig. 6.9).

The main point to note about eqn 6.13 is that for a first-order reaction, the half-life of a reactant is independent of its concentration. It follows that if the concentration of A at some arbitrary stage of the reaction is [A], then the concentration will fall to 12[A] after an interval of (ln 2)/kr whatever the actual value of [A] (Fig. 6.10).

A brief illustration

The molar concentration of properly folded hemoglobin after a succession of half-lives.

Fig. 6.9

In acidic solution, the disaccharide sucrose (cane sugar, Atlas S5) is converted to a mixture of the monosaccharides glucose (Atlas S4), and fructose (Atlas S3) in a pseudo-first-order reaction. Under certain conditions of pH, the

6.5 INTEGRATED RATE LAWS

231

half-life of sucrose is 28.4 min. To calculate how long it takes for the concentration of a sample to fall from 8.0 mmol dm−3 to 1.0 mmol dm−3, we note that molar concentration/(mmol dm−3): 8.0 ffg 4.0 ffg 2.0 ffg 1.0 28.4 min

28.4 min

28.4 min

The total time required is 3 × 28.4 min = 85.2 min.

Self-test 6.4 The half-life of a substrate in a certain enzyme-catalyzed firstorder reaction is 138 s. How long does it take for the concentration of substrate to fall from 1.28 mmol dm−3 to 0.040 mmol dm−3?

Answer: 690 s

Another indication of the rate of a first-order reaction is the time constant, t, the time required for the concentration of a reactant to fall to 1/e of its initial value. From eqn 6.12b it follows that krt = −ln

A [A]0/eD 1 = −ln = 1 C [A]0 F e

Hence, the time constant is the reciprocal of the rate constant: t=

1 kr

Time constant of a first-order decay

(6.14)

The longer the time constant of a first-order reaction, the slower the decay and the longer the reactants survive. (c) Second-order reactions

Now we need to see how concentration varies with time for a reaction with a second-order rate law of the form A → products

v=−

d[A] = kr[A]2 dt

Second-order rate law

(6.15a)

As before, we suppose that the concentration of A at t = 0 is [A]0 and, as shown in the following Justification, find that 1 1 − = −krt [A]0 [A]

Second-order integrated rate law

(6.15b)

Two alternative forms of eqn 6.15b are 1 1 = + krt [A] [A]0 [A] =

[A]0 1 + krt[A]0

Justification 6.2 Second-order integrated rate laws I

To solve the differential equation −

d[A] = kr[A]2 dt

(6.15c) (6.15d)

In each successive period of duration t1/2, the concentration of a reactant in a first-order reaction decays to half its value at the start of that period. After n such periods, the concentration is ( 12 )n of its initial concentration.

Fig. 6.10

232

6 THE RATES OF REACTIONS

we rearrange it into d[A] = krdt [A]2

and integrate it between t = 0, when the concentration of A is [A]0, and the time of interest t, when the concentration of A is [A]:

[A]

d[A] = −kr [A]2

[A]0

冮 dt t

The term on the right is −krt. We evaluate the integral on the left by using the standard form

冮dxx = − 1x + constant 2

which implies that

冮 dxx = !@− 1x + constant#$ − !@− 1x + constant#$ b

2

b

a

1 1 =− + b a a

and so obtain eqn 6.15b.

The determination of the rate constant of a second-order reaction. A straight line is obtained when 1/[A] is plotted against t; the slope is kr.

Fig. 6.11

Equation 6.15b shows that to test for a second-order reaction, we should plot 1/[A] against t and expect a straight line. If the line is straight, then the reaction is second order in A and the slope of the line is equal to the rate constant (Fig. 6.11). Equation 6.15c enables us to predict the concentration of A at any time after the start of the reaction (Fig. 6.12). We see that the concentration of A approaches zero more slowly in a second-order reaction than in a first-order reaction with the same initial rate (Fig. 6.13). It follows from eqn 6.15a by substituting t = t1/2 and [A] = 12[A]0 that the half-life of a species A that is consumed in a second-order reaction is t1/2 =

1 kr[A]0

Half-life for a secondorder reaction

(6.16)

Therefore, unlike a first-order reaction, the half-life of a substance in a secondorder reaction varies with the initial concentration. A practical consequence of this dependence is that species that decay by second-order reactions (which includes some environmentally harmful substances) may persist in low concentrations for long periods because their half-lives are long when their concentrations are low. Another type of second-order reaction is one that is first order in each of two reactants A and B: d[A] = −kr[A][B] dt The variation with time of the concentration of a reactant in a second-order reaction.

Fig. 6.12

Overall second-order rate law

(6.17a)

We have already seen that the rate of formation of DNA from two complementary strands can be modeled by this rate law. We cannot integrate eqn 6.17a until we know how the concentration of B is related to that of A. For example, if the

6.5 INTEGRATED RATE LAWS

Mathematical toolkit 6.2

Integration by the method of partial fractions

Then we integrate each term on the right by using the standard integral already given in Justification 6.1. It follows that

To solve an integral of the form I=

233

冮(a − x)(b1 − x) dx

where a and b are constants, we use the method of partial fractions in which a fraction that is the product of terms (as in the denominator of this integrand) is written as a sum of fractions. To implement this procedure we write the integrand as

I= =

1 ⎡ dx dx ⎤ − ⎢ ⎥ b−a⎣ a−x b − x⎦

1 A 1 1 D ln − ln + constant b−aC a−x b − xF

1 1 A 1 1 D = − (a − x)(b − x) b − a C a − x b − xF

reaction is A + B → P, where P denotes products and the initial concentrations are [A]0 and [B]0, then we show in the following Justification that at a time t after the start of the reaction, the concentrations satisfy the relation ln

A [B]/[B]0 D = ([B]0 − [A]0)krt C [A][A]0 F

Integrated overall second-order rate law

(6.17b)

Therefore, a plot of the expression on the left against t should be a straight line from which kr can be obtained. Note that if [A]0 = [B]0, then the solutions are those already given in eqn 6.13b (but this solution cannot be found simply by setting [A]0 = [B]0 in eqn 6.17b.)

Justification 6.3 Second-order integrated rate laws II

It follows from the reaction stoichiometry that when the concentration of A has fallen to [A]0 − x, the concentration of B will have fallen to [B]0 − x because each A that disappears entails the disappearance of one B. It follows that d[A] = −kr([A]0 − x)([B]0 − x) dt Then, because [A] = [A]0 − x and d[A]/dt = −dx/dt, the rate law is dx = kr([A]0 − x)([B]0 − x) dt The initial condition is that x = 0 when t = 0; so the integration required is

冮 dt

x

dx = kr ([A] − x)([B] 0 0 − x) 0

t

The integral on the right is simply krt. The integral on the left is evaluated by using the method of partial fractions (see Mathematical toolkit 6.2): dx 1 ! ln A [A] D − ln A [B] D # = 冮 ([A] − x)([B] C [B] − x)F $ − x) [B] − [A] @ C [A] − x F x

Although the initial decay of a second-order reaction may be rapid, later the concentration approaches zero more slowly than in a first-order reaction with the same initial rate (compare Fig. 6.7).

Fig. 6.13

234

6 THE RATES OF REACTIONS

The two logarithms can be combined as follows: ln

A [A]0 D A [B]0 D − ln = ln [A]0 − ln ([A]0 − x) − ln [B]0 + ln ([B]0 − x) C [A]0 − x F C [B]0 − x F = ln [A]0 − ln [A] − ln [B]0 + ln [B] = {ln [B] − ln [B]0} − {ln [A] − ln [A]0} = ln

A [B] D A [A] D − ln C [B]0F C [A]0F

= ln

A [B]/[B]0 D C [A]/[A]0F

where we have used [A] = [A]0 − x and [B] = [B]0 − x. Combining all the results so far gives eqn 6.17b.

Equation 6.17b can be rearranged to give the concentration of either reactant. To do this, we substitute [A] = [A]0 − x and [B] = [B]0 − x into the equation written in the form [B][A]0 ([B] −[A] )k t =e [A][B]0 0

r

and obtain [A]0([B]0 − x) ([B] −[A] )k t =e [B]0([A]0 − x) 0

r

This expression is then solved for x: x=

[A]0[B]0(e([B] −[A] )k t − 1) [B]0e([B] −[A] )k t − [A]0 0

r

r

At this point, we can form either [A] or [B]. For instance, from [A] = [A]0 − x we find [A] =

[A]0([B]0 − [A]0) [B]0e([B] −[A] )k t − [A]0 0

(6.18)

r

The time dependence of [A] and [B] is illustrated in Fig. 6.14.

Case study 6.1

The time dependence of the concentrations of the reactants in a reaction with the overall second-order rate law in eqn 6.17a. We have taken [B]0 = 2[A]0.

Fig. 6.14

Pharmaco*kinetics

Pharmaco*kinetics is the study of the rates of absorption and elimination of drugs by organisms. In most cases, elimination is slower than absorption and is a more important determinant of availability of a drug for binding to its target. A drug can be eliminated by many mechanisms, such as metabolism in the liver, intestine, or kidney followed by excretion of breakdown products through urine or feces. As an example of pharmaco*kinetic analysis, consider the elimination of b-adrenergic blocking agents (‘b-blockers’), drugs used in the treatment of hypertension. After intravenous administration of a b-blocker, the blood plasma of a patient was analyzed for remaining drug, and the data are shown

6.6 THE ARRHENIUS EQUATION

235

below, where c is the mass concentration of the drug measured at a time t after the injection. t/min c/(ng cm−3)

30 699

60 622

120 413

150 292

240 152

360 60

480 24

To see if the removal is a first-order process, we draw up the following table: t/min ln (c/(ng cm−3))

30 6.550

60 6.433

120 6.023

150 5.677

240 5.024

360 4.09

480 3.18

The graph of the data is shown in Fig. 6.8. The plot is straight, confirming a first-order process. Its least-squares best-fit slope is −7.6 × 10−3, so kr = 7.6 × 10−3 min−1 and t1/2 = 91 min at 310 K, body temperature. Most drugs are eliminated from the body by a first-order process. An essential aspect of drug development is the optimization of the half-life of elimination, which needs to be long enough to allow the drug to find and act on its target organ but not so long that harmful side effects become important.

The temperature dependence of reaction rates The rates of most chemical reactions increase as the temperature is raised. Many organic reactions in solution lie somewhere in the range spanned by the hydrolysis of methyl ethanoate (for which the rate constant at 35°C is 1.8 times that at 25°C) and the hydrolysis of sucrose (for which the factor is 4.1). Reactions in the gas phase typically have rates that are only weakly sensitive to the temperature. Enzyme-catalyzed reactions may show a more complex temperature dependence because raising the temperature may provoke conformational changes and even denaturation and degradation, which lower the effectiveness of the enzyme. We saw in the discussion of the hydrophobic effect (Section 2.7) that lowering the temperature can also result in denaturation, so an enzyme may lose its effectiveness at low temperatures too. 6.6 The Arrhenius equation The balance of reactions in organisms depends strongly on the temperature: that is one function of a fever, which modifies reaction rates in the infecting organism and hence destroys it. To discuss the effect quantitatively, we need to know the factors that make a reaction rate more or less sensitive to temperature.

As data on reaction rates were accumulated toward the end of the nineteenth century, the Swedish chemist Svante Arrhenius noted that almost all of them showed a similar dependence on temperature. In particular, he noted that a graph of ln kr, where kr is the rate constant for the reaction, against 1/T, where T is the (absolute) temperature at which kr is measured, gives a straight line with a slope that is characteristic of the reaction (Fig. 6.15). The mathematical expression of this conclusion is that the rate constant varies with temperature as ln kr = intercept + slope ×

1 T

Empirical temperature dependence

(6.19a)

Fig. 6.15 The general form of an Arrhenius plot of ln kr against 1/T. The slope is equal to −Ea /R and the intercept at 1/T = 0 is equal to ln A.

236

6 THE RATES OF REACTIONS

This expression is normally written as the Arrhenius equation: ln kr = ln A −

Ea RT

Arrhenius equation

(6.19b)

or alternatively as kr = Ae−E /RT

(6.19c)

a

These Arrhenius plots correspond to three different activation energies. Note the fact that the plot corresponding to the higher activation energy indicates that the rate of that reaction is more sensitive to temperature.

Fig. 6.16

The parameter A (which has the same units as kr) is called the pre-exponential factor and the parameter Ea (which is a molar energy and normally expressed as kilojoules per mole) is called the activation energy. Collectively, A and Ea are called the Arrhenius parameters of the reaction. A practical point to note by comparing eqns 6.19a and 6.19b is that a high activation energy corresponds to a reaction rate that is very sensitive to temperature (the Arrhenius plot has a steep slope, Fig. 6.16). Conversely, a small activation energy indicates a reaction rate that varies only slightly with temperature (the slope is shallow). A reaction with zero activation energy (so kr = A), such as for some radical recombination reactions in the gas phase, has a rate that is largely independent of temperature. Marked deviations from a straight line at high or low temperature may indicate that an enzyme has lost its effectiveness through denaturation.

Example 6.2

Determining the Arrhenius parameters

The rate constant of the acid hydrolysis of sucrose discussed in Section 6.6b varies with temperature as follows. Find the activation energy and the preexponential factor. T/K kr /(10−3 s−1)

297 4.8

301 7.8

305 13

309 20

313 32

Strategy We plot ln kr against 1/T and expect a straight line. The slope is −Ea /R and the intercept of the extrapolation to 1/T = 0 is ln A. It is best to do a leastsquares fit of the data to a straight line. Note that, as remarked in the text, A has the same units as kr. Solution The Arrhenius plot is shown in Fig. 6.17. The least-squares best fit of the line has slope −1.10 × 104 and intercept 31.7 (which is well off the graph), therefore

Ea = −R × slope = −(8.3145 J K−1 mol−1) × (−1.10 × 104 K) = 91.5 kJ mol−1 and A = e31.7 s−1 = 5.8 × 1013 s−1

The Arrhenius plot for the acid hydrolysis of sucrose, and the best (least-squares) straight line fitted to the data points. The data are from Example 6.2.

Fig. 6.17

Self-test 6.5

Determine A and Ea from the following data:

T/K kr /(dm3 mol−1 s−1)

300 7.9 × 106

350 3.0 × 107

400 7.9 × 107

450 1.7 × 108

500 3.2 × 108

Answer: 8 × 1010 dm3 mol−1 s−1, 23 kJ mol−1

6.7 PRELIMINARY INTERPRETATION OF THE ARRHENIUS PARAMETERS

Once the activation energy of a reaction is known, it is a simple matter to predict the value of a rate constant kr,2 at a temperature T2 from its value kr,1 at another temperature T1 that lies within the functional range of the enzyme. To do so, we write ln kr,2 = ln A −

Ea RT2

and then subtract eqn 6.19b (with T identified as T1 and kr as kr,1), so obtaining ln kr,2 − ln kr,1 = −

Ea E + a RT2 RT1

and therefore ln

kr,2 Ea A 1 1 D = − kr,1 R C T1 T2 F

Temperature dependence of the rate constant

(6.20)

A brief illustration

For a reaction with an activation energy of 50 kJ mol−1, an increase in the temperature from 25°C to 37°C (body temperature) corresponds to ln

kr,2 50 × 103 J mol−1 A 1 1 D 50 × 103 A 1 1 D = − = − −1 −1 C F C kr,1 8.3145 J K mol 298 K 310 K 8.3145 298 310F

By taking natural antilogarithms (that is, by forming e x ), kr,2 = 2.18kr,1. This result corresponds to slightly more than a doubling of the rate constant.

Self-test 6.6 The activation energy of one of the reactions in the citric acid cycle (Case study 4.3) is 87 kJ mol−1. What is the change in rate constant when the temperature falls from 37°C to 15°C?

Answer: kr,2 = 0.076kr,1

6.7 Preliminary interpretation of the Arrhenius parameters Once we know the molecular interpretation of the pre-exponential factor and the activation energy, we can identify the strategies that special biological macromolecules adopt to accelerate and regulate the rates of biochemical reactions.

The simplest interpretation of a chemical reaction is that it takes place when two molecules collide either in a gas or, more relevantly to biology, as they move through a solution. We shall refine this picture in Chapter 7, but it is adequate for our present purpose. The pre-exponential factor A is a measure of the rate at which collisions occur, irrespective of their energy. More precisely, A is the constant of proportionality between the collision rate and the product of the molar concentrations of the reactants. To interpret Ea, we consider how the energy of the reactant molecules A and B (specifically, their total molecular potential energy) changes in the course of a collision. As the reaction proceeds, A and B come into contact, distort, and begin to exchange or discard atoms. The potential energy rises to a maximum and the

237

238

6 THE RATES OF REACTIONS

A potential energy profile for an exothermic reaction. The graph depicts schematically the changing potential energy of two species that approach, collide, and then go on to form products. The activation energy is the height of the barrier above the potential energy of the reactants.

Fig. 6.18

cluster of atoms that corresponds to the region close to the maximum is called the activated complex (Fig. 6.18). After the maximum, the potential energy falls as the atoms rearrange in the cluster and reaches a value characteristic of the products. The climax of the reaction is at the peak of the potential energy, which corresponds to the activation energy, Ea. Here two reactant molecules have come to such a degree of closeness and distortion that a small further distortion will send them in the direction of products. This crucial configuration is called the transition state of the reaction. Although some molecules entering the transition state might revert to reactants, if they pass through this configuration it is inevitable that products will emerge from the encounter.2 We can infer from the preceding discussion that to react when they meet, two reactant molecules must have sufficient kinetic energy to surmount the barrier and pass through the transition state. It follows that the activation energy is the minimum relative kinetic energy that reactants must have in order to form products. For example, in a gas phase reaction there are numerous collisions each second, but only a tiny proportion are sufficiently energetic to lead to reaction. Hence, the exponential factor in eqn 6.19c can be interpreted as the fraction of collisions that have enough kinetic energy to lead to reaction. An enzyme, and a catalyst in general, lowers the activation energy of the reaction either by providing a different reaction pathway with a lower activation energy or by lowering the energy of the transition state to make it more accessible (Fig. 6.19). Enzymes are very specific and can have a dramatic effect on the reactions they control. For example, the enzyme catalase reduces the activation energy for the decomposition of hydrogen peroxide from 76 kJ mol−1 to 8 kJ mol−1, corresponding to an acceleration of the reaction by a factor of 1012 at 298 K.

A brief illustration

The effect of catalase on the rate of decomposition of H2O2 can be assessed by evaluating the ratio of rate constants as follows: kr,catalyzed Ae−E = −E kr,uncatalyzed Ae = e(68 × 10

a,catalyzed

/RT

3

A catalyst acts by providing a new reaction pathway between reactants and products, with a lower activation energy than the original pathway.

= e−(E

a,catalyzed

a,uncatalyzed

/RT −1

J mol )/(8.3145 J K

−1

− Ea,uncatalyzed)/RT −1

mol ) × (298 K)

= 8.3 × 1011

Fig. 6.19

2 The terms activated complex and transition state are often used as synonyms; however, we shall preserve a distinction.

DISCUSSION QUESTIONS

239

Checklist of key concepts 1. The rates of chemical reactions are measured by using techniques that monitor the concentrations of species present in the reaction mixture (Table 6.1).

6. An integrated rate law is an expression for the rate of a reaction as a function of time. 7. The half-life t1/2 of a reaction is the time it takes for the concentration of a species to fall to half its initial value.

2. Techniques for the study of reactions include real-time and quenching procedures, flow and stopped-flow techniques, and flash photolysis.

8. The temperature dependence of the rate constant of a reaction typically follows the Arrhenius law.

3. The instantaneous rate of a reaction is the slope of the tangent to the graph of concentration against time (expressed as a positive quantity).

9. The greater the activation energy, the more sensitive the rate constant is to the temperature.

4. A rate law is an expression for the reaction rate in terms of the concentrations of the species that occur in the overall chemical reaction.

10. The activation energy is the minimum relative kinetic energy that reactants must have in order to form products; the pre-exponential factor is a measure of the rate at which collisions occur irrespective of their energy.

5. For a rate law of the form rate = kr[A]a[B]b . . . , the order with respect to A is a and the overall order is a+b+···.

Checklist of key equations Property

Equation

Comment

Half-life

t1/2 = (ln 2)/kr

First-order reaction

t1/2 = 1/kr[A]0

Second-order reaction

Arrhenius equation

ln kr = ln A − Ea /RT

Integrated rate laws

Zeroth order

[A] = [A]0 − krt for krt ≤ [A]0

First order

[A] = [A]0e

A → P v = kr[A]

Second order

[A] = [A]0/(1 + krt[A]0)

A → P v = kr[A]2

[A] = [A]0([B]0 − [A]0)/f(t)

A + B → P v = kr[A][B]

−krt

f(t) = [B]0e

([B]0−[A]0)krt

A → P v = kr

− [A]0

Discussion questions 6.1 Consult literature sources and list the observed timescales during which the following processes occur: proton transfer reactions, the initial event of vision, energy transfer in photosynthesis, the initial electron transfer events of photosynthesis, and the helix-to-coil transition in polypeptides. 6.2 Write a brief report on a recent research article in which at least

one of the following techniques was used to study the kinetics of a biochemical reaction: stopped-flow techniques, flash photolysis, chemical quench-flow methods, or freeze-quench methods. Your report should be similar in content and extent to one of the Case studies found throughout this book.

6.3 Describe the main features, including advantages and disadvantages, of the following experimental methods for determining the rate law of a reaction: the isolation method, the method of initial rates, and fitting data to integrated rate law expressions. 6.4 Distinguish between zeroth-order, first-order, second-order, and pseudo-first-order reactions. 6.5 Define the terms in and limit the generality of the expression ln kr = ln A − Ea/RT. 6.6 Provide molecular interpretations of the activation energy and the pre-exponential factor.

240

6 THE RATES OF REACTIONS

Exercises 6.7 The molar absorption coefficient of a substance dissolved in water is known to be 855 dm3 mol−1 cm−1 at 270 nm. To determine the rate of decomposition of this substance, a solution with a concentration of 3.25 mmol dm−3 was prepared. Calculate the percentage reduction in intensity when light of that wavelength passes through 2.5 mm of this solution. 6.8 The molar absorption coefficient of cytochrome P450, an enzyme involved in the breakdown of harmful substances in the liver and small intestine, at 522 nm is 291 dm3 mol−1 cm−1. In a study of its enzymatic activity, a solution of cytochrome P450 was prepared in a cell of length 6.5 mm. When light of 522 nm passes through this cell, 39.8 per cent of the light is absorbed. What is the molar concentration of cytochrome P450 in the solution?

6.15 The elimination of carbon dioxide from pyruvate ions by a decarboxylase enzyme was monitored by measuring the partial pressure of the gas as it was formed in a 250 cm3 flask at 293. In one experiment, the partial pressure increased from 0 to 100 Pa in 522 s in a first-order reaction when the initial concentration of pyruvate ions in 100 cm3 of solution was 3.23 mmol dm−3. What is the rate constant of the reaction? 6.16 Carbonic anhydrase is a zinc-based enzyme that catalyzes the

conversion of carbon dioxide to carbonic acid. In an experiment to study its effect, it was found that the molar concentration of carbon dioxide in solution decreased from 220 mmol dm−3 to 56.0 mmol dm−3 in 1.22 × 104 s. What is the rate constant of the first-order reaction?

6.9 (a) The rate of formation of C in the reaction 2 A + B → 3 C + 2 D is 2.2 mol dm−3 s−1. State the rates of formation and consumption of A, B, and D. (b) The rate law for this reaction was reported as rate = kr[A][B][C] with the molar concentrations in mol dm−3 and the time in seconds. What are the units of kr?

6.17 The formation of NOCl from NO in the presence of a large excess of chlorine is pseudo-second order in NO. When the initial pressure of NO was 300 Pa, the partial pressure of NOCl increased from zero to 100 Pa in 522 s. What is the rate constant of the reaction?

6.10 If the rate laws are expressed with (a) concentrations in numbers of molecules per cubic meter (molecules m−3) and (b) pressures in kilopascals, what are the units of the second-order and third-order rate constants?

6.18 The following data were obtained on the initial rate of isomerization of a compound S catalyzed by an enzyme E:

[S]0 /(mmol dm−3) V0 /(mmol dm−3 s−1)

(a) (b) (c)

1.00 4.5 14.8 58.9

2.00 9.0 25.0 120.0

3.00 15.0 45.0 180.0

4.00 18.0 59.7 238.0

6.11 The growth of microorganisms may be described in general terms as follows: (a) initially, cells do not grow appreciably; (b) after the initial period, cells grow rapidly with first-order kinetics; (c) after this period of growth, the number of cells reaches a maximum level and then begins to decrease. Sketch a plot of log(number of microorganisms) against t that reflects the kinetic behavior just described.

The enzyme concentrations are (a) 1.00 mmol dm−3, (b) 3.00 mol dm−3, and (c) 10.0 mmol dm−3. Find the orders of reactions with respect to S and E, and the rate constant.

6.12 Laser flash photolysis is often used to measure the binding

solution. An experiment on the hydrolysis of sucrose in 0.50 m HCl(aq) produced the following data:

rate of CO to heme proteins, such as myoglobin (Mb), because CO dissociates from the bound state relatively easily on absorption of energy from an intense and short pulse of light. The reaction is usually run under pseudo-first-order conditions. For a reaction in which [Mb]0 = 10 mmol dm−3, [CO] = 400 mmol dm−3, and the rate constant is 5.8 × 105 dm3 mol−1 s−1, plot a curve of [Mb] against time. The observed reaction is Mb + CO → MbCO. 6.13 The oxidation of ethanol to acetaldehyde (ethanal) by NAD+ in the liver in the presence of the enzyme liver alcohol dehydrogenase:

CH3CH2OH(aq) + NAD (aq) + H2O(l) → CH3CHO(aq) + NADH(aq) + H3O+(aq) +

is effectively zeroth order overall as the ethanol is in excess and the concentration of the NAD+ is maintained at a constant level by normal metabolic processes. Calculate the rate constant for the conversion of ethanol to ethanal in the liver if the concentration of ethanol in body fluid drops by 50 per cent from 1.5 g dm−3, a level that results in lack of coordination and slurring of speech, in 49 min at body temperature. Express your answer in units of g dm−3 h−1. 6.14 In a study of the alcohol dehydrogenase catalyzed oxidation of

ethanol, the molar concentration of ethanol decreased in a first-order reaction from 220 mmol dm−3 to 56.0 mmol dm−3 in 1.22 × 104 s. What is the rate constant of the reaction?

6.19 Sucrose is readily hydrolyzed to glucose and fructose in acidic

t/min [sucrose]/(mol dm−3) t/min [sucrose]/(mol dm−3)

0 0.316 110 0.211

14 0.300 140 0.190

39 0.274 170 0.170

60 0.256 210 0.146

80 0.238

Determine the order of the reaction with respect to sucrose and the rate constant of the reaction. 6.20 Iodoacetamide and N-acetylcysteine react with 1:1

stoichiometry. The following data were collected at 298 K for the reaction of 1.00 mmol dm−3 N-acetylcysteine with 1.00 mmol dm−3 iodoacetamide: t/s [N-acetylcysteine]/(mmol dm−3) t/s [N-acetylcysteine]/(mmol dm−3)

10 0.770 60 0.315

20 0.580 100 0.210

40 0.410 150 0.155

(a) Explain why analysis of these data yield the overall order of the reaction and not the order with respect to N-acetylcysteine (or iodoacetamide). (b) Plot the data in an appropriate fashion to determine the overall order of the reaction. (c) From the graph, determine the rate constant. 6.21 The following data were collected at 298 K for the reaction of 1.00 mmol dm−3 N-acetylcysteine with 2.00 mmol dm−3

EXERCISES

241

iodoacetamide under conditions that are different from those in Exercise 6.20:

of phenobarbital must be re-injected to restore the original level of anesthetic in the dog?

t/s 5 10 25 35 50 60 [N-acetylcysteine]/(mmol dm−3) 0.74 0.58 0.33 0.21 0.12 0.09

6.31 The second-order rate constant for the reaction CH3COOC2H5(aq) + OH−(aq) → CH3CO−2 (aq) + CH3CH2OH(aq) is 0.11 dm3 mol−1 s−1. What is the concentration of ester after (a) 15 s and (b) 15 min when ethyl acetate is added to sodium hydroxide so that the initial concentrations are [NaOH] = 0.055 mol dm−3 and [CH3COOC2H5] = 0.150 mol dm−3?

(a) Use these data and your result from Exercise 6.20a to determine the order of the reaction with respect to each reactant. (b) Determine the rate constant. 6.22 The composition of a liquid phase reaction 2 A → B was followed spectrophotometrically with the following results:

t/min [B]/(mol dm−3)

0 0

10 0.089

20 0.153

30 0.200

40 0.230

∞ 0.312

Determine the order of the reaction with respect to sucrose and the rate constant of the reaction. 6.23 Establish the integrated form of a third-order rate law of the

form v = kr[A]3. What would it be appropriate to plot to confirm that a reaction is third order? 6.24 Derive an integrated expression for a second-order rate law

v = kr[A][B] for a reaction of stoichiometry 2 A + 3 B → P. 6.25 Derive the integrated form of a third-order rate law v = kr[A]2[B]

in which the stoichiometry is 2 A + B → P and the reactants are initially present in (a) their stoichiometric proportions and (b) with B present initially in twice the amount.

6.32 A reaction 2 A → P has a second-order rate law with

kr = 1.24 cm3 mol−1 s−1. Calculate the time required for the concentration of A to change from 0.260 mol dm−3 to 0.026 mol dm−3.

6.33 Show that the ratio t1/2/t3/4, where t1/2 is the half-life and t3/4 is the time for the concentration of A to decrease to 34 of its initial value (implying that t3/4 < t1/2), can be written as a function of n alone and can therefore be used as a rapid assessment of the order of a reaction. 6.34 (a) Show that, for a reaction that is n-order in A, t1/2 is given by

t1/2 =

2n−1 − 1 (n − 1)kr[A]n−1 o

(b) Deduce an expression for the time it takes for the concentration of a substance to fall to one-third the initial value in an nth-order reaction. 6.35 A rate constant is 1.78 × 104 dm3 mol−1 s−1 at 19°C and

1.38 × 10−3 dm3 mol−1 s−1 at 37°C. Evaluate the Arrhenius parameters of the reaction.

6.26 The half-life of pyruvic acid in the presence of an aminotransferase enzyme (which converts it to alanine) was found to be 221 s. How long will it take for the concentration of pyruvic acid to fall to 641 of its initial value in this first-order reaction?

6.36 The activation energy for the denaturation of the O2-binding protein hemocyanin is 408 kJ mol−1. At what temperature will the rate be 10 per cent greater than its rate at 25°C?

6.27 Radioactive decay of unstable atomic nuclei is a first-order

6.37 Which reaction responds more strongly to changes of

process. The half-life for the (first-order) radioactive decay of 14C is 5730 a (1 a is the SI unit 1 annum, for 1 year; the nuclide emits b particles, high-energy electrons, with an energy of 0.16 MeV). An archaeological sample contained wood that had only 69 per cent of the 14C found in living trees. What is its age?

temperature, one with an activation energy of 52 kJ mol−1 or one with an activation energy of 25 kJ mol−1?

6.28 One of the hazards of nuclear explosions is the generation of 90

6.38 The rate constant of a reaction increases by a factor of 1.23 when the temperature is increased from 20°C to 27°C. What is the activation energy of the reaction? 6.39 Make an appropriate Arrhenius plot of the following data for

Sr and its subsequent incorporation in place of calcium in bones. This nuclide emits b particles of energy 0.55 MeV and has a half-life of 28.1 a (1 a is the SI unit 1 annum, for 1 year). Suppose 1.00 mg was absorbed by a newborn child. How much will remain after (a) 19 a and (b) 75 a if none is lost metabolically?

T/K 289.0 kr /(106 dm3 mol−1 s−1) 1.04

6.29 The estimated half-life for P–O bonds is 1.3 × 105 a (1 a is the SI

6.40 Food rots about 40 times more rapidly at 25°C than when it is

unit 1 annum, for 1 year). Approximately 109 such bonds are present in a strand of DNA. How long (in terms of its half-life) would a single strand of DNA survive with no cleavage in the absence of repair enzymes?

stored at 4°C. Estimate the overall activation energy for the processes responsible for its decomposition.

6.30 To prepare a dog for surgery, about 30 mg (kg body mass)−1 of phenobarbital must be administered intravenously. The anesthetic is metabolized with first-order kinetics and a half-life of 4.5 h. After about 2 h, the drug begins to lose its effect in a 15-kg dog. What mass

the binding of an inhibitor to the enzyme carbonic anhydrase and calculate the activation energy for the reaction. 293.5 1.34

298.1 303.2 1.53 1.89

308.0 2.29

313.5 2.84

6.41 The enzyme urease catalyzes the reaction in which urea is hydrolyzed to ammonia and carbon dioxide. The half-life of urea in the pseudo-first-order reaction for a certain amount of urease doubles when the temperature is lowered from 20°C to 10°C and the equilibrium constant for binding of urea to the enzyme is largely unchanged. What is the activation energy of the reaction?

242

6 THE RATES OF REACTIONS

Project 6.42* Prebiotic reactions are reactions that might have occurred

under the conditions prevalent on the Earth before the first living creatures emerged and can lead to analogs of molecules necessary for life as we now know it. To qualify, a reaction must proceed with a favorable rate and have a reasonable value for the equilibrium constant. An example of a prebiotic reaction is the formation of 5-hydroxymethyluracil (HMU) from uracil and formaldehyde (HCHO). Amino acid analogs can be formed from HMU under prebiotic conditions by reaction with various nucleophiles, such as H2S, HCN, indole, and imidazole. For the synthesis of HMU at pH = 7, the temperature dependence of the rate constant is given by * Adapted from an exercise provided by Charles Trapp, Carmen Giunta, and Marshall Cady.

log kr /(dm3 mol−1 s−1) = 11.75 − 5488/(T/K) and the temperature dependence of the equilibrium constant is given by log K = −1.36 + 1794/(T/K) (a) Calculate the rate constants and equilibrium constants over a range of temperatures corresponding to possible prebiotic conditions, such as 0–50°C, and plot them against temperature. (b) Calculate the activation energy and the standard reaction Gibbs energy and enthalpy at 25°C. (c) Prebiotic conditions are not likely to be standard conditions. Speculate about how the actual values of the reaction Gibbs energy and enthalpy might differ from the standard values. Do you expect that the reaction would still be favorable?

7

Accounting for the rate laws Even quite simple rate laws can give rise to complicated behavior. The observation that the heart maintains a steady pulse throughout a lifetime, but may break into fibrillation during a heart attack, is one sign of that complexity. On a less personal scale, intermediates come and go in the course of reactions, and all reactions approach equilibrium with the forward and reverse reactions proceeding at the same rate. However, the complexity of the behavior of reaction rates means that the study of reaction rates can give deep insight into the way that reactions actually take place. As we remarked in Chapter 6, rate laws are a window on to the mechanism, the sequence of elementary molecular events that leads from the reactants to the products, of the reactions they summarize. In this chapter, we see how to interpret an observed rate law in terms of a proposed mechanism in preparation for dealing with biological systems in Chapter 8.

Reaction mechanisms So far, we have considered very simple rate laws, in which reactants are consumed or products formed. However, all reactions actually proceed toward a state of equilibrium in which the reverse reaction becomes increasingly important. Moreover, many reactions—particularly those in organisms—proceed to products through a series of intermediates. In organisms (and in chemical plants), one of these intermediates may be of crucial importance and the ultimate products may represent waste.

Reaction mechanisms 243 7.1 The approach to

equilibrium In the laboratory 7.1

Relaxation techniques in biochemistry

Many biochemical reactions take place in steps that reach equilibrium quickly, and to understand their role we need to understand the relation between their kinetic behavior and their equilibrium composition.

All forward reactions are accompanied by their reverse reactions. At the start of a reaction, when little or no product is present, the rate of the reverse reaction is negligible. However, as the concentration of products increases, the rate at which they decompose into reactants becomes greater. At equilibrium, the reverse rate matches the forward rate and the reactants and products are present in abundances given by the equilibrium constant for the reaction. (a) The relation between equilibrium constants and rate constants

We can analyze the relation between rates and equilibrium by thinking of a very simple reaction of the form

245

7.2 Elementary

reactions

247

7.3 Consecutive

reactions

249

Case study 7.1

Mechanisms of protein folding and unfolding

254

7.4 Diffusion control

256

7.5 Kinetic and

thermodynamic control

258

Reaction dynamics

259

7.6 Collision theory

259

7.7 Transition state

theory

7.1 The approach to equilibrium

243

261

In the laboratory 7.2

Time-resolved spectroscopy for kinetics 263 7.8 The kinetic salt

effect

264

Checklist of key concepts 267 Checklist of key equations 267 Further information 7.1 Collisions in the gas phase 267 Discussion questions

270

Exercises

270

Projects

272

244

7 ACCOUNTING FOR THE RATE LAWS Forward: A → B

Rate of formation of B = kr[A]

Reverse: B → A

Rate of decomposition of B = k r′[B]

For instance, we could envisage this scheme as the interconversion of coiled (A) and uncoiled (B) DNA molecules. The net rate of formation of B, the difference of its rates of formation and decomposition, is Net rate of formation of B =

d[B] = kr[A] − k ′[B] r dt

When the reaction has reached equilibrium, the concentrations of A and B are [A]eq and [B]eq and there is no net formation of either substance. It follows that d[B]/dt = 0 and hence that kr[A]eq = k ′[B] r eq. Therefore, the equilibrium constant for the reaction is related to the forward and reverse rate constants by The reaction profile for an exothermic reaction. The activation energy is greater for the reverse reaction than for the forward reaction, so the rate of the forward reaction increases less sharply with temperature. As a result, the equilibrium constant shifts in favor of the reactants as the temperature is raised.

Fig. 7.1

Mathematical toolkit 7.1

K=

[B]eq kr = [A]eq k ′r

(7.1)

If the forward rate constant is much larger than the reverse rate constant, then K >> 1. If the opposite is true, then K > 1 is now kr c 3 >> k ′,r and the units of each term match. A brief illustration

The rates of the forward and reverse reactions for the dimerization of proflavin (1), an antibacterial agent that inhibits the biosynthesis of DNA by intercalating between adjacent base pairs, were found to be 8.1 × 108 dm3 mol−1 s−1 (second order) and 2.0 × 106 s−1 (first order), respectively. The equilibrium constant for the dimerization is therefore K=

(8.1 × 108 dm3 mol−1 s−1) × (1 mol dm−3) = 4.0 × 102 2.0 × 106 s−1

7.3 Consecutive reactions In general, biological processes have complex mechanisms and we need to know how to build an overall rate law from the rate law of each step of a proposed mechanism.

A reactant commonly produces an intermediate, a species that does not appear in the overall chemical equation for the reaction but which has been invoked in the mechanism. Biochemical processes are often elaborate versions of this simple model. For instance, the restriction enzyme EcoRI catalyzes the cleavage of DNA at a specific sequence of nucleotides (at GAATTC, making the cut between G and A on both strands). The reaction sequence it brings about is supercoiled DNA → open-circle DNA → linear DNA We can discover the characteristics of this type of reaction by setting up the rate laws for the net rate of change of the concentration of each substance. (a) The variation of concentration with time

To illustrate the kinds of considerations involved in dealing with a mechanism, let’s suppose that a reaction takes place in two first-order steps, in one of which the intermediate I (the open-circle DNA, for instance) is formed from the

249

250

7 ACCOUNTING FOR THE RATE LAWS

reactant A (the supercoiled DNA) in a first-order reaction, and then I decays in a first-order reaction to form the product P (the linear DNA): A→I

Rate of formation of I = ka[A]

I→P

Rate of formation of P = kb[I]

For simplicity, we are ignoring the reverse reactions, which is permissible if they are slow. The first of these rate laws implies that A decays with a first-order rate law and therefore that [A] = [A]0e−k t

(7.10)

a

The net rate of formation of I is the difference between its rate of formation and its rate of consumption, so we can write Net rate of formation of I =

d[I] = ka[A] − kb[I] dt

(7.11)

with [A] given by eqn 7.10. This equation is more difficult to solve, but has the form of the second example in Mathematical toolkit 7.1 and the solution is [I] =

ka (e−k t − e−k t )[A]0 kb − ka a

(7.12)

b

Show that eqn 7.12 follows from eqn 7.11 by (i) recasting eqn 7.11 so that it resembles the standard form in the second example of Mathematical toolkit 7.1 and (ii) showing that the solution to the standard form can be rearranged into eqn 7.12 with the following substitutions: x = t, f(x) = ka[A] (with [A] given by eqn 7.10), F(x) = kbt, a = kb, b = ka[A]0, b′ = ka, and c = ka[A]0/(ka − kb). Self-test 7.1

Finally, because [A] + [I] + [P] = [A]0 at all stages of the reaction, the concentration of P is A k e−k t − kbe−k t D [P] = 1 + a [A]0 C kb − ka F b

a

(7.13)

These solutions are illustrated in Fig. 7.6. We see that the intermediate grows in concentration initially, then decays as A is exhausted. Meanwhile, the concentration of P rises smoothly to its final value. As we see in the following Justification, the intermediate reaches its maximum concentration at t= The concentrations of the substances involved in a consecutive reaction of the form A → I → P, where I is an intermediate and P a product. We have used k1 = 5k2. Note how at each time the sum of the three concentrations is a constant.

1 k ln a ka − kb kb

Time of maximum concentration

(7.14)

Fig. 7.6

A brief illustration

Consider a manufacturing process of a pharmaceutical in which ka = 0.120 h−1 and kb = 0.012 h−1. It follows that the intermediate is at a maximum at t = 21 h after the start of the process. This is the optimum time for a manufacturer trying to make the intermediate in a batch process to extract it.

7.3 CONSECUTIVE REACTIONS

251

Justification 7.3 The time of maximum concentration

To find the time corresponding to the maximum concentration of intermediate, we differentiate eqn 7.12 and look for the time at which d[I]/dt = 0. First we obtain d[I] ka = (−kae−k t + kbe−k t)[A]0 = 0 dt kb − ka a

b

This equation is satisfied if kae−k t = kbe−k t a

b

Because eatebt = e(a+b)t, this relation can be written as ka = e(k −k )t kb a

b

Taking logarithms of both sides leads to eqn 7.14.

(b) The rate-determining step

We now suppose that the second step in the reaction we are considering is very fast, so that whenever an I molecule is formed, it decays rapidly into P. We can use the condition k b >> k a to write e−k t > k a′[I]. For the equilibrium between the intermediate and the reactants, we write (see Section 7.2) K=

[I]c 3 [A][B]

K=

k ac 3 k a′

(7.21)

A brief comment

For the integration we have used the standard result

1 e−kxdx = − e−kx + constant k

254

7 ACCOUNTING FOR THE RATE LAWS

In writing these equations, we are presuming that the rate of reaction of I to form P is too slow to affect the maintenance of the pre-equilibrium (see the example below). The rate of formation of P may now be written d[P] = k b[I] = (k b K/c 3)[A][B] dt

(7.22)

This rate law has the form of a second-order rate law with a composite rate constant: d[P] = kr[A][B] dt

For a reaction with a pre-equilibrium, there are three activation energies to take into account, two referring to the reversible steps of the preequilibrium and one for the final step. The relative magnitudes of the activation energies determine whether the overall activation energy is (a) positive or (b) negative.

Fig. 7.10

kr =

kbK kakb = c3 k a′

(7.23)

One feature to note is that although each of the rate constants in eqn 7.23 increases with temperature, that might not be true of kr itself. Thus, if the rate constant k a′ increases more rapidly than the product kakb increases, then kr will decrease with increasing temperature and the reaction will go more slowly as the temperature is raised. Mathematically, we would say that the composite reaction had a ‘negative activation energy’. For example, suppose that each rate constant in eqn 7.23 exhibits an Arrhenius temperature dependence. It follows from the Arrhenius equation (eqn 6.19, kr = Ae−E /RT ) that a

kr = =

(Aae−E

)(Abe−E A a′e−E′ /RT

a,a

/RT

a,b

/RT

a,a

Aa Ab −(E e Aa′

a,a

)

=

AaAb e−E /RTe−E Aa′ e−E′ /RT a,a

a,b

/RT

a,a

+Ea,b−E′a,a)/RT

where we have used the relations: e x+y = e xe y and e x−y = e x/e y. The effective activation energy of the reaction is therefore Ea = Ea,a + Ea,b − E′a,a

(7.24)

This activation energy is positive if Ea,a + Ea,b > E′a,a (Fig. 7.10a) but negative if E′a,a > Ea,a + Ea,b (Fig. 7.10b). An important consequence of this discussion is that we have to be very cautious about making predictions about the effect of temperature on reactions that are the outcome of several steps. Enzyme-catalyzed reactions may also exhibit strongly non-Arrhenius temperature dependence if the enzyme denatures at high temperatures and ceases to function.

Case study 7.1

Mechanisms of protein folding and unfolding

Much of the kinetic work on the mechanism of unfolding of a helix into a random coil has been conducted on small synthetic polypeptides rich in alanine, an amino acid known to stabilize helical structures. Experimental and theoretical results suggest that the mechanism of unfolding consists of at least two steps: a very fast step, in which amino acids at either end of a helical segment undergo transitions to coil regions, and a slower rate-determining step, which corresponds to the cooperative melting of the rest of the chain and loss of helical content. Using h and c to denote an amino acid residue belonging to a helical and coil region, respectively, the mechanism may be summarized as follows: hhhh . . . 7 chhh

very fast

chhh . . . → cccc

rate-determining step

7.3 CONSECUTIVE REACTIONS

The rate-determining step is thought to account for the relaxation time of 160 ns measured with a laser-induced temperature jump from 280 K to 300 K in an alanine-rich polypeptide containing 21 residues. It is thought that the limitation on the rate of the helix–coil transition in this peptide arises from an activation energy barrier of 1.7 kJ mol−1 associated with initial events of the form . . . hhhh . . . → . . . hhch . . . in the middle of the chain. Therefore, initiation is not only thermodynamically unfavorable but also kinetically slow. Theoretical models also suggest that a hhhh . . . → chhh . . . transition at either end of a helical segment has a significantly lower activation energy on account of the converting residue not being flanked by h regions. The kinetics of unfolding have also been measured in naturally occurring proteins. In the engrailed homeodomain (En-HD) protein, which contains three short helical segments (Fig. 7.11), unfolding occurs with a half-life of about 630 ms at 298 K. It is difficult to interpret these results because we do not yet know how the amino acid sequence or interactions between helices in a folded protein affect the helix–coil relaxation time. As remarked in the Prologue, a protein does not fold into its active conformation by sampling every possible three-dimensional arrangement of the chain, as the process would take far too long—up to 1021 years for a protein with 100 amino acids. Moreover, folding times have been measured in synthetic peptides and naturally occurring proteins and have been found to be very fast. For example, the En-HD protein folds with a half-life of 18 ms at 298 K. In fact, Nature’s search for the active conformation of a large polypeptide appears to be highly streamlined, and the identification of specific mechanisms of protein folding is a major focus of current research in biochemistry. Although it is unlikely that a single model can describe the folding of every protein, progress has been made in the identification of some general mechanistic features. Two models have received attention. In the framework model, regions with well-defined and stable secondary structure form independently and then coalesce to yield the correct tertiary structure. The En-HD protein and other proteins that are predominantly helical fold according to the framework model. In the nucleation–condensation model, rather loose and unstable helices and sheets are thought to form early in the folding process. However, the molecule can be stabilized by interactions that also give rise to some degree of tertiary structure. That is, formation of secondary structure is fostered by the formation of tertiary structure and vice versa. It is easy to imagine that some regions, called ‘nuclei’, of the loosely packed protein resemble the active conformation of the protein rather closely, whereas other regions do not. Far away from the nuclei, similarities to the active conformation are thought to be less prominent, but these regions eventually coalesce, or ‘condense,’ around nuclei to give the properly folded protein. Proteins containing mostly a-helices, mostly b-sheets, or a mixture of the two have been observed to fold in a manner consistent with the nucleation condensation model. A key feature of the framework and nucleation–condensation models is the formation of secondary structure—which might or might not be coupled to the formation of tertiary structure—early in the folding process. It follows that a full description of the mechanism of protein folding also requires an understanding of the rules that stabilize molecular interactions in polypeptides. We consider these rules in Chapter 11.

255

The engrailed homeodomain (En-HD) protein contains three short helical segments.

Fig. 7.11

256

7 ACCOUNTING FOR THE RATE LAWS

7.4 Diffusion control Most biochemical processes require that two or more molecules encounter each other as they travel through the aqueous environment of the cell, so one contribution to the overall rate of enzyme-catalyzed reactions that we need to analyze is the rate at which species diffuse through a solution.

The concept of the rate-determining step plays an important role for reactions in solution, where it leads to the distinction between ‘diffusion control’ and ‘activation control.’ To develop this point, let’s suppose that a reaction between two solute molecules A and B occurs by the following mechanism. First, we assume that A and B drift into each other’s vicinity by diffusion,1 the process by which the molecules of different substances mingle with each other, and form an encounter pair, AB: A + B → AB

Rate of formation of AB = kd[A][B]

The subscript d reminds us that this process is diffusional. The encounter pair persists for some time as a result of the cage effect, the trapping of A and B near each other by their inability to escape rapidly through the surrounding solvent molecules. However, the encounter pair can break up when A and B have the opportunity to diffuse apart, and so we must allow for the following process: AB → A + B

Rate of loss of AB = k d′[AB]

We have supposed that this process is first order in AB. Competing with this process is the reaction between A and B while they exist as an encounter pair. This process depends on their ability to acquire sufficient energy to react. That energy might come from the jostling of the thermal motion of the solvent molecules. We shall assume that the reaction of the encounter pair is first order in AB, but if the solvent molecules are involved, it is more accurate to regard it as pseudo first order with the solvent molecules in great and constant excess. In any event, we can suppose that the reaction is AB → products

Rate of reactive loss of AB = ka[AB]

The subscript a on k reminds us that this process is activated in the sense that it depends on the acquisition by AB of at least a minimum energy. Now we use the steady-state approximation to set up the rate law for the formation of products and deduce in the following Justification that the rate v = d[P]/dt and rate constant kr of formation of products are given by v = kr[A][B]

kr =

kakd ka + kd′

Rate constant in the presence of diffusion and activation

Justification 7.4 Rate in the presence of diffusion and activation

The net rate of formation of AB is d[AB] = kd[A][B] − kd′[AB] − ka[AB] dt In a steady state, this rate is zero, so we can write kd[A][B] − kd′[AB] − ka[AB] = 0 1

Diffusion is treated in more detail in Chapter 8.

(7.25)

7.4 DIFFUSION CONTROL

which we can rearrange to find [AB]: [AB] =

kd[A][B] ka + kd′

The rate of formation of products (which is the same as the rate of reactive loss of AB) is therefore v = ka[AB] =

kakd[A][B] ka + kd′

which is eqn 7.25.

Now we distinguish two limits. Suppose the rate of reaction is much faster than the rate at which the encounter pair breaks up. In this case, ka >> kd′ and we can neglect kd′ in the denominator of the expression for kr in eqn 7.25. The ka in the numerator and denominator then cancel, and we are left with v = kd[A][B]

Diffusion-controlled limit

(7.26a)

In this diffusion-controlled limit the rate of the reaction is controlled by the rate at which the reactants diffuse together (as expressed by kd), for the reaction once they have encountered is so fast that they will certainly go on to form products rather than diffuse apart before reacting. Alternatively, we may suppose that the rate at which the encounter pair accumulated enough energy to react is so low that it is highly likely that the pair will break up. In this case, we can set ka 1 if specific molecular interactions during collisions (for example, Coulomb interactions between charged species) effectively extend the cross-section for the reactive encounter beyond the value expected from simple mechanical contact between reactants. 7.7 Transition state theory The concepts we introduce here form the basis of a theory that explains the rates of biochemical reactions in fluid environments.

261

Table 7.1 Collision cross-sections of atoms and molecules

Species

s/nm2*

Argon, Ar

0.36

Benzene, C6H6

0.88

Carbon dioxide, CO2

0.52

Chlorine, Cl2

0.93

Ethene, C2H4

0.64

Helium, He

0.21

Hydrogen, H2

0.27

Methane, CH4

0.46

Nitrogen, N2

0.43

Oxygen, O2

0.40

Sulfur dioxide, SO2

0.58

*1 nm2 = 10 −18 m2.

Although collision theory is a useful starting point for the discussion of reactions in the atmosphere, it has little relevance to the reactions that interest biologists the most: those taking place in the aqueous environment of a cell. A more sophisticated theory, ‘transition state theory’ (or activated complex theory), builds on collision theory but is applicable to a wider range of reaction environments and introduces a more sophisticated interpretation of the empirical Arrhenius parameters A and Ea. (a) Formulation of the theory

In the transition state theory of reactions it is supposed that as two reactants approach, their potential energy rises and reaches a maximum, as illustrated by the reaction profile in Fig. 7.1. This maximum corresponds to the formation of an activated complex, a cluster of atoms that is poised to pass on to products or to collapse back into the reactants from which it was formed (Fig. 7.16). The concept of an activated complex is applicable to reactions in solutions as well as to the gas phase because we can think of the activated complex as perhaps involving any solvent molecules that may be present. To describe the essential features of transition state theory, we follow the progress of a bimolecular reaction, possibly occurring in solution. Initially only the reactants A and B are present. As the reaction event proceeds, A and B come into contact, distort, and begin to exchange or discard atoms. The potential energy rises to a maximum, and the cluster of atoms that corresponds to the region close to the maximum is the activated complex. The potential energy falls as the atoms rearrange in the cluster and reaches a value characteristic of the products. The climax of the reaction is at the peak of the potential energy. Here two reactant molecules have come to such a degree of closeness and distortion that a small further distortion will send them in the direction of products. This crucial configuration is called the transition state of the reaction. Although some molecules entering the transition state might revert to reactants, if they pass through this configuration it is probable that products will emerge from the encounter. The reaction coordinate is an indication of the stage reached in this process. On the left we have undistorted, widely separated reactants. On the right are the products. Somewhere in the middle is the stage of the reaction corresponding

Energy is not the only criterion of a successful reactive encounter, for relative orientation can also play a role. (a) In this collision, the reactants approach in an inappropriate relative orientation and no reaction occurs even though their energy is sufficient. (b) In this encounter, both the energy and the orientation are suitable for reaction.

Fig. 7.15

262

7 ACCOUNTING FOR THE RATE LAWS

to the formation of the activated complex. The principal goal of transition state theory is to write an expression for the rate constant by tracking the history of the activated complex from its formation by encounters between the reactants to its decay into product. Here we outline the steps involved in the calculation, with an eye toward gaining insight into the molecular events that affect the rate constant. The activated complex C‡ is formed from the reactants A and B and it is supposed—without much justification—that there is an equilibrium between the concentrations of A, B, and C‡: A + B 7 C‡

In the transition state theory of chemical reactions, two reactants encounter each other (either in a gas-phase collision or as a result of diffusing together through a solvent) and, if they have sufficient energy, form an activated complex. The activated complex is depicted here by a relatively loose cluster of atoms that may undergo rearrangement into products. In an actual reaction, only some atoms— those at the actual reaction site—might be significantly loosened in the complex; the bonding of the others remaining almost unchanged. This would be the case for CH3 groups attached to a carbon atom that was undergoing substitution. Fig. 7.16

K‡ =

[C‡]c 3 [A][B]

At the transition state, motion along the reaction coordinate corresponds to some complicated collective vibration-like motion of all the atoms in the complex (and the motion of the solvent molecules if they are involved too). However, it is possible that not every motion along the reaction coordinate takes the complex through the transition state and to the product P. By taking into account the equilibrium between A, B, and C‡ and the rate of successful passage of C‡ through the transition state, it is possible to derive the Eyring equation for the rate constant kr:4 kr = k ×

kT K ‡ × h c3

Eyring equation

(7.32)

where k = R/NA = 1.381 × 10−23 J K−1 is Boltzmann’s constant and h = 6.626 × 10−34 J s is Planck’s constant (Fundamentals F3). The factor k (kappa) is the transmission coefficient, which takes into account the fact that the activated complex does not always pass through to the transition state. In the absence of information to the contrary, k is assumed to be about 1. The term kT/h in eqn 7.32 has the dimensions of a frequency, as kT is an energy and division by Planck’s constant turns an energy into a frequency (with kT in joules, kT/h has the units s−1). It arises from a consideration of the motions of atoms that lead to the decay of C ‡ into products, as specific bonds are broken and formed. Calculation of the equilibrium constant K ‡ is very difficult, except in certain simple model cases. For example, if we suppose that the reactants are two structureless atoms and that the activated complex is a weakly bound diatomic molecule of bond length R, then kr turns out to be the same as for collision theory, provided we interpret the collision cross-section in eqn 7.31 as pR2. (b) Thermodynamic parameterization

It is more useful, especially for biological reactions in aqueous environments, to express the Eyring equation in terms of thermodynamic parameters and to discuss reactions in terms of their empirical values. Thus, we saw in Section 4.3 that an equilibrium constant may be expressed in terms of the standard reaction Gibbs energy (−RT ln K = DrG 3). We do the same here, and express K ‡ in terms of the activation Gibbs energy, D‡G: D‡G = −RT ln K ‡

and

K ‡ = e−D G/RT ‡

Therefore, by writing D‡G = D‡H − TD‡S 4

(7.33)

In some expositions, you will see Boltzmann’s constant denoted kB to emphasize its significance.

7.7 TRANSITION STATE THEORY

263

where D‡H and D‡S are the enthalpy of activation and the entropy of activation, respectively, we conclude that (with k = 1) kr =

kT −(D H−TD S)/RT A kT D S/R D −D H/RT e = e e C h F h ‡

The Eyring equation in terms of thermodynamic parameters

(7.34)

This expression has the form of the Arrhenius expression, eqn 7.29, if we identify the D‡H with the activation energy and the term in parentheses with the preexponential factor. The advantage of transition state theory over collision theory is that it is applicable to reactions in solution as well as in the gas phase. It also gives some clue to the calculation of the steric factor P, for the orientation requirements are carried in the entropy of activation. Thus, if there are strict orientation requirements (for example, in the approach of a substrate molecule to an enzyme), then the entropy of activation will be strongly negative (representing a decrease in disorder when the activated complex forms), and the pre-exponential factor will be small. In practice, it is occasionally possible to estimate the sign and magnitude of the entropy of activation and hence to estimate the rate constant. The general importance of transition state theory is that it shows that even a complex series of events—not only a collisional encounter in the gas phase—displays Arrheniuslike behavior and that the concept of activation energy is widely applicable.

In a certain reaction in water, it is proposed that two ions of opposite charge come together to form an electrically neutral activated complex. Is the contribution of the solvent to the entropy of activation likely to be positive or negative?

Self-test 7.5

Answer: Positive, as H2O is less organized around the neutral species

In the laboratory 7.2

Time-resolved spectroscopy for kinetics

The ability of lasers to produce pulses as brief as 1 fs (10−15 s) is particularly useful in chemistry when we want to monitor very fast processes in time. In time-resolved spectroscopy, a form of flash photolysis (In the laboratory 6.1), laser pulses are used to obtain the spectra of reactants, intermediates, products, and even activated complexes of reactions. Lasers that produce nanosecond pulses are generally suitable for the observation of reactions with rates controlled by the speed with which reactants can move through a fluid medium. However, pulses in the range 1 fs to 1 ps are needed to study energy transfer, molecular vibrations, and conversion from one mode of motion into another. The arrangement shown in Fig. 7.17 is often used to study ultrafast chemical reactions that can be initiated by light. An intense and short laser pulse, the pump, initiates the process, possibly by elevating a molecule A to a high-energy state, A*, which can either release the additional energy, perhaps as light, or react with another species B to yield a product C: A f f g A* (absorption of energy from light) A* → A (energy release) A* + B → AB → C (reaction) laser pulse

where AB denotes either an intermediate or an activated complex.

A configuration used for time-resolved absorption spectroscopy, in which an intense laser pulse is used to generate a monochromatic pump pulse and a second laser pulse, the probe, arrives at the sample some time later to measure a spectroscopic feature of the reaction mixture.

Fig. 7.17

264

7 ACCOUNTING FOR THE RATE LAWS

The rates of appearance and disappearance of the various species are determined by observing time-dependent changes in the spectrum of the sample during the course of the reaction. This monitoring may be done by passing a second, weaker laser pulse, the probe, through the sample at different times after the laser pulse. For example, the wavelength of the probe pulse may be set at an absorption of an intermediate (to probe its formation and decay) or a product (to probe the rate of its formation). Biological processes that are open to study by time-resolved spectroscopy include the energy-converting processes of photosynthesis (Section 5.11 and Case study 12.3) and the light-induced processes of vision (Case study 12.2). In other experiments, the laser-induced ejection of carbon monoxide from myoglobin and the attachment of O2 to the exposed heme site have been studied to obtain rate constants for the two processes. The technique may also be used to detect and study clusters of atoms that resemble activated complexes. In a typical experiment, energy from a femtosecond pump pulse is used to dissociate a molecule, and then a femtosecond probe pulse is fired at an interval after the pulse. The frequency of the probe pulse is set at an absorption of one of the free fragmentation products, so its absorption is a measure of the abundance of the dissociation product. For example, when ICN is dissociated by the first pulse, the emergence of CN can be monitored by watching the growth of the free CN absorption. In this way it has been found that the CN signal remains zero until the fragments have separated by about 600 pm, which takes about 205 fs. Time-resolved techniques have also been used to examine analogs of the activated complex involved in more complex reactions, such as the Diels–Alder reaction, nucleophilic substitution reactions, and pericyclic addition and cleavage reactions.

7.8 The kinetic salt effect Many biochemical reactions in solution are between ions; to treat them, we need to combine transition state theory and the Debye–Hückel limiting law.

The thermodynamic version of transition state theory simplifies the discussion of reactions in solution, particularly those involving ions. For instance, the kinetic salt effect is the effect on the rate of a reaction of adding an inert salt to the reaction mixture. The physical origin of the effect is the difference in stabilization of the reactant ions and the activated complex by the ionic atmosphere (Section 5.1) formed around each of them by the added ions. Thus, in a reaction in which the activated complex forms in the pre-equilibrium A+ + B− 7 C ‡ both reactants are stabilized by their atmospheres, but the activated complex C‡ is not, less C‡ is present at the (presumed) equilibrium, so the rate of formation of products is decreased. On the other hand, if the reaction is between ions of like charge, as in A+ + B+ 7 C ‡2+ the ionic atmosphere around the doubly charged activated complex has a greater effect than around each singly charged ion, the complex is stabilized more than either ion, so its abundance at equilibrium is increased and the rate of formation of products is increased too. We show in the following Justification that quantitative treatment of the problem leads to the result that

7.8 THE KINETIC SALT EFFECT log kr = log k or + 2AzAzB I 1/2

The kinetic salt effect

(7.35)

where k or is the rate constant in the absence of added salt and A = 0.509 for water at 25°C. The charge numbers of A and B are zA and zB, so the charge number of the activated complex is zA + zB; the zJ are positive for cations and negative for anions. The quantity I is the ionic strength due to the added salt (Section 5.1), and for a 1:1 electrolyte (such as NaCl) is equal to the numerical value of the molality (that is, I = b/b3, with b3 = 1 mol kg−1).

Justification 7.5 The kinetic salt effect

Consider a reaction with the mechanism A + B 7 C ‡ → products, where A and B are charged species, and the rate of formation of products is v = kr[A][B] Our goal is to write an expression for the rate constant kr. We begin by writing the thermodynamic equilibrium constant in terms of activities a and activity coefficients g: K=

aC [C‡]c 3 = Kg , aAaB [A][B] ‡

where

Kg =

gC gAgB ‡

It follows that [A][B] =

K gc 3 ‡ [C ] K

The combination of this expression with v = kr[A][B] gives v = kr

A K gc 3 ‡ D [C ] C K F

If we let k ‡ be the rate constant for formation of products from the activated complex C‡, then we may also write v = k ‡[C‡] It follows that kr =

k ‡K K gc 3

If k or is the rate constant when the activity coefficients are 1 (that is, k or = k‡K/c 3), we can write kr =

k ro Kg

At low concentrations the activity coefficients can be expressed in terms of the ionic strength, I, of the solution by using the Debye–Hückel limiting law (eqn 5.4, log gJ = −Az 2J I 1/2). Then log kr = log k or − A{zA2 + zB2 − (zA + zB)2}I 1/2 = log k or + 2AzAzBI 1/2 as in eqn 7.35.

265

266

7 ACCOUNTING FOR THE RATE LAWS

Equation 7.35 confirms that if the reactants have opposite charges (so zAzB is negative), then the rate decreases as the ionic strength is increased (Fig. 7.18), just as the qualitative description suggested. However, if the charges of the reactant ions have the same sign (and zAzB is positive), then the rate increases when a salt is added. Information of this kind is useful in unraveling the reaction mechanism of reactions in solution and identifying the nature of the activated complex.

Example 7.2

Analyzing the kinetic salt effect

The study of conditions that optimize the association of proteins in solution guides the design of protocols for the formation of large crystals that are amenable to analysis by the X-ray diffraction techniques discussed in Chapter 11. It is important to characterize protein dimerization because the process is considered to be the rate-determining step in the growth of crystals of many proteins. Consider the variation with ionic strength of the rate constant of dimerization in aqueous solution of a cationic protein P: An illustration of the kinetic salt effect. If the reactants have opposite charges, then the rate decreases as the ionic strength, I, is increased. However, if the charges of the reactant ions have the same sign, then the rate increases when a salt is added.

Fig. 7.18

I kr /k or

0.0100 0.0150 0.0200 8.10 13.30 20.50

0.0250 27.80

0.0300 0.0350 38.10 52.00

What can be deduced about the charge of P? Strategy Although the dimer is not an activated complex in the same sense as in transition state theory, eqn 7.35 applies if we assume that the activated complex and the product (the dimer) are similar in the sense that two protein molecules associate to form the activated complex. Thus, the equilibrium constant for the dimerization is related to the rate constants for the formation of the dimer and its decomposition by K = kr /k r′c 3 and K = KcK g . Hence kr = Kc K gk r′c 3 = Kg k or . It then follows, as in Justification 7.5, that

log (kr /k or ) = 1.02z 2I 1/2 Therefore, to infer the protein charge number z from the slope, 1.02z 2, we need to plot log (kr /k or ) against I 1/2. Answer: We draw up the following table: I 1/2 log (kr /k or )

0.100 0.908

0.122 0.141 1.124 1.312

0.158 0.173 1.444 1.581

0.187 1.716

These points are plotted in Fig. 7.19. The slope of the straight line is 9.2, indicating that z 2 = 9. Because the protein is cationic, its charge number is +3. An ion of charge number +1 is known to be involved in the activated complex of a reaction. Deduce the charge number of the other ion from the following data: Self-test 7.6

The plot for the data in Example 7.2. Fig. 7.19

I kr /k or

0.0050 0.010 0.850 0.791

0.015 0.750

0.020 0.717

0.025 0.689

0.030 0.666 Answer: –1

FURTHER INFORMATION

267

Checklist of key concepts 1. In relaxation methods of kinetic analysis, the equilibrium position of a reaction is first shifted suddenly and then allowed to readjust to the equilibrium composition characteristic of the new conditions.

5. In the steady-state approximation, it is assumed that the concentrations of all reaction intermediates remain constant and small throughout the reaction. 6. Provided a reaction has not reached equilibrium, the products of competing reactions are controlled by kinetics.

2. An elementary unimolecular reaction has first-order kinetics; an elementary bimolecular reaction has second-order kinetics.

7. In collision theory, it is supposed that the rate is proportional to the collision frequency, a steric factor, and the fraction of collisions that occur with at least the kinetic energy Ea along their lines of centers.

3. The molecularity of an elementary reaction is the number of molecules coming together to react.

8. In transition state theory, it is supposed that an activated complex is in equilibrium with the reactants and that the rate at which that complex forms products depends on the rate at which it passes through a transition state.

4. The rate-determining step is the slowest step in a reaction mechanism that controls the rate of the overall reaction.

Checklist of key equations Property

Equation

Comment

Equilibrium constant in terms of rate constants

K = kr /kr′

First order in each direction

Relaxation time

x = x0e

Rate constant of a diffusion-controlled reaction

kr = kd = 8RT/3h

Rate constant of an activation-controlled reaction

kr = kakd /kd′

−t/t

1/t = kr + kr′

Temperature jump experiment

Kinetic control

[P2]/[P1] = k2/k1

Eyring equation

kr = k(kT/h)(K ‡/c 3)

Thermodynamic parameterization

kr = (kT/h)e−D G/RT = {(kT/h)eD S/R}e−D H/RT

Eyring equation with k = 1

Kinetic salt effect

log kr = log k + 2AzAzB I

Dilute solutions

o r

1/2

Further information Further information 7.1 Collisions in the gas phase

To gain insight into collisions in the gas phase, first we need to explore a model that can explain the properties of perfect gases. Then we consider the distribution of molecular speeds, a concept we previewed in Fundamentals F.2 and F.3. Finally, we define some of the parameters used in the collision theory of chemical reactions in the gas phase. (a) The kinetic model of gases

The basis for our discussion is the kinetic model of gases (also called the ‘kinetic molecular theory’, KMT, of gases), which makes the following three assumptions:

1. A gas consists of molecules in ceaseless random motion (Fig. F.9). 2. The size of the molecules is negligible in the sense that their diameters are much smaller than the average distance traveled between collisions. 3. The molecules do not interact, except during collisions. The assumption that the molecules do not interact unless they are in contact implies that the potential energy of the molecules (their energy due to their position) is independent of their separation and may be set equal to zero. The total energy of a sample of gas is therefore the sum of the kinetic energies

268

7 ACCOUNTING FOR THE RATE LAWS

(the energy due to motion) of all the molecules present in it. It follows that the faster the molecules travel (and hence the greater their kinetic energy), the greater the total energy of the gas. The kinetic model accounts for the steady pressure exerted by a gas in terms of the collisions the molecules make with the walls of the container. Each collision gives rise to a brief force on the wall, but as billions of collisions take place every second, the walls experience a virtually constant force, and hence the gas exerts a steady pressure. On the basis of this model, the pressure exerted by a gas of molar mass M in a volume V is5 p=

nMc 2 3V

The pressure of a perfect gas according to the kinetic model

(7.36)

where c is the root-mean-square speed (r.m.s. speed) of the molecules. This quantity is defined as the square root of the mean value of the squares of the speeds, v, of the molecules. That is, for a sample consisting of N molecules with speeds v1, v2, . . . , vN, we square each speed, add the squares together, divide by the total number of molecules (to get the mean, denoted by 〈· · ·〉), and finally take the square root of the result: A v 2 + v 2 + · · · + vN2 D 1/2 c = 〈v 〉 = 1 2 C F N 2 1/2

Definition of the r.m.s. speed

(7.37)

The r.m.s. speed might at first encounter seem to be a rather peculiar measure of the mean speeds of the molecules, but its significance becomes clear when we make use of the fact that the kinetic energy of a molecule of mass m traveling at a speed v is E k = 12 mv 2, which implies that the mean kinetic energy, 〈Ek 〉, is the average of this quantity, or 12 mc 2. It follows that c=

A 2〈Ek 〉 D 1/2 C m F

The r.m.s. speed in terms of the mean kinetic energy

(7.38)

Therefore, wherever c appears, we can think of it as a measure of the mean kinetic energy of the molecules of the gas. We also note that eqn 7.36 resembles the perfect gas equation of state, pV = nRT, for we can rearrange it into pV = 13 nMc 2 Equating the expression on the right to nRT gives 1 3

nMc 2 = nRT

where the ns now cancel. The great usefulness of this expression is that we can rearrange it into a formula for the r.m.s. speed of the gas molecules at any temperature: c=

A 3RT D 1/2 C M F

The r.m.s. speed in terms of the temperature

(7.39)

A brief illustration

Substitution into eqn 7.39 of the molar mass of O2 (32.0 g mol−1) and a temperature corresponding to 25°C (that is, 298 K) gives an r.m.s. speed for these molecules of 482 m s−1. The same calculation for nitrogen molecules gives 515 m s−1.

5

For a derivation of eqn 7.36, see our Physical chemistry (2010).

The important conclusion to draw from eqn 7.39 is that the r.m.s. speed of molecules in a gas is proportional to the square root of the temperature. Because the mean speed is proportional to the r.m.s. speed, the same is true of the mean speed too. Therefore, doubling the temperature (on the Kelvin scale) increases the mean and the r.m.s. speed of molecules by a factor of 21/2 = 1.414 . . . . (b) The Maxwell distribution of speeds

The mathematical expression that tells us the fraction of molecules that have a particular speed at any instant is called the distribution of molecular speeds. The precise form of the distribution was worked out by James Clerk Maxwell towards the end of the nineteenth century, and his expression is known as the Maxwell distribution of speeds. According to Maxwell, the fraction f of molecules that have a speed in a narrow range between s and s + Ds (for example, between 300 m s−1 and 310 m s−1, corresponding to s = 300 m s−1 and Ds = 10 m s−1) is f = F(s)Ds F(s) = 4p

with

A M D 3/2 2 −Ms /2RT se C 2pRT F 2

Maxwell distribution of speeds

(7.40)

Although eqn 7.40 looks complicated, its features can be picked out quite readily: 1. Because f is proportional to the range of speeds Ds, we see that the fraction in the range Ds increases in proportion to the width of the range. If at a given speed we double the range of interest (but still ensure that it is narrow), then the fraction of molecules in that range doubles too. 2. Equation 7.40 includes a decaying exponential function, the term e−Ms /2RT. Its presence implies that the fraction of molecules with very high speeds will be very small because e−x becomes very small when x 2 is large. 2

2

3. The factor M/2RT multiplying s2 in the exponent is large when the molar mass, M, is large, so the exponential factor goes most rapidly towards zero when M is large. That tells us that heavy molecules are unlikely to be found with very high speeds. 4. The opposite is true when the temperature, T, is high: then the factor M/2RT in the exponent is small, so the exponential factor falls towards zero relatively slowly as s increases. This tells us that at high temperatures, a greater fraction of the molecules can be expected to have high speeds than at low temperatures. 5. A factor s 2 (the term before the e) multiplies the exponential. This factor goes to zero as s goes to zero, so the fraction of molecules with very low speeds will also be very small. The remaining factors (the term in parentheses in eqn 7.40 and the 4p) simply ensure that when we add together the fractions over the entire range of speeds from zero to infinity, then we get 1. These features are summarized in Figs F.10 and F.11.

FURTHER INFORMATION

269

(c) Molecular collisions

The average distance that a molecule travels between collisions is called its mean free path, l (lambda). The mean free path in a liquid is less than the diameter of the molecules because a molecule in a liquid meets a neighbor even if it moves only a fraction of a diameter. However, in gases, the mean free paths of molecules can be several hundred molecular diameters. If we think of a molecule as the size of a tennis ball, then the mean free path in a typical gas would be about the length of a tennis court. The collision frequency, z, is the average rate of collisions made by one molecule. Specifically, z is the average number of collisions one molecule makes in a given time interval divided by the length of the interval. It follows that the inverse of the collision frequency, 1/z, is the time of flight, the average time that a molecule spends in flight between two collisions (for instance, if there are 10 collisions per second, so the collision frequency is 10 s−1, then the average time between collisions is 101 of a second and the time of flight is 101 s). As we shall see, the collision frequency in a typical gas is about 109 s−1 at 1 atm and room temperature, so the time of flight in a gas is typically 1 ns. Because speed is distance traveled divided by the time taken for the journey, the r.m.s. speed c, which we can loosely think of as the average speed, is the average length of the flight of a molecule between collisions (that is, the mean free path, l) divided by the time of flight (1/z). It follows that the mean free path and the collision frequency are related by mean free path l c= = = lz time of flight 1/z

The r.m.s. speed in terms of the mean free path and the collision frequency

(7.41)

To find expressions for l and z, we need a slightly more elaborate version of the kinetic model of gases. The basic kinetic model supposes that the molecules are effectively pointlike; however, to obtain collisions, we need to assume that two ‘points’ score a hit whenever they come within a certain range d of each other, where d can be thought of as the diameter of the molecules (Fig. 7.20). The collision cross-section, s (sigma), the target area presented by one molecule to another, is therefore the area of a circle of radius d, so s = pd 2. When this quantity is built into the kinetic model, we find that kT l = 1/2 2 sp

21/2sp z= c kT

The mean free path and the collision frequency in terms of the collision cross-section

(7.42)

Fig. 7.20 To calculate features of a perfect gas that are related to collisions, a point is regarded as being surrounded by a sphere of diameter d. A molecule will hit another molecule if the center of the former lies within a circle of radius d. The collision cross-section is the target area, pd 2.

We can identify the following features: 1. Because l ∝ 1/p, we see that the mean free path decreases as the pressure increases. This decrease is a result of the increase in the number of molecules present in a given volume as the pressure is increased, so each molecule travels a shorter distance before it collides with a neighbor. For example, the mean free path of an O2 molecule decreases from 73 nm to 36 nm when the pressure is increased from 1.0 bar to 2.0 bar at 25°C. 2. Because l ∝ 1/s, the mean free path is shorter for molecules with large collision cross-sections. For instance, the collision cross-section of a benzene molecule (0.88 nm2) is about four times greater than that of a helium atom (0.21 nm2), and at the same pressure and temperature its mean free path is four times shorter. 3. Because z ∝ p, the collision frequency increases with the pressure of the gas. This dependence follows from the fact that, provided the temperature is the same, the molecule take less time to travel to its neighbor in a denser, higher pressure gas. For example, although the collision frequency for an O2 molecule in oxygen gas at 298.15 K and 1.0 bar is 6.2 × 109 s−1, at 2.0 bar and the same temperature the collision frequency is doubled, to 1.2 × 1010 s−1.

270

7 ACCOUNTING FOR THE RATE LAWS

Discussion questions 7.1 Sketch, without carrying out the calculation, the variation of concentration with time for the approach to equilibrium when both forward and reverse reactions are second order. How does your graph differ from that in Fig. 7.2? 7.2 Write a brief report on a recent research article in which at least one of the following techniques was used to study the kinetics of a biochemical reaction: stopped-flow techniques, time-resolved spectroscopy, chemical quench-flow methods, freeze-quench methods, temperature-jump methods. Your report should be similar in content and extent to one of the Case studies found throughout this text. 7.3 Assess the validity of the following statement: the ratedetermining step is the slowest step in a reaction mechanism.

7.4 Distinguish between a pre-equilibrium approximation and a steady-state approximation. 7.5 Distinguish between a diffusion-controlled reaction and an activation-controlled reaction. 7.6 Distinguish between kinetic and thermodynamic control of a reaction. Suggest criteria for expecting one over the other. 7.7 Describe the formulation of the Eyring equation and interpret its form. 7.8 Is it possible for the activation energy of a reaction to be negative? Explain your conclusion and provide a molecular explanation. 7.9 Discuss the physical origin of the kinetic salt effect.

Exercises 7.10 The equilibrium constant for the attachment of a substrate to the active site of an enzyme was measured as 235. In a separate experiment, the rate constant for the second-order attachment was found to be 7.4 × 107 dm3 mol−1 s−1. What is the rate constant for the loss of the unreacted substrate from the active site? 7.11 Find the solutions of the same rate laws that led to eqn 7.2, but

for some B present initially. Go on to confirm that the solutions you find reduce to those in eqn 7.2 when [B]0 = 0. 7.12 The reaction H2O(l) 7 H+(aq) + OH−(aq) (pKw = 14.01) relaxes

to equilibrium with a relaxation time of 37 ms at 298 K and pH ≈ 7. (a) Given that the forward reaction (with rate constant kr) is first order and the reverse is second order overall (with rate constant k r′), show that 1 = kr + k r′([H+]eq + [OH−]eq) t

(b) Calculate the rate constants for the forward and reverse reactions. 7.13 A protein dimerizes according to the reaction 2A 7 A2 with

forward rate constant kr and reverse rate constant k r′. Show that the relaxation time is t=

1 k r′ + 4kr[A]eq

7.14 Consider the dimerization of a protein, as in Exercise 7.13.

(a) Derive the following expression for the relaxation time in terms of the total concentration of protein, [A]tot = [A] + 2[A2]: 1 = k r′ 2 + 8kr k r′[A]tot t2 (b) Describe the computational procedures that lead to the determination of the rate constants kr and k r′ from measurements of t for different values of [A]tot. 7.15 An understanding of the kinetics of formation of molecular

complexes held together by hydrogen bonds gives insight into the formation of base pairs in nucleic acids. Use the data provided below

and the procedure you outlined in Exercise 7.14 to calculate the rate constants kr and k r′ and the equilibrium constant K for formation of hydrogen-bonded dimers of 2-pyridone (2): [P]/(mol dm−3) t/ns

0.500 2.3

0.352 2.7

0.251 3.3

0.151 4.0

0.101 5.3

7.16 Confirm (by differentiation) that the three expressions in eqns 7.10, 7.12, and 7.13 are correct solutions of the rate laws for consecutive first-order reactions. 7.17 Two radioactive nuclides decay by successive first-order

processes: Xfg Yfg Z 22.5 d

33.0 d

(The times are half-lives in days.) Suppose that Y is an isotope that is required for medical applications. At what stage after X is first formed will Y be most abundant? 7.18 Use mathematical software or an electronic spreadsheet to examine the time dependence of [I] in the reaction mechanism A → I → P (ka, kb) by plotting the expression in eqn 7.12. In the following calculations, use [A]0 = 1 mol dm−3 and a time range of 0 to 5 s. (a) Plot [I] against t for ka = 10 s−1 and kb = 1 s−1. (b) Increase the ratio kb /ka steadily by decreasing the value of ka and examine the plot of [I] against t at each turn. What approximation about d[I]/dt becomes increasingly valid? 7.19 The reaction 2 H2O2(aq) → H2O(l) + O2(g) is catalyzed by Br − ions. If the mechanism is as shown below give the predicted order of the reaction with respect to the various participants.

EXERCISES

H2O2(aq) + Br−(aq) → H2O(l) + BrO−(aq)

(slow)

BrO−(aq) + H2O2(aq) → H2O(l) + O2(g) + Br−(aq)

(fast)

(fast)

A+B→P

(slow)

diffusion controlled. Estimate the rate constant for the reaction at 298 K, given that the viscosity of ethanol is 1.06 × 10−3 kg m−1 s−1. 7.27 Collision theory demands knowing the fraction of molecular

7.20 The reaction mechanism

A2 7 A + A

collisions having at least the kinetic energy Ea along the line of flight. What is this fraction when (a) Ea = 10 kJ mol−1 and (b) Ea = 100 kJ mol−1 at (i) 300 K and (ii) 1000 K?

involves an intermediate A. Deduce the rate law for the formation of P. 7.21 Consider the following mechanism for formation of a double

7.28 Calculate the percentage increase in the fractions in Exercise 7.27

when the temperature is raised by 10 K. 7.29 Calculate the ratio of rates of catalyzed to non-catalyzed

helix from its strands A and B: A + B 7 unstable helix

(fast)

unstable helix → stable double helix

(slow)

Derive the rate equation for the formation of the double helix and express the rate constant of the reaction in terms of the rate constants of the individual steps. 7.22 The following mechanism has been proposed for the

decomposition of ozone in the atmosphere: (1) O3 → O2 + O and its reverse (ka, k a′) (2) O + O3 → O2 + O2 (kb; the reverse reaction is negligibly slow) Use the steady-state approximation, with O treated as the intermediate, to find an expression for the rate of decomposition of O3. Show that if step 2 is slow, then the rate is second order in O3 and −1 order in O2. 7.23 The condensation reaction of acetone, (CH3)2CO (propanone), in aqueous solution is catalyzed by bases, B, which react reversibly with acetone to form the carbanion C3H5O−. The carbanion then reacts with a molecule of acetone to give the product. A simplified version of the mechanism is

reactions at 37°C given that the Gibbs energy of activation for a particular reaction is reduced from 100 kJ mol−1 to 10 kJ mol−1. 7.30 Estimate the pre-exponential factor for the reaction between

molecular hydrogen and ethene at 400°C. Hint: The steric factor is P = 1.7 × 10−6. 7.31 The mechanism of a composite reaction consists of a fast pre-equilibrium step with forward and reverse activation energies of 25 kJ mol−1 and 38 kJ mol−1, respectively, followed by an elementary step of activation energy 10 kJ mol−1. What is the activation energy of the composite reaction? 7.32 Rhodopsin is the protein in the retina that absorbs light, starting

a cascade of chemical events that we call vision (see Case study 12.2 for additional information). Bovine rhodopsin undergoes a transition from one form (metarhodopsin I) to another form (metarhodopsin II) with a half-life of 600 ms at 37°C to 1 s at 0°C. On the other hand, studies of a frog retina show that the same transformation has a half-life that increases by only a factor of 6 over the same temperature range. Suggest an explanation and speculate on the survival advantages that this difference represents for the frog. 7.33 Estimate the activation Gibbs energy for the decomposition

of urea in the reaction (NH2)2CO(aq) + 2 H2O(l) → 2 NH 4+(aq) + CO 2− 3 (aq) for which the pseudo-first-order rate constant is 1.2 × 10−7 s−1 at 60°C and 4.6 × 10−7 s−1 at 70°C.

(1) AH + B → BH+ + A− (2) A− + BH+ → AH + B (3) A− + HA → product where AH stands for acetone and A− its carbanion. Use the steadystate approximation to find the concentration of the carbanion and derive the rate equation for the formation of the product. 7.24 Consider the acid-catalyzed reaction

7.34 Calculate the entropy of activation of the reaction in Exercise

7.33 at the two temperatures. 7.35 Calculate the Gibbs energy, enthalpy, and entropy of activation

(at 300 K) for the binding of an inhibitor to the enzyme carbonic anhydrase using the following data:

HA + H+ 7 HAH+

(fast)

HAH + B → BH + AH

(slow)

T/ K kr /(106 dm3 mol−1 s−1)

289.0 1.04

293.5 1.34

298.1 1.53

Deduce the rate law and show that it can be made independent of the specific term [H+].

T/ K kr /(106 dm3 mol−1 s−1)

303.2 1.89

308.0 2.29

313.5 2.84

7.25 Models of population growth are analogous to chemical

7.36 The reaction A− + H+ → P has a rate constant given by the

+

+

reaction rate equations. In the model due to Malthus (1798) the rate of change of the population N of the planet is assumed to be given by dN/dt = births − deaths. The numbers of births and deaths are proportional to the population, with proportionality constants b and d. Obtain the integrated rate law. How well does it fit the (very approximate) data below on the population of the planet as a function of time? Year 1750 1825 N/109 0.5 1

271

1922 1960 2 3

1974 1987 2000 4 5 6

7.26 The compound a-tocopherol, a form of vitamin E (Atlas M3), is a powerful antioxidant that may help to maintain the integrity of biological membranes. The light-induced reaction between duroquinone and the antioxidant in ethanol is bimolecular and

empirical expression kr = (8.72 × 1012)e(6134 K)/T dm3 mol−1 s−1. Evaluate the energy and entropy of activation at 25°C.

7.37 The conversion of fumarate ion to malate ion is catalyzed by the enzyme fumarase:

fumarate2−(aq) + H2O(l) 7 malate2−(aq) (a) Sketch the reaction profile for this reaction given that (i) the standard enthalpy of formation of the fumarate–fumarase complex from fumarate ion and enzyme is 17.6 kJ mol−1, (ii) the enthalpy of activation of the forward reaction is 41.3 kJ mol−1, (iii) the standard enthalpy of formation of the malate–fumarase complex from malate ion and enzyme is −5.0 kJ mol−1, and (iv) the standard reaction enthalpy is −20.1 kJ mol−1. (b) What is the enthalpy of activation of the reverse reaction?

272

7 ACCOUNTING FOR THE RATE LAWS

7.38 The activation Gibbs energy is composed of two terms: the activation enthalpy and the activation entropy. Differences in the latter can lead to the activation Gibbs energy for a process having the same values despite species inhabiting environments that differ widely in temperature. Show how the data depicted in Fig. 7.21 support this remark. The data relate to the enthalpy and entropy of activation of myosin ATPase in different species of fish living in environments ranging from the Arctic to hot springs. 7.39 At 25°C, kr = 1.55 dm6 mol−2 min−1 at an ionic strength of 0.0241

for a reaction in which the rate-determining step involves the encounter of two singly charged cations. Use the Debye–Hückel limiting law to estimate the rate constant at zero ionic strength.

(right) The correlation of the enthalpy and entropy of activation of the reaction catalyzed by myosin ATPase in a variety of fish species. (Data from I.A. Johnson and G. Goldspink, Nature 257, 620 (1970), recalculated by H. Guttfreund, Kinetics for the life sciences. Cambridge University Press (1995).

Fig. 7.21

Projects 7.40 The absorption and elimination of a drug in the body may be modeled with a mechanism consisting of two consecutive reactions:

A → drug at site of administration

B drug dispersed in blood

C eliminated drug

where the rate constants of absorption (A → B) and elimination are, respectively, ka and kb. (a) Consider a case in which absorption is so fast that it may be regarded as instantaneous, and a dose of A at an initial concentration [A]0 immediately leads to a drug concentration in blood of [B]0. Also, assume that elimination follows first-order kinetics. (i) Show that, after the administration of N equal doses separated by a time interval t, the peak concentration of drug B in the blood, [P]N, rises beyond the value of [B]0 and eventually reaches a constant, maximum peak value given by [P]∞ = [B]0(1 − e−k t)−1 b

[P]N is the (peak) concentration of B immediately after administration of the Nth dose and [P]∞ is the value at very large N. (ii) Write a mathematical expression for the residual concentration of B, [R]N, which we define to be the concentration of drug B immediately before the administration of the (N + 1)th dose. Note that [R]N is always smaller than [P]N because of drug elimination during the period t between drug administrations. Show that [P]∞ − [R]∞ = [B]0. (b) Consider a drug for which kb = 0.0289 h−1. (i) Calculate the t value required to achieve [P]∞ /[B]0 = 10. Prepare a graph that plots both [P]N/[B]0 and [R]N /[B]0 against N. (ii) How many doses must be administered to achieve a [P]N value that is 75 per cent of the maximum value? What time has passed during the administration of these doses? (iii) What actions can be taken to reduce the variation [P]∞ − [R]∞ while maintaining the same value of [P]∞?

(c) Now consider the administration of a single dose [A]0 for which absorption follows first-order kinetics and elimination follows zeroth-order kinetics. Show that with the initial concentration [B]0 = 0, the concentration of drug in the blood is given by [B] = [A]0(1 − e−k t) − kbt a

Plot [B]/[A]0 against t for the case ka = 10 h−1, kb = 4.0 × 10−3 mmol dm−3 h−1, and [A]0 = 0.1 mmol dm−3. Comment on the shape of the curve. (d) Using the model from part (c), set d[B]/dt = 0 and show that the maximum value of [B] occurs at the time tmax =

1 A k [A] D ln a 0 C kb F ka

Also, show that the maximum concentration of drug in blood is given by [B]max = [A]0 − kb /ka − kbtmax. 7.41 Consider a mechanism for the helix–coil transition in which

nucleation occurs in the middle of the chain: hhhh . . . 7 hchh

hchh . . . 7 cccc

We saw in Case study 7.1 that this type of nucleation is relatively slow, so neither step may be rate determining. (a) Set up the rate equations for this alternative mechanism. (b) Apply the steady-state approximation and show that, under these circ*mstances, the mechanism is equivalent to hhhh . . . 7 cccc . . . (c) Use your knowledge of experimental techniques and your results from parts (a) and (b) to support or refute the following statement: It is very difficult to obtain experimental evidence for intermediates in protein folding by performing simple rate measurements and one must resort to special flow, relaxation, or trapping techniques to detect intermediates directly.

8

Complex biochemical processes Biochemical processes use a number of strategies to achieve kinetic control. Chief among them is the use of enzymes to accelerate and regulate the rates of chemical reactions that, although thermodynamically favorable under intracellular conditions, would be too slow to account for the observed rate of growth of organisms and the processes of life in general. With the constant development of powerful experimental techniques, biochemists are beginning to decipher the mechanisms of even the most complex biological processes, such as the transport of nutrients across cell membranes and the transfer of electrons between proteins during glucose metabolism and photosynthesis. In this chapter we describe these processes and develop the physical and chemical concepts that will be used throughout the remainder of the text.

Enzymes Enzymes are hom*ogeneous biological catalysts that work by lowering the activation energy of a reaction pathway or providing a new pathway with a low activation energy. Enzymes are special biological polymers that contain an active site, which is responsible for binding the substrates, the reactants, and processing them into products. As is true of any catalyst, the active site returns to its original state after the products are released. Many enzymes consist primarily of proteins, some featuring organic or inorganic cofactors in their active sites. However, certain ribonucleic acid (RNA) molecules can also be biological catalysts, forming ribozymes. A very important example of a ribozyme is the ribosome, a large assembly of proteins and catalytically active RNA molecules responsible for the synthesis of proteins in the cell. The structure of the active site is specific to the reaction that it catalyzes, with groups in the substrate interacting with groups in the active site through intermolecular interactions, such as hydrogen bonding, electrostatic, or van der Waals interactions (see Chapter 11). Figure 8.1 shows two models that explain the binding of a substrate to the active site of an enzyme. In the lock-and-key model, the active site and substrate have complementary three-dimensional structures and dock perfectly without the need for major atomic rearrangements. Experimental evidence favors the induced fit model, in which binding of the substrate induces a conformational change in the active site. Only after the change does the substrate fit snugly in the active site. Enzyme-catalyzed reactions are prone to inhibition by molecules that interfere with the formation of product. As we remarked in the Prolog, many drugs for the treatment of disease inhibit enzymes of infectious agents, such as bacteria and viruses. Here we focus on the kinetic analysis of enzyme inhibition, and in

Enzymes

273

8.1

The Michaelis–Menten mechanism of enzyme catalysis 274

8.2

The analysis of complex mechanisms 277

8.3

The catalytic efficiency of enzymes 279

8.4

Enzyme inhibition

280

Case study 8.1 The molecular basis of catalysis by hydrolytic enzymes 284

Transport across biological membranes 8.5 8.6 8.7

285

Molecular motion in liquids

285

Molecular motion across membranes

288

The mobility of ions

290

In the laboratory 8.1

Electrophoresis 8.8

Transport across ion channels and ion pumps

Electron transfer in biological systems 8.9

The rates of electron transfer processes

291

294 296

296

8.10 The theory of electron

transfer processes

298

8.11 Experimental tests

of the theory

299

8.12 The Marcus

cross-relation

300

Checklist of key concepts 303 Checklist of key equations 303 Further information 8.1 Fick’s laws of diffusion

304

Discussion questions

305

Exercises

305

Projects

308

274

8 COMPLEX BIOCHEMICAL PROCESSES

Chapters 10 and 11 we shall see how computational methods contribute to the design of efficient inhibitors and potent drugs. 8.1 The Michaelis–Menten mechanism of enzyme catalysis Because enzyme-controlled reactions are so important in biochemistry, we need to build a model of their mechanism. The simplest approach proposed by Michaelis and Menten is our starting point.

Fig. 8.1 Two models that explain the binding of a substrate to the active site of an enzyme. In the lock-and-key model, the active site and substrate have complementary threedimensional structures and dock perfectly without the need for major atomic rearrangements. In the induced fit model, binding of the substrate induces a conformational change in the active site. The substrate fits well in the active site after the conformational change has taken place.

Experimental studies of enzyme kinetics are typically conducted by monitoring the initial rate of product formation in a solution in which the enzyme is present at very low concentration. Indeed, enzymes are such efficient catalysts that significant accelerations may be observed even when their concentrations are more than three orders of magnitude smaller than those of their substrates. The principal features of many enzyme-catalyzed reactions are as follows (Fig. 8.2): 1. For a given initial concentration of substrate, [S]0, the initial rate of product formation is proportional to the total concentration of enzyme, [E]0. 2. For a given [E]0 and low values of [S]0, the rate of product formation is proportional to [S]0. 3. For a given [E]0 and high values of [S]0, the rate of product formation becomes independent of [S]0, reaching a maximum value known as the maximum velocity, vmax. The Michaelis–Menten mechanism accounts for these features.1 According to this mechanism, an enzyme–substrate complex, ES, is formed in the first step and the substrate is released either unchanged or after modification to form products: Michaelis–Menten mechanism: E + S → ES

v = k a[E][S]

ES → E + S

v = k a′[ES]

ES → P + E

v = k b[ES]

As we show in the following Justification, this mechanism implies that the rate of product formation is given by the Michaelis–Menten equation: v=

k b[E]0 1 + KM/[S]0

or

v=

kb[E]0[S]0 [S]0 + KM

Michaelis–Menten equation

(8.1)

where KM = (k a′ + kb)/ka is the Michaelis constant, characteristic of a given enzyme acting on a given substrate, [E]0 is the molar concentration of enzyme added and [S]0 is the molar concentration of the substrate. Justification 8.1 The Michaelis–Menten equation

The variation of the rate of an enzyme-catalyzed reaction with substrate concentration. The approach to a maximum rate, vmax, for large [S]0 is explained by the Michaelis–Menten mechanism. (The constant KM is explained shortly.) Fig. 8.2

To derive eqn 8.1, we note that the rate of product formation (v = kb[ES]) requires us to know [ES]. We can obtain the concentration of the enzyme– substrate complex by invoking the steady-state approximation (Section 7.3c) and writing 1 Michaelis and Menten derived their rate law in 1913 in a more restrictive way, by assuming a rapid equilibrium. The approach we take is a generalization using the steady-state approximation made by Briggs and Haldane in 1925.

8.1 THE MICHAELIS–MENTEN MECHANISM OF ENZYME CATALYSIS

275

d[ES] = k a[E][S] − k a′[ES] − k b[ES] = 0 dt It follows that [ES] =

A ka D [E][S] C k a′ + k bF

where [E] and [S] are the concentrations of free enzyme and substrate, respectively. Now we define the Michaelis constant as KM =

k a′ + k b [E][S] = [ES] ka

and note that KM has the units of a molar concentration. To express the rate law in terms of the concentrations of enzyme, we note that [E]0 = [E] + [ES], so [E]0 = [E] +

[E][S] A [S] D = 1+ [E] C KM KM F

Moreover, because the substrate is typically in large excess relative to the enzyme, the free substrate concentration is approximately equal to the initial substrate concentration and we can write [S] ≈ [S]0. It then follows that [ES] =

1 [E]0[S]0 [E]0 = KM 1 + [S]0 /KM (1 + KM)/[S]0

We obtain eqn 8.1 when we substitute this expression for [ES] into that for the rate of product formation.

Equation 8.1 shows that, in accord with experimental observations: 1. When [S]0 >KM, the rate reaches its maximum value and is independent of [S]0: v = vmax = kb[E]0

(8.2b)

We can rearrange eqn 8.1 into a form that is amenable to data analysis by linear regression. Substitution of the definition of vmax into eqn 8.2b gives v=

vmax 1 + KM/[S]0

Then, on taking reciprocals of both sides, we obtain 1 1 A KM D 1 + = v vmax C vmax F [S]0

Lineweaver–Burk plot

(8.3)

A Lineweaver–Burk plot is a plot of 1/v against 1/[S]0 and, according to eqn 8.3, it should yield a straight line with slope of KM/vmax, a y-intercept at 1/vmax, and an x-intercept at −1/KM (Fig. 8.3). The value of k b is then calculated from the y-intercept and eqn 8.2b. However, the plot cannot give the individual rate constants ka and k a′ that appear in the expression for KM. The stopped-flow technique

Fig. 8.3 A Lineweaver–Burk plot is used to analyze kinetic data on enzyme-catalyzed reactions. The reciprocal of the rate of formation of products (1/v) is plotted against the reciprocal of the substrate concentration (1/[S]0). All the data points (which typically lie in the full region of the line) correspond to the same overall enzyme concentration, [E]0. The intercept of the extrapolated (dotted) straight line with the horizontal axis is used to obtain the Michaelis constant, KM. The intercept with the vertical axis is used to determine vmax = kb[E]0 and hence kb. The slope may also be used, for it is equal to KM/vmax.

276

8 COMPLEX BIOCHEMICAL PROCESSES

described in Section 6.1b gives the additional data needed because we can find the rate of formation of the enzyme–substrate complex by monitoring the concentration after mixing the enzyme and substrate. This procedure gives a value for ka, and k a′ is then found by combining this result with the values of kb and KM. Example 8.1

Analyzing a Lineweaver–Burk plot

The enzyme carbonic anhydrase (Atlas P2) catalyzes the hydration of CO2 in red blood cells to give bicarbonate (hydrogencarbonate) ion: CO2(g) + H2O(l) → HCO3−(aq) + H+(aq) The following data were obtained for the reaction at pH = 7.1, 273.5 K, and an enzyme concentration of 2.3 nmol dm−3: The Lineweaver–Burk plot based on the data in Example 8.1.

Fig. 8.4

[CO2]0 /(mmol dm−3) 1.25 2.5 V/(mmol dm−3 s−1) 2.78 × 10−2 5.00 × 10−2

5 8.33 × 10−2

20 1.67 × 10−1

Determine the maximum velocity and the Michaelis constant for the reaction. Strategy We construct a Lineweaver–Burk plot by drawing up a table of 1/[S]0 and 1/v. The intercept at 1/[S]0 = 0 is 1/vmax and the slope of the line through the points is KM/vmax, so KM is found from the slope divided by the intercept. Solution We draw up the following table:

A note on good practice

The slope and the intercept are unit-less: we have remarked previously that all graphs should be plotted as pure numbers by dividing the physical variables by their units (not just by ignoring the units!).

1/([CO2]0 /(mmol dm−3)) 1/(V/(mmol dm−3 s−1))

0.800 36.0

0.400 20.0

0.200 12.0

0.0500 6.00

The graph is plotted in Fig. 8.4. A least-squares analysis gives an intercept at 4.00 and a slope of 40.0. It follows that vmax /(mmol dm−3 s−1) =

1 1 = = 0.250 intercept 4.00

and KM/(mmol dm−3) =

slope 40.0 = = 10.0 intercept 4.00

Self-test 8.1 The enzyme a-chymotrypsin (Atlas P3) is secreted in the pancreas of mammals and cleaves peptide bonds made between certain amino acids. Several solutions containing the small peptide N-glutaryl-lphenylalanine-p-nitroanilide at different concentrations were prepared, and the same small amount of a-chymotrypsin was added to each one. The following data were obtained on the initial rates of the formation of product:

[S]0 /(mmol dm−3) V/(mmol dm−3 s−1)

0.334 0.450 0.152 0.201

0.667 1.00 1.33 0.269 0.417 0.505

1.67 0.667

Determine the maximum velocity and the Michaelis constant for the reaction. Answer: 2.76 mmol dm−3 s−1, 5.77 mmol dm−3

8.2 THE ANALYSIS OF COMPLEX MECHANISMS

Many enzyme-catalyzed reactions are consistent with a modified version of the Michaelis–Menten mechanism, in which the release of product from the ES complex is also reversible with the step P → ES

v = k b′[P]

added to the mechanism. In Exercise 8.10 you are invited to show that application of the steady-state approximation for [ES] then results in the following expression for the rate of the reaction: v=

(vmax /KM)[S]0 − (v′max /K′M)[P] 1 + [S]0 /KM + [P]/K′M

(8.4a)

where vmax = kb[E]0 KM =

k a′ + kb ka

v′max = k a′[E]0

(8.4b)

k a′ + kb k b′

(8.4c)

K′M =

Equation 8.4a tells us that the reaction rate depends on the concentration of product. However, at the early stages of the reaction, when [S] = [S]0 >> [P], terms containing [P] can be ignored and it is easy to show that eqn 8.4a reduces to eqn 8.1. 8.2 The analysis of complex mechanisms The simple mechanism described in the previous section is only a starting point: to account for the full range of enzyme-controlled reactions, we need to consider more involved mechanisms.

Many enzymes can generate several intermediates as they process a substrate into one or more products. An example is the enzyme chymotrypsin, which we treat in detail in Case study 8.1. Other enzymes act on multiple substrates. An example is hexokinase, which catalyzes the reaction between ATP and glucose (the two substrates of the enzyme), the first step of glycolysis (Section 4.8). The same strategies developed in Section 8.1 can be used to deal with such complex reaction schemes, and we shall focus on reactions involving two substrates. (a) Sequential reactions

In sequential reactions the active site binds all the substrates before processing them into products. The binding can be ordered: Sequential reaction mechanism [E][S1] [ES1]

E + S1 7 ES1

KM1 =

ES1 + S2 7 ES1S2

KM12 =

ES1S2 → E + P

v = kb[ES1S2]

[ES1][S2] [ES1S2]

for the two substrates S1 and S2. (Note that the Michaelis constants are, apart from their units, the reciprocals of the equilibrium constants for each step; we are supposing that there are two fast pre-equilibrium steps.) Alternatively, substrate binding can be random and the following steps can also lead to formation of the ES1S2 complex:

277

278

8 COMPLEX BIOCHEMICAL PROCESSES

Sequential mechanism with random attachment [E][S2] [ES2]

E + S2 7 ES2

KM2 =

ES2 + S1 7 ES1S2

KM21 =

[ES2][S1] [ES1S2]

The resulting rate law, based on the relation [E]0 = [E] + [ES1] + [ES2] + [ES1S2] for the total concentration of enzyme in its bound and unbound forms, is v=

vmax[S1]0[S2]0 KM1KM12 + KM12[S1]0 + KM12[S2]0 + [S1]0[S2]0

(8.5a)

where vmax = kb[E]0 and we have supposed that both S1 and S2 are in such excess over the enzyme concentration that they are equal to their nominal concentrations. This equation can be rearranged into a form more suitable for plotting by holding the concentration of one substrate (S2, for instance) constant and writing first v=

vmax[S1]0 KM1KM12/[S2]0 + KM12[S1]0/[S2]0 + KM21 + [S1]0

and then forming the reciprocal of both sides: 1 1 + KM12 /[S2]0 KM21 + KM1KM12 /[S2]0 1 = + v vmax vmax [S1]0

Analysis of sequential reaction

(8.5b)

It follows that a plot of 1/v against 1/[S1]0 for constant [S2]0 is linear with slope =

KM21 + KM1KM12 /[S2]0 vmax

y-intercept =

1 + KM12 /[S2]0 vmax

(8.5c)

(b) Ping-pong reactions

In so-called ping-pong reactions products are released in a stepwise fashion. In a two-substrate reaction, the first substrate (S1) binds to the enzyme E and a product (P1) is released, leaving the enzyme chemically modified (denoted E*), perhaps by a fragment of the substrate. Then the second substrate (S2) binds to the modified enzyme and is processed into a second product, P2, returning the enzyme to its native form. The scheme can be summarized as follows: Ping-pong mechanism [E][S1] [ES1]

E + S1 7 ES1

KM1 =

ES1 → E* + P1

v1 = k b1[ES1]

E* + S2 7 E*S2

KM2 =

E*S2 → E + P2

v2 = k b2[E*S2]

[E*][S2] [E*S2]

8.3 THE CATALYTIC EFFICIENCY OF ENZYMES

279

Enzymes with ping-pong mechanisms include various transferases, oxidoreductases, and proteases. The intermediate E* in the action of the protease chymotrypsin (Atlas P3), for instance, is formed by modification of a serine residue in the active site. If we suppose that ES1 rapidly turns into E*, we may identify [E*] with the value of [ES1] due to the first rapid equilibrium and write [E]0 = [E] + [E*] + [E*S2] = [E] + [ES1] + [E*S2] = [E] +

[E][S1] [E*][S2] [E][S1] [ES1][S2] + = [E] + + KM1 KM2 KM1 KM2

A [S ] [S ][S ] D = 1 + 1 + 1 2 [E] C KM1 KM1KM2 F The rate of the production of P2 is then given by v2 = k b2[E*S2] =

kb2 k kb2 [E*][S2] = b2 [ES1][S2] = [E][S1][S2] KM2 KM2 KM1KM2

which, with the value of [E] replaced by the expression derived above, becomes v2 =

v2max[S1][S2] KM2[S1] + KM1[S2] + [S1][S2]

(8.6a)

with v2max = kb2[E]0. Again, we can rearrange this equation to obtain 1 1 + KM2/[S2] KM1 1 = + · v2 v2max v2max [S1]

Analyzing a ping-pong mechanism

(8.6b)

It follows that a plot of 1/v2 against 1/[S1] for constant [S2] is linear with slope =

KM1 v2max

y-intercept =

1 + KM2/[S2] v2max

(8.6c)

Equations 8.5 and 8.6 form the basis of a graphical method for distinguishing between sequential and ping-pong reactions. For sequential reactions, the slope of a plot of 1/v against 1/[S1] depends on [S2], so a series of such plots for different values of [B] form a family of nonparallel lines (Fig. 8.5a). However, for ping-pong reactions the lines described by plots of 1/v2 against 1/[S1] for different values of [S2] are parallel because the slopes are independent of [S2] (Fig. 8.5b). 8.3 The catalytic efficiency of enzymes To discuss the effectiveness of enzymes, it is useful to have a quantitative measure of their kinetic efficiencies for the acceleration of biochemical reactions.

The turnover frequency, or catalytic constant, of an enzyme, kcat, is the number of catalytic cycles (turnovers) performed by the active site in a given interval divided by the duration of the interval. This quantity has the same units as a first-order rate constant and, in terms of the Michaelis–Menten mechanism, is numerically equivalent to k b, the rate constant for release of product from the enzyme–substrate complex. It follows from the identification of kcat with kb and from eqn 8.2b that

Fig. 8.5 The analysis of kinetic data for enzyme-catalyzed reactions involving two substrates. Plots of 1/v against 1/[S1]0 for different values of [S2]0 can be used to distinguish between (a) a sequential reaction, which gives rise to a family of nonparallel lines, and (b) a ‘ping-pong’ reaction, which give rise to a family of parallel lines.

280

8 COMPLEX BIOCHEMICAL PROCESSES

kcat = k b =

vmax [E]0

Turnover frequency

(8.7)

The catalytic efficiency, h (eta), of an enzyme is the ratio kcat /KM. The higher the value of h, the more efficient is the enzyme. We can think of the catalytic activity as the effective rate constant of the enzymatic reaction. From KM = (k a′ + k b)/ka and eqn 8.7, it follows that h=

kcat kakb = KM k a′ + kb

Catalytic efficiency

(8.8)

The catalytic efficiency reaches its maximum value of ka when kb >> k a′. Because ka is the rate constant for the formation of a complex from two species that are diffusing freely in solution, the maximum efficiency is related to the maximum rate of diffusion of E and S in solution (Section 7.5). This limit leads to rate constants of about 108–109 dm3 mol−1 s−1 for molecules as large as enzymes at room temperature. The enzyme catalase has h = 4.0 × 108 dm3 mol−1 s−1 and is said to have attained ‘catalytic perfection’ in the sense that the rate of the reaction it catalyzes is essentially diffusion controlled: it acts as soon as a substrate makes contact. Self-test 8.2 Calculate kcat and the catalytic efficiency of carbonic anhydrase by using the data from Example 8.1.

Answer: kcat = 1.1 × 105 s−1, h = 1.1 × 104 dm3 mmol−1 s−1

8.4 Enzyme inhibition We now need to take the analysis a stage further to see how to accommodate reaction steps that prevent an enzyme from forming product.

An inhibitor, I, decreases the rate of product formation from the substrate by binding to the enzyme, to the ES complex, or to the enzyme and ES complex simultaneously. The most general kinetic scheme for enzyme inhibition is then Reaction with inhibition E + S → ES

v = ka[E][S]

ES → E + S

v = k a′[ES]

ES → E + P

v = kb[ES]

EI 7 E + I

KI =

[E][I] [EI]

ESI 7 ES + I

K I′ =

[ES][I] [ESI]

The lower the values of KI and K I′, the more efficient is the inhibition. As shown in the following Justification, the rate of reaction in the presence of an inhibitor is v=

vmax a′ + aKM/[S]0

Inhibited rate

(8.9a)

where a=1+

[I] KI

a′ = 1 +

[I] K I′

(8.9b)

8.4 ENZYME INHIBITION

281

This equation is very similar to the Michaelis–Menten equation for the uninhibited enzyme (eqn 8.1) and is also amenable to analysis by a version of the Lineweaver–Burk plot: 1 a′ aKM 1 + = v vmax vmax [S]0

Lineweaver–Burk plot with inhibition

(8.9c)

Justification 8.2 Enzyme inhibition

By mass balance, the total concentration of enzyme is [E]0 = [E] + [EI] + [ES] + [ESI] By using the definitions in eqn 8.9 and the two equilibrium constants it follows that [E]0 = a[E] + a′[ES] Then, because KM = [E][S]/[ES] and [S] ≈ [S]0, we can write [E]0 =

aKM[ES] A aKM D + a′[ES] = + a′ [ES] C [S]0 F [S]0

The expression for the rate of product formation is then v = kb[ES] =

kb[E]0 aKM/[S]0 + a′

which, on rearrangement, gives eqn 8.9c.

There are three major modes of inhibition that give rise to distinctly different kinetic behavior (Fig. 8.6): • Competitive inhibition: the inhibitor binds only to the active site of the enzyme and thereby inhibits the attachment of the substrate. This condition corresponds to a > 1 and a′ = 1 (because ESI does not form). The slope of the Lineweaver–Burk plot increases by a factor of a relative to the slope for data on the uninhibited enzyme (a = a′ = 1). The y-intercept does not change as a result of competitive inhibition. • Uncompetitive inhibition: the inhibitor binds to a site of the enzyme that is removed from the active site but only if the substrate is already present. The inhibition occurs because ESI reduces the concentration of ES, the active type of complex. In this case a = 1 (because EI does not form) and a′ > 1. The y-intercept of the Lineweaver–Burk plot increases by a factor of a′ relative to the y-intercept for data on the uninhibited enzyme, but the slope does not change. • Non-competitive inhibition (or mixed inhibition): the inhibitor binds to a site other than the active site, and its presence reduces the ability of the substrate to bind to the active site. Inhibition occurs at both the E and ES sites. This condition corresponds to a > 1 and a′ > 1. Both the slope and y-intercept of the Lineweaver–Burk plot increase on addition of the inhibitor. Figure 8.6c shows the special case of KI = K I′ and a = a′, which results in intersection of the lines at the x-axis.

Fig. 8.6 Lineweaver–Burk plots characteristic of the three major modes of enzyme inhibition: (a) competitive inhibition, (b) uncompetitive inhibition, and (c) noncompetitive inhibition, showing the special case a = a′ > 1.

282

8 COMPLEX BIOCHEMICAL PROCESSES

A brief comment

Because several plots are involved in the extraction of information from the data, it is sometimes difficult to keep track of units. The following lengthy Example shows in detail how to keep track of them in analyses of this kind (and in general).

In all cases, the efficiency of the inhibitor may be obtained by determining KM and vmax from a control experiment with uninhibited enzyme and then repeating the experiment with a known concentration of inhibitor. From the slope and y-intercept of the Lineweaver–Burk plot for the inhibited enzyme (eqn 8.9), the mode of inhibition, the values of a or a′, and the values of KI or K I′ can be obtained.

Example 8.2

Distinguishing between types of inhibition

Five solutions of a substrate, S, were prepared with the concentrations given in the first column below, and each one was divided into five equal volumes. The same concentration of enzyme was present in each one. An inhibitor, I, was then added in different concentrations to the samples, and the initial rate of formation of product was determined with the results given below. Does the inhibitor act competitively or noncompetitively? Determine KI and KM. −3

[S]/(mmol dm ) 0 0.050 0.033 0.10 0.055 0.20 0.083 0.40 0.111 0.60 0.126

[I]/(mmol dm−3) 0.20 0.40 0.60 0.026 0.021 0.018 0.045 0.038 0.033 0.071 0.062 0.055 0.100 0.091 0.084 0.116 0.108 0.101

0.80 0.016 ⎤ 0.029 ⎥ 0.050 ⎥ v/(mmol dm−3 s−1) ⎥ 0.077 ⎥ 0.094 ⎦

Strategy We draw a series of Lineweaver–Burk plots for different inhibitor concentrations. If the plots resemble those in Fig. 8.6a, then the inhibition is competitive. On the other hand, if the plots resemble those in Fig. 8.6c, then the inhibition is noncompetitive. To find KI, we need to determine the slope at each value of [I], which is equal to aKM/vmax, or KM/vmax + KM[I]/KI vmax, then plot this slope against [I]: the intercept at [I] = 0 is the value of KM/vmax and the slope is KM/KIvmax. To conform to the rule that all graphs should be plots of dimensionless quantities, we express eqn 8.9c as

1 a′ aKM/mmol dm−3 1 + = −3 −1 −3 −1 −3 −1 v/mmol dm s vmax /mmol dm s vmax /mmol dm s [S]0 /mmol dm−3 and plot 1/(v/mmol dm−3 s−1) against 1/([S]0/mmol dm−3), then the slope (which we shall call slope1 for this first graph) is identified with slope1 =

aKM/mmol dm−3 , vmax/mmol dm−3 s−1

so

a = slope1 ×

vmax/mmol dm−3 s−1 KM/mmol dm−3

and the y-intercept with intercept1 =

a′ , vmax/mmol dm−3 s−1

so

vmax =

a′ mmol dm−3 s−1 intercept1

For the determination of KI and KM we write eqn 8.9b, a = 1 + [I]/KI, which becomes slope1 ×

vmax /mmol dm−3 s−1 [I]/mmol dm−3 =1+ −3 KM/mmol dm KI/mmol dm−3

8.4 ENZYME INHIBITION

283

and therefore slope1 =

KM/mmol dm−3 KM/mmol dm−3 [I]/mmol dm−3 + × vmax/mmol dm−3 s−1 vmax /mmol dm−3 s−1 KI/mmol dm−3

=

KM/mmol dm−3 KM/KI + × [I]/mmol dm−3 vmax /mmol dm−3 s−1 vmax/mmol dm−3 s−1

We conclude that when in the second graph slope1 is plotted against [I]/mmol dm−3, slope2 =

KM/KI , vmax/mmol dm−3 s−1

intercept2 =

KM/KI = slope2 × vmax/mmol dm−3 s−1

so

KM/mmol dm−3 , vmax/mmol dm−3 s−1

so

KM dm−3 = intercept2 × vmax /mmol dm−3 s−1 mmol Solution First, we draw up a table of 1/[S] and 1/v for each value of [I]:

[I]/(mmol dm−3) 0.20 0.40 0.60 1/([S]/(mmol dm )) 0 20 30 38 48 56 10 18 22 26 30 5.0 12 14 16 18 2.5 9.01 10.0 11.0 11.9 1.7 7.94 8.62 9.26 9.90 −3

0.80 62 ⎤ 34 ⎥ 20 ⎥ 1/(v/(mmol dm−3 s−1)) ⎥ 13.0 ⎥ 10.6 ⎦

The five plots (one for each [I]) are given in Fig. 8.7. We see that they pass through the same intercept on the vertical axis, so the inhibition is competitive and we can set a′ = 1. The mean of the (least-squares) intercepts is intercept1 = 5.83, so vmax =

Fig. 8.7 Lineweaver–Burk plots for the data in Example 8.2. Each line corresponds to a different concentration of inhibitor.

1 mmol dm−3 s−1 = 0.17 mmol dm−3 s−1 5.83

The (least-squares) slopes of the lines are as follows: [I]/mmol dm−3 slope1

0 1.219

0.20 1.627

0.40 2.090

0.60 2.489

0.80 2.832

These values are plotted in Fig. 8.8. The intercept at [I] = 0 is intercept2 = 1.234, so KM/mmol dm−3 = 1.234 × 0.17 = 0.21 and therefore KM = 0.21 mmol dm−3. The (least-squares) slope of the line is slope2 = 2.045, so KM/KI = 2.045 × 0.17 = 0.348 It follows that KI =

KM 0.21 mmol dm−3 = 2.045 × 0.17 2.045 × 0.17 = 0.60 mmol dm−3

Fig. 8.8 Plot of the slopes of the plots in Fig. 8.7 against [I] based on the data in Example 8.2.

284

8 COMPLEX BIOCHEMICAL PROCESSES

Repeat the question using the following data:

Self-test 8.3

−3

[S]/(mmol dm ) 0 0.050 0.020 0.10 0.035 0.20 0.056 0.40 0.080 0.60 0.093

[I]/(mmol dm−3) 0.20 0.40 0.60 0.015 0.012 0.0098 0.026 0.021 0.017 0.042 0.033 0.028 0.059 0.047 0.039 0.069 0.055 0.046

0.80 0.0084 ⎤ 0.015 ⎥ 0.024 ⎥ v/(mmol dm−3 s−1) ⎥ 0.034 ⎥ 0.039 ⎦

Answer: Noncompetitive, KM = 0.30 mmol dm−3, KI = 0.57 mmol dm−3

Case study 8.1

The molecular basis of catalysis by hydrolytic enzymes

One protein enzyme that has been studied in considerable detail is chymotrypsin (Atlas P3), which functions by hydrolyzing peptide bonds in polypeptides in the small intestine. The sequence of steps by which the enzyme carries out the first part of its task—to snip through the C–N bond of the peptide link—is shown in Fig. 8.9. The crucial point to notice is the formation of a tetrahedral transition state in the course of the reaction. The second sequence by which the carboxylic acid group is eliminated from the polypeptide is shown in Fig. 8.10. This step involves the attack by an H2O molecule on the carboxyl group and the subsequent cleavage of the original C–O bond. Once again, the crucial point is the formation of a tetrahedral transition state. In each case, the catalytic activity of the enzyme can be traced to the structure of the active site, in this case featuring a catalytic triad, which enhances reactivity of the enzyme toward the substrate, and an oxoanion hole, which stabilizes the tetrahedral transition state.

Fig. 8.9 The sequence of steps by which chymotrypsin cuts through the C–N bond of a peptide link and releases an amine.

The catalytic triad consists of the serine, histidine, and aspartic acid residues shown in Figs 8.9 and 8.10. There, proton transfer between the residues deprotonates serine’s hydroxyl group, resulting in an alkoxide ion that is particularly reactive toward the carbonyl group of the polypeptide. In the oxoanion hole, NH groups from the peptide backbone of the enzyme are strategically placed to form hydrogen bonds with the negatively charged oxygen atom (formerly the carbonyl oxygen of the polypeptide substrate) of the tetrahedral transition state. By helping to accommodate a nascent negative charge, the oxoanion hole lowers the energy of the transition state and enhances the rate of hydrolysis. The entities known as ‘catalytic antibodies’ combine the insight that studies on molecules such as chymotrypsin provide with an organism’s natural defence system. In that way, they open routes to alternative enzymes for carrying out particular reactions. The key idea is that an organism generates a flood of antibodies when an antigen—a foreign body—is introduced. The organism maintains a wide range of latent antibodies, but they proliferate in the presence of the antigen. It follows that, if we can introduce an antigen that emulates the tetrahedral transition state typical of a peptide hydrolysis reaction, then an organism should produce a supply of antibodies that may be able to act as enzymes for that and related functions. This procedure has been applied to the search for enzymes for the hydrolysis of esters. The compound used to mimic the tetrahedral transition state is a tetrahedral phosphonate (1). When the antibody stimulated to interact with

8.5 MOLECULAR MOTION IN LIQUIDS

285

this antigen is used to catalyze the hydrolysis of an ester, pronounced activity is indeed found, with KM = 1.9 mmol dm−3 and an enhancement of rate over the uncatalyzed reaction by a factor of 103. The hope is that catalytic antibodies can be formed that catalyze reactions currently untouched by enzymes, such as those that target destruction of viruses and tumors.

Transport across biological membranes At this stage we can begin to explore the molecular features that govern the rates of reactions. We saw in Chapter 5 that many cellular processes, such as the propagation of impulses in neurons and the synthesis of ATP by ATPases, are controlled by the transport of molecules and ions across biological membranes. Passive transport is the spontaneous movement of species down concentration and membrane potential gradients; active transport is nonspontaneous movement against these gradients and driven by the hydrolysis of ATP. Here we complement the thermodynamic treatment of Chapter 5 with a kinetic analysis that begins with a consideration of the laws governing the motion of molecules and ions in liquids and then describes modes of transport across cell membranes. 8.5 Molecular motion in liquids Because the rate at which molecules move in solution may be a controlling factor of the maximum rate of a biochemical reaction in the intracellular medium, we need to understand the factors that limit molecular motion in a liquid.

A molecule in a liquid is surrounded by other molecules and can move only a fraction of a diameter in each step it takes, perhaps because its neighbors briefly move aside. Molecular motion in liquids is a series of short steps, with ever-changing directions, like people in an aimless, milling crowd. The process of migration by means of a random jostling motion through a liquid is called diffusion. We can think of the motion of the molecule as a series of short jumps in random directions, a so-called random walk (Fig. 8.11). If there is an initial concentration gradient in the liquid (for instance, a solution may have a high concentration of solute in one region), then the rate at which the molecules spread out is proportional to the concentration gradient and we write

Fig. 8.10 The following sequence of steps by which chymotrypsin cuts through the C–O bond and releases a carboxylic acid.

rate of diffusion ∝ concentration gradient To express this relation mathematically, we introduce the flux, J, which is the number of particles passing through an imaginary window in a given time interval, divided by the area of the window and the duration of the interval: J=

number of particles passing through window area of window × time interval

Definition of flux

(8.10a)

The flux may also be expressed in terms of the amount (in moles) of molecules: J=

amount of particles passing through window area of window × time interval

One possible path of a random walk in three dimensions. In this general case, the step length is also a random variable. (Available at http:// www.ki.inf.tu-dresden. de/~fritzke/research/TS/ example1.html.) Fig. 8.11

(8.10b)

To calculate the number or amount of molecules passing through a given window in a given time interval, we multiply the flux by the area of the window

286

8 COMPLEX BIOCHEMICAL PROCESSES

Diffusion coefficients in water, D/(10−9 m2 s−1) Table 8.1

Water, H2O*

2.26

Glycine, NH2CH2COOH*

1.055

Sucrose, C12H22O11*

0.522

Lysozyme†

0.112

Serum albumin†

0.0594

Catalase†

0.0410

Fibrinogen† Bushy stunt virus

0.0202 †

0.0115

*Measured at 5°C. † Measured at 20°C.

and the time interval. Fick’s first law of diffusion (see Further information 8.1 for a derivation) then states: J = −D

dc dx

Fick’s first law

(8.11)

where dc/dx is the gradient of either the number concentration (molecules m−3, for instance) when the flux is in terms of numbers or the molar concentration (mol dm−3, for instance) when the flux is in terms of amounts. The coefficient D, which has the dimensions of area divided by time (typically with units m2 s−1), is called the diffusion coefficient (Table 8.1) and depends on the solute species, the solvent, and the temperature. For a given concentration gradient, large values of D correspond to rapid diffusion. The negative sign in eqn 8.11 simply means that if the concentration gradient is negative (down from left to right, Fig. 8.12), then the flux is positive (flowing from left to right).

A brief illustration

For sucrose in water at 25°C, D = 5.22 × 10−10 m2 s−1. Suppose that in a region of an unstirred aqueous solution of sucrose the molar concentration gradient is −0.10 mol dm−3 cm−1. Then, because 1 dm = 10−1 m (so 1 dm−3 = 103 m−3) and 1 cm = 10−2 m (so 1 cm−1 = 102 m−1), the flux arising from this gradient is J = −(5.22 × 10−10 m2 s−1) × (−0.10 mol dm−3 cm−1) = 5.22 × 0.10 × 10−10 m2 s−1 mol × (103 m−3) × (102 m−1) = 5.2 × 10−6 mol m−2 s−1 The amount of sucrose molecules passing through a 1.0-cm square window in 10 minutes is therefore n = JADt = (5.2 × 10−6 mol m−2 s−1) × (1.0 × 10−2 m)2 × (10 × 60 s) = 3.1 × 10−7 mol, or 0.31 mmol

The flux of solute particles is proportional to the concentration gradient. Here we see a solution in which the concentration falls from left to right. The gradient is negative (down from left to right) and the flux is positive (towards the right). The greatest flux is found where the gradient is steepest (at the left).

Fig. 8.12

Diffusion coefficients are of the greatest importance for discussing the spread of pollutants in lakes and through the atmosphere. In both cases, the spread of pollutant may be assisted—and is normally greatly dominated—by bulk motion of the fluid as a whole (as when a wind blows in the atmosphere). This motion is called convection. Because diffusion is often a slow process, we speed up the spread of solute molecules by inducing convection by stirring a fluid or turning on an extractor fan. One of the most important equations in the physical chemistry of fluids is the diffusion equation, which enables us to predict the rate at which the concentration of a solute changes in a nonuniform solution. In essence, the diffusion equation expresses the fact that wrinkles in the concentration tend to disperse. The formal statement of the diffusion equation, which is also known as Fick’s second law of diffusion, is ∂c ∂ 2c =D 2 ∂t ∂x

Fick’s second law

(8.12)

8.5 MOLECULAR MOTION IN LIQUIDS

Mathematical toolkit 8.1

Partial derivatives

When a function depends on more than one variable, such as the function f(x,y), there are two first derivatives: one with respect to x with y held constant, and the other with respect to y with x held constant. These derivatives are referred to as ‘partial derivatives’ and denoted A ∂f D C ∂x F y

and

287

Higher derivatives are defined analogously, as in A ∂ 2f D C ∂x 2 F y

and

A ∂ 2f D C ∂y 2 F x

A ∂f D C ∂y F x

respectively (note the ‘curly d’) and the variable held constant as a right-subscript (this may be omitted if there is no ambiguity). The first of these expressions is the slope of the function parallel to the x-axis and the second is the slope parallel to the y-axis (see the illustration). For instance, if f = x 2y 3, then A ∂f D = 2xy 3 C ∂x F y

and

A ∂f D = 3x 2y 2 C ∂y F x

where ∂c/∂t is the rate of change of concentration in a region and ∂ 2c/∂x 2 may be thought of as the curvature of the concentration in the region. Because the concentration c depends on both position and time, we have to express Fick’s second law using ‘partial derivative’ notation (see Mathematical toolkit 8.1). The ‘curvature’ ∂ 2c/∂x 2 is a measure of the wrinkliness of the concentration (see below). The derivation of this expression from Fick’s first law is also given in Further information 8.1. The concentrations on the left and right of this equation may be either number concentrations or molar concentrations. The diffusion equation tells us that a uniform concentration and a concentration with unvarying slope through the region (so ∂ 2c/∂x 2 = 0 in each case) results in no net change in concentration in the region (∂c/∂t = 0) because the rate of influx through one wall of the region is equal to the rate of efflux through the opposite wall. Only if the slope of the concentration varies through a region— only if the concentration is wrinkled—is there a change in concentration. Where the curvature is positive (a dip, Fig. 8.13), the change in concentration is positive: the dip tends to fill. Where the curvature is negative (a heap), the change in concentration is negative: the heap tends to spread. The diffusion coefficient increases with temperature because an increase in temperature enables a molecule to escape more easily from the attractive forces exerted by its neighbors. If we suppose that the rate of random motion follows an Arrhenius temperature dependence with an activation energy Ea, then the diffusion coefficient will follow the relation D = D0e−E /RT a

Temperaturedependence of D

(8.13)

Nature abhors a wrinkle. The diffusion equation tells us that peaks in a distribution (regions of negative curvature) spread and troughs (regions of positive curvature) fill in. Fig. 8.13

288

8 COMPLEX BIOCHEMICAL PROCESSES

The rate at which particles diffuse through a liquid is related to the viscosity, and we should expect a high diffusion coefficient to be found for fluids that have a low viscosity. That is, we can suspect that D ∝ 1/h, where h (eta) is the coefficient of viscosity. In fact, the Stokes–Einstein relation states that D=

kT 6pha

Stokes–Einstein relation

(8.14)

where a is the effective radius of the molecule. Figure 8.14 shows the observed temperature dependence of the viscosity of water. Self-test 8.4 Estimate the activation energy for the diffusion of a solute molecule in water from the graph in Fig. 8.14 by using the viscosities at 40°C and 80°C. Hint: Use an equation such as eqn 8.13 to formulate an expression for the logarithm of the ratio of the two diffusion coefficients.

Answer: 19 kJ mol−1

The experimental temperature dependence of the viscosity of water. As the temperature is increased, more molecules are able to escape from the potential wells provided by their neighbors, so the liquid becomes more fluid.

Fig. 8.14

8.6 Molecular motion across membranes A crucial aspect of biochemical change is the rate at which species are transported across a membrane, so we need to understand the kinetic factors that facilitate or impede transport.

Consider the passive transport of an uncharged species A across a lipid bilayer of thickness l. To simplify the problem, we assume that the concentration of A is always maintained at [A] = [A]0 on one surface of the membrane and at [A] = 0 on the other surface, perhaps by a perfect balance between the rate of the process that produces A on one side and the rate of another process that consumes A completely on the other side. Then ∂[A]/∂t = 0 because the two boundary conditions ensure that the interior of the membrane is maintained at a constant but not necessarily uniform concentration, and eqn 8.12 simplifies to D

d2[A] =0 dx 2

(8.15)

where D is the diffusion coefficient. (We can use d in place of the partial ∂ because this [A] is independent of time, so x is the only variable.) We use the conditions [A](0) = [A]0 and [A](l) = 0 to solve this differential equation and the result, which may be verified by differentiation, is A xD [A](x) = [A]0 1 − C lF

Concentration profile

(8.16)

which implies that [A] decreases linearly inside the membrane. We now use Fick’s first law to calculate the flux J of A through the membrane. From eqn 8.16, it follows that d[A] [A]0 =− dx l and from this result and eqn 8.11 obtain J=D

[A]0 l

(8.17)

8.6 MOLECULAR MOTION ACROSS MEMBRANES

289

Before using this simple result we need to take into account the fact that the concentration of A on the surface of a membrane is not always equal to its concentration measured in the bulk solution, which we assume to be aqueous. This difference arises from the significant difference in the solubility of A in an aqueous environment and in the solution–membrane interface. One way to deal with this problem is to define a partition coefficient k (kappa) as k=

[A]0 [A]s

Definition of partition coefficient

(8.18)

where [A]s is the molar concentration of A in the bulk aqueous solution. It follows that J = kD

[A]s l

Diffusion flux

(8.19)

We see, as intuition would suggest, that the flux is high when the concentration of A in the bulk solution is high and the membrane is thin. In spite of the assumptions that led to its final form, eqn 8.19 describes adequately the passive transport of many nonelectrolytes through membranes of blood cells. In many cases, however, eqn 8.19 underestimates the flux, which suggests that the membrane is more permeable than expected. However, because the permeability increases only for certain species, we can infer that in these cases, transport is facilitated by carrier molecules. One example is the transporter protein that carries glucose into cells. But we issue a word of caution: there is little justification for supposing that D in the membrane is equal to its value in aqueous solution or that k has any particular value, and the conclusion that facilitated transport is involved needs additional evidence before it can be accepted. To treat facilitated transport we suppose that a characteristic of a carrier C is that it binds to the transported species A and that the dissociation of the AC complex is described by AC 7 A + C

K=

[A][C] [AC]c 3

(8.20a)

where we have used concentrations instead of activities (and c 3 = 1 mol dm−3, the standard molar concentration). After writing [C]0 = [C] + [AC], where [C]0 is the total concentration of carrier, it follows that [AC] =

[A][C]0 [A] + Kc 3

(8.20b)

Then the flux through the membrane of the species AC is given by a version of eqn 8.19 as J = kAC DAC

[AC] kACDAC[C]0 [A] [A] = = Jmax 3 l l [A] + Kc [A] + Kc3

Mediated flux

(8.21)

where kAC and DAC are the partition coefficient and diffusion coefficient of the species AC, respectively and Jmax = kACDAC[C]0 /l. We see from Fig. 8.15 that: • when [A] > Kc 3, J = Jmax and the flux has its maximum value. This behavior is characteristic of mediated transport.

Fig. 8.15 The flux of the species AC through a membrane varies with the concentration of the species A. The behavior shown in the figure and explained in the text is characteristic of mediated transport of A, with C as a carrier molecule.

290

8 COMPLEX BIOCHEMICAL PROCESSES

8.7 The mobility of ions Ion transport through membranes is central to the operation of many biological processes, particularly signal transduction in neurons, and we need to be equipped to describe ion migration quantitatively.

An ion in solution responds to the presence of an electric field, migrates through the solution, and carries charge from one location to another. The study of the motion of ions down a potential gradient gives an indication of their size, the effect of solvation, and details of the type of motion they undergo. When an ion is subjected to an electric field E, it accelerates. However, the faster it travels through the solution, the greater the retarding force it experiences from the viscosity of the medium. As a result, as we show in the following Justification, the ion settles down into a limiting velocity called its drift velocity, s, which is proportional to the strength of the applied field: s = uE

Definition of mobility

(8.22)

The mobility, u, depends on the radius, a, and charge number, z, of the ion and the viscosity, h, of the solution: u=

ez 6pha

Relation of mobility to size and viscosity

(8.23)

Justification 8.3 The ionic mobility

An ion of charge ze in an electric field E (typically in volts per meter, V m−1) experiences a force of magnitude zeE, which accelerates it. However, the ion experiences a frictional force due to its motion through the medium, and that retarding force increases the faster the ion travels. The viscous drag (the retarding force), F, on a spherical particle of radius a traveling at a speed s is given by Stokes’ law: F = 6phas When the particle has reached its drift speed, the accelerating and viscous retarding forces are equal, so we can write zeE = 6phas and solve this expression for s: s=

ezE 6pha

At this point we can compare this expression for the drift speed with eqn 8.22 and hence find the expression for mobility given in eqn 8.23.

Equation 8.23 tells us that the mobility of an ion is high if it is highly charged, is small, and is in a solution with low viscosity. These features appear to contradict the trends in Table 8.2, which lists the mobilities of a number of ions. For instance, the mobilities of the Group 1 cations increase down the group despite their increasing radii (Section 9.12). The explanation is that the radius to use in eqn 8.23 is the hydrodynamic radius, the effective radius for the migration of the ions taking into account the entire object that moves. When an ion migrates, it carries

8.7 THE MOBILITY OF IONS

Table 8.2

291

Ionic mobilities in water at 298 K, u/(10−8 m2 s−1 V−1)

Cations

Anions

H+ (H3O+)

36.23

+

OH−

20.64

Li

4.01

F−

Na+

5.19

Cl−

7.92

+

K

7.62

Br

8.09

Rb+

8.06

I−

Cs+

8.00

CO32−

2+

5.74

7.96 7.41 8.29

Mg

5.50

NO

Ca2+

6.17

SO42−

Sr2+

6.16

NH4+

7.62

[N(CH3)4]+ [N(CH2CH3)4]

7.46

− 3

4.65 +

3.38

its hydrating water molecules with it, and as small ions are more extensively hydrated than large ions (because they give rise to a stronger electric field in their vicinity), ions of small radius actually have a large hydrodynamic radius. Thus, hydrodynamic radius decreases down Group 1 because the extent of hydration decreases with increasing ionic radius. One significant deviation from this trend is the very high mobility of the proton in water. It is believed that this high mobility reflects an entirely different mechanism for conduction, the Grotthus mechanism, in which the proton on one H2O molecule migrates to its neighbors, the proton on that H2O molecule migrates to its neighbors, and so on along a chain (Fig. 8.16). The motion is therefore an effective motion of a proton, not the actual motion of a single proton.

In the laboratory 8.1

Electrophoresis

An important application of the preceding material is to the determination of the molar mass of biological macromolecules. Electrophoresis is the motion of a charged species, such as DNA and ionic forms of amino acids, in response to an electric field. Electrophoretic mobility is a result of a constant drift speed, so the mobility of a macromolecule in an electric field depends on its net charge, size (and hence molar mass), and shape. Electrophoresis is a very valuable tool for the separation of biopolymers from complex mixtures, such as those resulting from fractionation of biological cells. We shall consider several strategies controlling the drift speeds of biomolecules in order to achieve separation of a mixture into its components. In gel electrophoresis, migration takes place through a slab of a porous gel, a semi-rigid dispersion of a solid in a liquid. Because the molecules must pass through the pores in the gel, the larger the macromolecule, the less mobile it is in the electric field and, conversely, the smaller the macromolecule, the more swiftly it moves through the pores. In this way, gel electrophoresis allows for

Fig. 8.16 A simplified version of the Grotthus mechanism of proton conduction through water. The proton entering the chain at the top is not the same as the proton leaving the chain at the bottom.

292

8 COMPLEX BIOCHEMICAL PROCESSES

the separation of components of a mixture according to their molar masses. Two common gel materials for the study of proteins and nucleic acids are agarose and cross-linked polyacrylamide. Agarose has large pores and is better suited for the study of large macromolecules, such as DNA and enzyme complexes. Polyacrylamide gels with varying pore sizes can be made by changing the concentration of acrylamide in the polymerization solution. In general, smaller pores form as the concentration of acrylamide is increased, making possible the separation of relatively small macromolecules by polyacrylamide gel electrophoresis (PAGE). The separation of very large pieces of DNA, such as chromosomes, by conventional gel electrophoresis is not effective, making the analysis of genomic material rather difficult. Double-stranded DNA molecules are thin enough to pass through gel pores, but long and flexible DNA coils can become trapped in the pores and the result is impaired mobility along the direction of the applied electric field. This problem can be avoided with pulsed-field electrophoresis, in which a brief burst of the electric field is applied first along one direction and then along a perpendicular direction. In response to the switching back and forth between field directions, the DNA coils writhe about and eventually pass through the gel pores. In this way, the mobility of the macromolecule can be related to its molar mass. We have seen that charge also determines the drift speed. For example, proteins of the same size but different net charge travel along the slab at different speeds. One way to avoid this problem and to achieve separation by molar mass is to denature the proteins in a controlled way. Sodium dodecyl sulfate is an anionic detergent that is very useful in this respect: it denatures proteins, whatever their initial shapes, into rods by forming a complex with them. Moreover, most protein molecules bind a constant number of ions, so the net charge per protein is well regulated. Under these conditions, different proteins in a mixture may be separated according to size only. The molar mass of each constituent protein is estimated by comparing its mobility in its rodlike complex form with a standard sample of known molar mass. However, molar masses obtained by this method, often referred to as SDS-PAGE when polyacrylamide gels are used, are not as accurate as those obtained by the sophisticated techniques discussed in Chapter 11.

The plot of the speed of a moving macromolecule against pH allows the isoelectric point to be detected as the pH at which the speed is zero. The data are from Example 8.3.

Fig. 8.17

Another technique that deals with the effect of charge on drift speed takes advantage of the fact that the overall charge of proteins and other biopolymers depends on the pH of the medium. For instance, in acidic environments protons attach to basic groups and the net charge is positive; in basic media the net charge is negative as a result of proton loss. At the isoelectric point, the pH is such that there is no net charge on the biopolymer. Consequently, the drift speed of a biopolymer depends on the pH of the medium, with s = 0 at the isoelectric point (see Example 8.3 and Fig. 8.17). Isoelectric focusing is an electrophoresis method that exploits the dependence of drift speed on pH. In this technique, a mixture of proteins is dispersed in a medium with a pH gradient along the direction of an applied electric field. Each protein in the mixture will stop moving at a position in the gradient where the pH is equal to the isoelectric point. In this manner, the protein mixture can be separated into its components.

8.7 THE MOBILITY OF IONS

Example 8.3

293

The isoelectric point of a protein

The speed with which bovine serum albumin (BSA) moves through water under the influence of an electric field was monitored at several values of pH, and the data are listed below. What is the isoelectric point of the protein? pH Velocity/(mm s−1)

4.20 0.50

4.56 0.18

5.20 −0.25

5.65 −0.65

6.30 −0.90

7.00 −1.25

Strategy If we plot speed against pH, we can use interpolation to find the pH

at which the speed is zero, which is the pH at which the molecule has zero net charge. Solution The data are plotted in Fig. 8.17. The velocity passes through zero at pH = 4.8; hence pH = 4.8 is the isoelectric point.

Self-test 8.5

The following data were obtained for another protein:

pH Velocity/(mm s−1)

4.5 −0.10

5.0 −0.20

5.5 −0.30

6.0 −0.35

Estimate the pH of the isoelectric point. Answer: 4.1

The separation of complicated mixtures of macromolecules may be difficult by SDS-PAGE or isoelectric focusing alone. However, the two techniques can be combined in two-dimensional (2D) electrophoresis. In a typical experiment, a protein mixture is separated first by isoelectric focusing, yielding a pattern of bands in a gel slab such as the one shown in Fig. 8.18a. To improve the separation of closely spaced bands, the first slab is attached to a second slab and SDS-PAGE is performed with the electric field being applied in a direction that is perpendicular to the direction in which isoelectric focusing was performed. The macromolecules separate according to their molar masses along this second dimension of the experiment, and the result is that spots are spread widely over the surface of the slab, leading to enhanced separation of the mixture’s components (Fig. 8.18b). The techniques described so far give good separations, but the drift speeds attained by macromolecules in traditional electrophoresis methods are rather low; as a result, several hours are often necessary to achieve good separation of complex mixtures. According to eqn 8.22, one way to increase the drift speed is to increase the electric field strength. However, there are limits to this strategy because very large electric fields can heat the large surfaces of an electrophoresis apparatus unevenly, leading to a nonuniform distribution of electrophoretic mobilities and poor separation. In capillary electrophoresis, the sample is dispersed in a medium (such as methylcellulose) and held in a thin glass or plastic tube with diameters ranging from 20 to 100 mm. The small size of the apparatus makes it easy to dissipate heat when large electric fields are applied. Excellent separations may be achieved in minutes rather than hours.

The experimental steps taken during separation of a mixture of biopolymers by two-dimensional electrophoresis. (a) Isoelectric focusing is performed on a thin gel slab, resulting in separation along the vertical direction of the illustration. (b) The first slab is attached to a second, larger slab and SDS-PAGE is performed with the electric field oriented in the horizontal direction of the illustration, resulting in further separation by molar mass. The dashed horizontal lines show how the bands in the twodimensional gel correspond to the bands in the gel on which isoelectric focusing was performed. Fig. 8.18

294

8 COMPLEX BIOCHEMICAL PROCESSES

8.8 Transport across ion channels and ion pumps We now have enough background information about ion transport to consider the centrally important processes of ion transport mediated by ion channels and ion pumps, which are involved in the propagation of action potentials and the synthesis of ATP.

A representation of the patch clamp technique for the measurement of ionic currents through membranes in intact cells. (a) A section of membrane containing an ion channel is in tight contact with the tip of a micropipette containing an electrolyte solution and the patch electrode. (b) A schematic representation of the crosssection of a membrane-spanning K+ ion channel and (c) the protein. (d) The selectivity filter has a number of carbonyl groups that grip K+ ions. As explained in the text, electrostatic repulsions between two bound K+ ions encourage ionic movement through the selectivity filter and across the membrane. Fig. 8.19

The thermodynamic treatment of ion transport in Chapter 5 does not explain the fact that ion channels and pumps discriminate between ions. For example, it is found experimentally that a K+ ion channel is not permeable to Na+ ions. We shall see that the key to the selectivity of an ion channel or pump lies in the mechanism of transport and, consequently, in the structure of the protein and the size of the ion. The structures of a number of channel proteins have been obtained by the X-ray diffraction techniques that will be described in greater detail in Chapter 12. Information about the flow of ions across channels and pumps is supplied by the patch clamp technique. One of many possible experimental arrangements is shown in Fig. 8.19. With mild suction, a ‘patch’ of membrane from a whole cell or a small section of a broken cell can be attached tightly to the tip of a micropipette filled with an electrolyte solution and containing an electronic conductor, the patch electrode. A potential difference (the ‘clamp’) is applied between the patch electrode and an intracellular electronic conductor in contact with the cytosol of the cell. If the membrane is permeable to ions at the applied potential difference, a current flows through the completed circuit. Using narrow micropipette tips with diameters of less than 1 mm, ion currents of a few picoamperes (1 pA = 10−12 A) have been measured across sections of membranes containing only one ion channel protein. (a) The potassium channel

A detailed picture of the mechanism of action of ion channels has emerged from analysis of patch clamp data and structural data. Here we focus on the K+ ion

8.8 TRANSPORT ACROSS ION CHANNELS AND ION PUMPS

channel protein, which, like all other mediators of ion transport, spans the membrane bilayer. The pore through which ions move has a length of 3.4 nm and is divided into two regions: a wide region with a length of 2.2 nm and diameter of 1.0 nm, and a narrow region with a length of 1.2 nm and diameter of 0.3 nm. The narrow region is called the selectivity filter of the K+ ion channel because it allows only K+ ions to pass. Filtering is a subtle process that depends on ionic size and the thermodynamic tendency of an ion to lose its hydrating water molecules. On entering the selectivity filter, the K+ ion is stripped of its hydrating shell and is then gripped by carbonyl groups of the protein. Dehydration of the K+ ion is endergonic (DdehydG 3 = +203 kJ mol−1) but is driven by the energy of interaction between the ion and the protein. The Na+ ion, although smaller than the K+ ion, does not pass through the selectivity filter of the K+ ion channel because interactions with the protein are not sufficient to compensate for the high Gibbs energy of dehydration of Na+ (DdehydG 3 = +301 kJ mol−1). More specifically, a dehydrated Na+ ion is too small and cannot be held tightly by the protein carbonyl groups, which are positioned for ideal interactions with the larger K+ ion. In its hydrated form, the Na+ ion is too large (larger than a dehydrated K+ ion), does not fit in the selectivity filter, and does not cross the membrane. Although very selective, a K+ ion channel can still let other ions pass through. For example, K+ and Tl+ ions have similar radii and Gibbs energies of dehydration, so Tl+ can cross the membrane. As a result, Tl+ is a neurotoxin because it replaces K+ in many neuronal functions and suppresses them. The efficiency of transfer of K+ ions through the channel can also be explained by structural features of the protein. For efficient transport to occur, a K+ ion must enter the protein but then must not be allowed to remain inside for very long, so that as one K+ ion enters the channel from one side, another K+ ion leaves from the opposite side. An ion is lured into the channel by water molecules about halfway through the length of the membrane. Consequently, the thermodynamic cost of moving an ion from an aqueous environment to the less hydrophilic interior of the protein is minimized. The ion is encouraged to leave the protein by electrostatic interactions in the selectivity filter, which can bind two K+ ions simultaneously, usually with a bridging water molecule. Electrostatic repulsion prevents the ions from binding too tightly, minimizing the residence time of an ion in the selectivity filter and maximizing the transport rate. (b) The proton pump

Now we turn our attention to a very important ion pump, the H+-ATPase responsible for coupling of proton flow to synthesis of ATP from ADP and Pi (Chapter 4). Structural studies show that the channel through which the protons flow is linked in tandem to a unit composed of six protein molecules arranged in pairs of a and b subunits to form three interlocked ab segments (Fig. 8.20). The conformations of the three pairs may be loose, (L), tight (T), or open (O), and one of each type is present at each stage. A protein at the centre of the interlocked structure, the subunit shown as an arrow, rotates and induces structural changes that cycle each of the three segments between L, T, and O conformations. At the start of a cycle, a T unit holds an ATP molecule. Then ADP and a Pi group migrate into the L site, and as it closes into T, the earlier T site opens into O and releases its ATP. The ADP and Pi in the T site meanwhile condense into ATP,

295

296

8 COMPLEX BIOCHEMICAL PROCESSES

Fig. 8.20 The mechanism of action of H+-ATPase, a molecular motor that transports protons across the mitochondrial membrane and catalyzes either the formation or hydrolysis of ATP. The yellow shapes represent the species ADP, ATP, and Pi.

and the new L site is ready for the cycle to begin again. The proton flux drives the rotation of a g subunit, and hence the conformational changes of the ab segments, as well as providing the energy for the condensation reaction itself. Several key aspects of this mechanism have been confirmed experimentally. For example, the rotation of a g subunit has been portrayed directly by using single-molecule spectroscopy (In the laboratory 12.6).

Electron transfer in biological systems We saw in Case studies 4.2 and 4.3 that exergonic electron transfer processes drive the synthesis of ATP in the mitochondrion during oxidative phosphorylation. Electron transfer between protein-bound co-factors or between proteins also plays a role in other biological processes, such as photosynthesis (Section 5.11 and Case study 12.3), nitrogen fixation, the reduction of atmospheric N2 to NH3 by certain microorganisms, and the mechanisms of action of oxidoreductases, which are enzymes that catalyze redox reactions. We begin by examining the features of a theory that describes the factors governing the rates of electron transfer. Then we discuss the theory in the light of experimental results on a variety of systems, including protein complexes. We shall see that relatively simple expressions can be used to predict the rates of electron transfer between proteins with reasonable accuracy. 8.9 The rates of electron transfer processes Electron transfer is of crucial importance in many biological reactions, and we need to see how to use the strategies we have developed to discuss them quantitatively.

Consider electron transfer from a donor species D to an acceptor species A in solution. The net reaction, the observed rate law, and the equilibrium constant are D + A → D+ + A−

v = kobs[D][A]

K=

[D+][A−] [D][A]

Electron transfer

(8.24)

8.9 THE RATES OF ELECTRON TRANSFER PROCESSES

The proposed mechanism is: Electron transfer D + A 7 DA

ka, k a′

DA → D+A−

vet = ket[DA]

D A → DA

vret = k′et[D+A−]

D+A− → D+ + A−

vd = kd[D+A−]

+

KDA =

kac 3 [DA]c 3 = k a′ [D][A]

In the first step of the mechanism, D and A must diffuse through the solution and encounter to form a complex DA, in which the donor and acceptor are separated by a distance comparable to r, the distance between the edges of each species. Next, electron transfer occurs within the DA complex to yield D+A−. The D+A− complex has two possible fates. One is the regeneration of DA. The other is to break apart and for the ions to diffuse through the solution. We show in the following Justification that kobs in eqn 8.24 is given by 1 1 k′ A k′ D = + a 1 + et kobs k a kaket C kd F

Electron transfer rate constant

(8.25)

Justification 8.4 The rate constant for electron transfer in solution

We begin by equating the rate of the net reaction (eqn 8.24) to the rate of formation of separated ions, the reaction products: v = kobs[D][A] = kd[D+A−] Next, we apply the steady-state assumption to the intermediate D+A−: d[D+A−] = ket[DA] − k′et[D+A−] − kd[D+A−] = 0 dt It follows that [D+A−] =

ket [DA] k′et + kd

However, DA is also an intermediate, so we apply the steady-state approximation again: d[DA] = ka[D][A] − k a′[DA] − ket[DA] + k′et[D+A−] = 0 dt Substitution of the initial expression for the steady-state concentration of D+A− into this expression for [DA] gives, after some algebra, a new expression for [D+A−]: [D+A−] =

kaket [D][A] k a′k′et + k a′kd + kd ket

When we multiply this expression by kd, we see that the resulting equation has the form of the rate of electron transfer, v = kobs[D][A], with kobs given by kobs =

kdka ket k a′k′et + k a′kd + kdket

To obtain eqn 8.25, we divide the numerator and denominator on the righthand side of this expression by kdket and solve for the reciprocal of kobs.

297

298

8 COMPLEX BIOCHEMICAL PROCESSES

To gain insight into eqn 8.25 and the factors that determine the rate of electron transfer reactions in solution, we assume that the main decay route for D+A− is dissociation of the complex into separated ions, or kd >> k′et. It then follows that 1 1 A k′ D ≈ 1+ a kobs ka C ket F

(8.26)

There are two limits to consider: • When ket >> k a′, kobs ≈ ka and the rate of product formation is controlled by diffusion of D and A in solution, which fosters formation of the DA complex. • When, ket > | DrG 3 |, kobs may be estimated by a special case of the Marcus cross-relation: kobs = (kDDkAAK)1/2

Marcus cross-relation

(8.34)

where K is the equilibrium constant for the net electron transfer reaction (eqn 8.24) and kDD and kAA (in general, kii ) are the experimental rate constants for the electron self-exchange processes (with the colors distinguishing one molecule from another): D + D+ → D+ + D A− + A → A + A−

kDD kAA

Justification 8.5 The Marcus cross-relation

To derive the Marcus cross-relation, we use eqn 8.33 to write the rate constants for the self-exchange reactions as kDD = ZDD e−D G ‡

DD

kAA = ZAA e−D G ‡

/RT

AA

/RT

For the net reaction and the self-exchange reactions, the Gibbs energy of activation may be written from eqn 8.29 as D‡G =

D rG 32 1 + 2 D rG 3 + 14 l 4l

3 3 = DrG AA = 0 and hence D‡GDD = 14 lDD and For the self-exchange reactions D rG DD D‡GAA = 14 lAA. It follows that

kDD = ZDD e−l

kAA = ZAA e−l

/4RT

DD

/4RT

AA

To make further progress, Marcus assumed that the reorganization energy of the net reaction is the arithmetic mean of the reorganization energies of the self-exchange reactions: l = 12 (lDD + lAA) Provided l >> DrG 3 for the net reaction, the first term in D‡G may be neglected and the Gibbs energy of activation of the net reaction is D‡G = 12 DrG 3 + 18 lDD + 18 lAA Therefore, the rate constant for the net reaction is 3

kobs = Ze−D G /2RT e−l r

/8RT −lAA/8RT

DD

e

3

We can use eqn 4.10 (ln K = −DrG 3/RT) in the form K = e−D G /RT to write r

kobs = (kDDkAAK)1/2 f where f=

Z (ZAAZDD)1/2

In practice, the factor f is usually set to 1 and we obtain eqn 8.34. The rate constants estimated by eqn 8.34 agree fairly well with experimental rate constants for electron transfer between proteins, as we see in the following example.

301

302

8 COMPLEX BIOCHEMICAL PROCESSES

Example 8.4

Using the Marcus cross-relation

The following data were obtained for cytochrome c and cytochrome c551, two proteins in which heme-bound iron ions shuttle between the oxidation states Fe(II) and Fe(III): cytochrome c cytochrome c551

kii /(dm3 mol−1 s−1) 1.5 × 102 4.6 × 107

E 9/V +0.260 +0.286

Estimate the rate constant kobs for the process cytochrome c551(red) + cytochrome c(ox) → cytochrome c551(ox) + cytochrome c(red) Then compare the estimated value with the observed value of 6.7 × 104 dm3 mol−1 s−1. Strategy We use the standard potentials and eqns 5.16 (ln K = nFE 3cell /RT )

and 5.17a (E 3cell = E 3R − E 3L ) to calculate the equilibrium constant K. Then we use eqn 8.34, the calculated value of K, and the self-exchange rate constants k ii to calculate the rate constant kobs. Solution The two reduction half-reactions are

Right: cytochrome c(ox) + e− → cytochrome c(red)

E 3R = +0.260 V

Left: cytochrome c551(ox) + e− → cytochrome c551(red)

E 3L = +0.286 V

The difference is E 3cell = (0.260 V) − (0.286 V) = −0.026 V It then follows from eqn 5.16 with v = 1 and RT/F = 25.69 mV that ln K = −

0.026V 2.6 =− 25.69 × 10−3 V 2.569

Therefore, K = 0.36. From eqn 8.34 and the self-exchange rate constants, we calculate kobs = {(1.5 × 102 dm3 mol−1 s−1) × (4.6 × 107 dm3 mol−1 s−1) × 0.36}1/2 = 5.0 × 104 dm3 mol−1 s−1 The calculated and observed values differ by only 25 per cent, indicating that the Marcus relation can lead to reasonable estimates of rate constants for electron transfer.

Self-test 8.6 Estimate kobs for the reduction by cytochrome c of plastocyanin, a protein containing a copper ion that shuttles between the +2 and +1 oxidation states and for which kAA = 6.6 × 102 dm3 mol−1 s−1 and E 3cell = +0.350 V.

Answer: 1.8 × 103 dm3 mol−1 s−1

CHECKLIST OF KEY EQUATIONS

303

Checklist of key concepts 1. Catalysts are substances that accelerate reactions but undergo no net chemical change. 2. A hom*ogeneous catalyst is a catalyst in the same phase as the reaction mixture. 3. Enzymes are hom*ogeneous, biological catalysts. 4. The Michaelis–Menten mechanism of enzyme kinetics accounts for the dependence of rate on the concentration of the substrate. 5. A Lineweaver–Burk plot is used to determine the parameters that occur in the Michaelis–Menten mechanism. 6. In sequential reactions, the active site binds all the substrates before processing them into products. In ‘ping-pong’ reactions, products are released in a stepwise fashion. 7. In competitive inhibition of an enzyme, the inhibitor binds only to the active site of the enzyme and thereby inhibits the attachment of the substrate. 8. In uncompetitive inhibition, the inhibitor binds to a site of the enzyme that is removed from the active site but only if the substrate is already present. 9. In noncompetitive inhibition, the inhibitor binds to a site other than the active site, and its presence reduces the ability of the substrate to bind to the active site.

10. Fick’s first law of diffusion states that the flux of molecules is proportional to the concentration gradient. 11. Fick’s second law of diffusion (the diffusion equation) states that the rate of change of concentration in a region is proportional to the curvature of the concentration in the region. 12. Diffusion is an activated process. 13. The flux of molecules through biological membranes is often mediated by carrier molecules. 14. Protons migrate by the Grotthus mechanism, Fig. 8.16. 15. Electrophoresis is the motion of a charged macromolecule, such as DNA, in response to an electric field. Important techniques are gel electrophoresis, isoelectric focusing, pulsed-field electrophoresis, two-dimensional electrophoresis, and capillary electrophoresis. 16. According to the Marcus theory, the rate constant of electron transfer in a donor–acceptor complex depends on the distance between electron donor and acceptor, the standard reaction Gibbs energy, and the reorganization energy, l.

Checklist of key equations Property

Equation

Comment

Michaelis–Menten rate law

v = vmax[S]0 /([S]0 + KM)

vmax = k b[E]0; assumes S is in excess

Lineweaver–Burk plot

1/v = 1/vmax + (KM/vmax)(1/[S]0)

Based on Michaelis–Menten mechanism

Fick’s first law

J = −Ddc/dx

Fick’s second law

∂c/∂t = D∂ 2c/∂x 2

Temperature-dependence of D

D = D0e

Mobility of an ion

u = ez/6pha

Marcus expression

ket ∝ e−bre−D G/RT ‡

with D‡G = (DrG 3 + l)2/4l Marcus cross-relation

Also known as the diffusion equation

−E a /RT

k obs = (kDDkAAK )1/2

Assumes the validity of Stokes’ law

304

8 COMPLEX BIOCHEMICAL PROCESSES

Further information Further information 8.1 Fick’s laws of diffusion 1. Fick’s first law of diffusion

Consider the arrangement in Fig. 8.22. In an interval Dt the number of molecules passing through the window of area A from the left is proportional to the number in the slab of thickness l and area A, and therefore volume lA, just to the left of the window where the average (number) concentration is c(x − 12 l), and to the length of the interval Dt: number coming from left ∝ c(x − 12 l)lADt

On writing the constant of proportionality as D (and absorbing l 2 into it), we obtain eqn 8.11. 2. Fick’s second law

Consider the arrangement in Fig. 8.23. The number of solute particles passing through the window of area A located at x in an infinitesimal interval dt is J(x)Adt, where J(x) is the flux at the location x. The number of particles passing out of the region through a window of area A at x + dx is J(x + dx)Adt, where J(x + dx) is the flux at the location of this window. The flux in and the flux out will be different if the concentration gradients are different at the two windows. The net change in the number of solute particles in the region between the two windows is net change in number = J(x)Adt − J(x + dx)Adt = {J(x) − J(x + dx)}Adt

The calculation of the rate of diffusion considers the net flux of molecules through a plane of area A as a result of arrivals from on average a distance 12 l in each direction.

Fig. 8.22

Likewise, the number coming from the right in the same interval is number coming from right ∝ c(x + l)lADt 1 2

The net flux is therefore proportional to the difference in these numbers divided by the area and the time interval: J∝

c(x − 12 l)lADt − c(x + 12 l)lADt = {c(x − 12 l) − c(x + 12 l)}l ADt

We now express the two concentrations in terms of the concentration at the window itself, c(x), as follows: dc dx dc c(x − 12 l) = c(x) − 12 l × dx

c(x + 12 l) = c(x) + 12 l ×

From which it follows that dc D A dc D # !A J ∝ c(x) − 12 l l − c(x) + 12 l F C @C dx dxF $ ∝ −l 2

dc dx

To calculate the change in concentration in the region between the two walls, we need to consider the net effect of the influx of particles from the left and their efflux toward the right. Only if the slope of the concentrations is different at the two walls will there be a net change.

Fig. 8.23

Now we express the flux at x + dx in terms of the flux at x and the gradient of the flux, dJ/dx: J(x + dx) = J(x) +

dJ × dx dx

It follows that net change in number = −

dJ × dx Adt dx

The change in concentration inside the region between the two windows is the net change in number divided by the volume of the region (which is Adx), and the net rate of change is obtained by dividing that change in concentration by the time interval dt. Therefore, on dividing by both Adx and dt, we obtain

305

EXERCISES

dJ A dc D rate of change of concentration = =− C dt F dx Finally, we express the flux by using Fick’s first law (at this point we need to acknowledge that c depends on both x and t and therefore use partial differential notation):

∂c ∂ A ∂c D ∂ 2c =− −D =D 2 ∂t ∂x C ∂x F ∂x which is eqn 8.12.

Discussion questions 8.1 Discuss the features and limitations of the Michaelis–Menten mechanism of enzyme action. 8.2 Prepare a report on the application of the experimental strategies described in Chapters 6 and 7 to the study of enzyme-catalyzed reactions. Devote some attention to the following topics: (a) the determination of reaction rates over a long time scale, (b) the determination of the rate constants and equilibrium constant of binding of substrate to an enzyme, and (c) the characterization of intermediates in a catalytic cycle. Your report should be similar in content and extent to one of the Case studies found throughout this text. 8.3 A plot of the rate of an enzyme-catalyzed reaction against temperature has a maximum, in an apparent deviation from the behavior predicted by the Arrhenius relation (eqn 6.19). Provide a molecular interpretation for this effect. 8.4 Describe graphical procedures for distinguishing between

(a) sequential and ping-pong enzyme-catalyzed reactions and (b) competitive, uncompetitive, and noncompetitive inhibition of an enzyme.

8.5 Some enzymes are inhibited by high concentrations of their own products. (a) Sketch a plot of reaction rate against concentration of substrate for an enzyme that is prone to product inhibition. (b) How does product inhibition of hexokinase, the enzyme that phosphorylates glucose in the first step of glycolysis, provide a mechanism for regulation of glycolysis in the cell? Hint: Review Case study 4.3. 8.6 Provide a molecular interpretation for the observation that mediated transport through biological membranes leads to a maximum flux Jmax when the concentration of the transported species becomes very large. 8.7 Discuss the mechanism of proton conduction in liquid water. For a more detailed account of the modern version of this mechanism, consult our Quanta, matter, and change (2009). 8.8 Discuss how the following factors determine the rate of electron transfer in biological systems: (a) the distance between electron donor and acceptor, and (b) the reorganization energy of redox active species and the surrounding medium.

Exercises 8.9 As remarked in the text, Michaelis and Menten derived their rate law by assuming a rapid pre-equilibrium of E, S, and ES. Derive the rate law in this manner, and identify the conditions under which it becomes the same as that based on the steady-state approximation (eqn 8.1). 8.10 Equation 8.4a gives the expression for the rate of formation of

product by a modified version of the Michaelis–Menten mechanism in which the second step is also reversible. Derive the expression and find its limiting behavior for large and small concentrations of substrate. 8.11 For many enzymes, such as chymotrypsin (Case study 8.1), the

mechanism of action involves the formation of two intermediates: E + S → ES

v = ka[E][S]

ES → E + S

v = ka′[ES]

ES → ES′

v = kb[ES]

ES′ → E + P

v = kc[ES′]

Show that the rate of formation of product has the same form as that shown in eqn 8.1, written as: v=

vmax 1 + KM/[S]0

but with vmax and KM given by vmax =

kbkc[E]0 kb + kc

and

KM =

kc(ka′ + kb) ka(kb + kc)

8.12 The enzyme-catalyzed conversion of a substrate at 25°C has a Michaelis constant of 0.045 mol dm−3. The rate of the reaction is 1.15 mmol dm−3 s−1 when the substrate concentration is 0.110 mol dm−3. What is the maximum velocity of this reaction? 8.13 Find the condition for which the reaction rate of an enzyme-

catalyzed reaction that follows Michaelis–Menten kinetics is half its maximum value. 8.14 Isocitrate lyase catalyzes the following reaction:

isocitrate ion → glyoxylate ion + succinate ion The rate, v, of the reaction was measured when various concentrations of isocitrate ion were present, and the following results were obtained at 25°C: [isocitrate]/(mmol dm−3) V/(pmol dm−3 s−1)

31.8 70.0

46.4 97.2

59.3 116.7

118.5 159.2

222.2 194.5

Determine the Michaelis constant and the maximum velocity of the reaction.

306

8 COMPLEX BIOCHEMICAL PROCESSES

8.15 The following results were obtained for the action of an ATPase on ATP at 20°C, when the concentration of the ATPase was 20 nmol dm−3:

[ATP]/(mmol dm−3) v/(mmol dm−3 s−1)

0.60 0.81

0.80 0.97

1.4 1.30

2.0 1.47

3.0 1.69

Determine the Michaelis constant, the maximum velocity of the reaction, the turnover number, and the catalytic efficiency of the enzyme. 8.16 Enzyme-catalyzed reactions are sometimes analyzed by use of

the Eadie–Hofstee plot, in which v/[S]0 is plotted against v. (a) Using the simple Michaelis–Menten mechanism, derive a relation between v/[S]0 and v. (b) Discuss how the values of KM and vmax are obtained from analysis of the Eadie–Hofstee plot. (c) Determine the Michaelis constant and the maximum velocity of the reaction from Exercise 8.14 by using an Eadie–Hofstee plot to analyze the data. 8.17 Enzyme-catalyzed reactions are sometimes analyzed by use of

the Hanes plot, in which [S]0/v is plotted against [S]0. (a) Using the simple Michaelis–Menten mechanism, derive a relation between [S]0/v and [S]0. (b) Discuss how the values of KM and vmax are obtained from analysis of the Hanes plot. (c) Determine the Michaelis constant and the maximum velocity of the reaction from Exercise 8.14 by using a Hanes plot to analyze the data. 8.18 An allosteric enzyme shows catalytic activity that changes on

noncovalent binding of small molecules called effectors. For example, consider a protein enzyme consisting of several identical subunits and several active sites. In one mode of allosteric behavior, the substrate acts as effector, so that binding of a substrate molecule to one of the subunits either increases or decreases the catalytic efficiency of the other active sites. Consequently, reactions catalyzed by allosteric enzymes show significant deviations from Michaelis–Menten behaviour. (a) Sketch a plot of reaction rate against substrate concentration for a multi-subunit allosteric enzyme, assuming that the catalytic efficiency changes in such a way that the enzyme with all its active sites occupied is more efficient than the enzyme with one fewer bound substrate molecule, and so on. Compare your sketch with Fig. 8.2, which illustrates Michaelis–Menten behavior. (b) Your plot from part (a) should have a sigmoidal shape (S shape) that is typical for allosteric enzymes. The mechanism of the reaction can be written as E + nS 7 ESn → E + nP and the reaction rate v is given by vmax v= 1 + K′/[S]n0 where K′ is a collection of rate constants analogous to the Michaelis constant and n is the interaction coefficient, which may be taken as the number of active sites that interact to give allosteric behavior. Plot v/vmax against [S]0 for a fixed value of K′ of your choosing and several values of n. Confirm that the expression for v does predict sigmoidal kinetics and provide a molecular interpretation for the effect of n on the shape of the curve. 8.19 (a) Show that the expression for the rate of a reaction catalyzed

by an allosteric enzyme of the type discussed in Exercise 8.18 may be rewritten as log

v vmax − v

= n log [S]0 − log K′

(b) Use the preceding expression and the following data to determine the interaction coefficient for an enzyme-catalyzed reaction showing sigmoidal kinetics:

[S]0 /(10−5 mol dm−3) V/(mmol dm−3 s−1)

0.10 0.0040

0.40 0.25

0.50 0.46

[S]0/(10−5 mol dm−3) V/(mmol dm−3 s−1)

0.60 0.75

0.80 1.42

1.0 2.08

[S]0/(10−5 mol dm−3) V/(mmol dm−3 s−1)

1.5 3.22

2.0 3.70

3.0 4.02

For substrate concentrations ranging between 0.10 mmol dm−3 and 10 mmol dm−3, the reaction rate remained constant at 4.17 mmol dm−3 s−1. 8.20 A simple method for the determination of the interaction

coefficient n for an enzyme-catalyzed reaction involves the calculation of the ratio [S]90 /[S]10, where [S]90 and [S]10 are the concentrations of substrate for which the reaction rates are 0.90vmax and 0.10vmax, respectively. (a) Show that [S]90 /[S]10 = 81 for an enzyme-catalyzed reaction that follows Michaelis–Menten kinetics. (b) Show that [S]90 /[S]10 = (81)1/n for an enzyme-catalyzed reaction that follows sigmoidal kinetics, where n is the interaction coefficient defined in Exercise 8.19. (c) Use the data from Exercise 8.19 to estimate the value of n. 8.21 Yeast alcohol dehydrogenase catalyzes the oxidation of ethanol

by NAD+ according to the reaction CH3CH2OH(aq) + NAD+(aq) → CH3CHO(aq) + NADH(aq) + H+(aq) The following results were obtained for the reaction: [CH3CH2OH]0 /(10−2 mol dm−3) V/(mol s−1 (kg protein)−1) V/(mol s−1 (kg protein)−1) V/(mol s−1 (kg protein)−1) V/(mol s−1 (kg protein)−1)

(a) (b) (c) (d)

1.0 0.30 0.51 0.89 1.43

2.0 0.44 0.75 1.32 2.11

4.0 0.57 0.99 1.72 2.76

20.0 0.76 1.31 2.29 3.67

where the concentrations of NAD+ are (a) 0.050 mmol dm−3, (b) 0.10 mmol dm−3, (c) 0.25 mmol dm−3, and (d) 1.0 mmol dm−3. Is the reaction sequential or ping-pong? Determine vmax and the appropriate K constants for the reaction. 8.22 One of the key events in the transmission of chemical messages

in the brain is the hydrolysis of the neurotransmitter acetylcholine by the enzyme acetylcholinesterase. The kinetic parameters for this reaction are kcat = 1.4 × 104 s−1 and KM = 9.0 × 10−5 mol dm−3. Is acetylcholinesterase catalytically perfect? 8.23 The enzyme carboxypeptidase catalyses the hydrolysis of polypeptides, and here we consider its inhibition. The following results were obtained when the rate of the enzymolysis of carbobenzoxy-glycyl-d-phenylalanine (CBGP) was monitored without inhibitor:

[CBGP]0 /(10−2 mol dm−3) Relative reaction rate

1.25 0.398

3.84 0.669

5.81 0.859

7.13 1.000

(All rates in this exercise were measured with the same concentration of enzyme and are relative to the rate measured when [CBGP]0 = 0.0713 mol dm−3 in the absence of inhibitor.) When 2.0 mmol dm−3 phenylbutyrate ion was added to a solution containing the enzyme and substrate, the following results were obtained: [CBGP]0 /(10−2 mol dm−3) Relative reaction rate

1.25 0.172

2.50 0.301

4.00 0.344

5.50 0.548

In a separate experiment, the effect of 50 mmol dm−3 benzoate ion was monitored and the results were [CBGP]0/(10−2 mol dm−3) Relative reaction rate

1.75 0.183

2.50 0.201

5.00 0.231

10.00 0.246

EXERCISES

Determine the mode of inhibition of carboxypeptidase by the phenylbutyrate ion and benzoate ion. 8.24 Consider an enzyme-catalyzed reaction that follows Michaelis–

307

8.34 The viscosity of water at 20°C is 1.0019 × 10−3 kg m−1 s−1 and at

30°C it is 7.982 × 10−4 kg m−1 s−1. What is the activation energy for the motion of water molecules?

Menten kinetics with KM = 3.0 mmol dm−3. What concentration of a competitive inhibitor characterized by KI = 20 mmol dm−3 will reduce the rate of formation of product by 50 per cent when the substrate concentration is held at 0.10 mmol dm−3?

8.35 The mobility of a Na+ ion in aqueous solution is 5.19 × 10−8 m2 s−1 V−1 at 25°C. The potential difference between two electrodes placed in the solution is 12.0 V. If the electrodes are 1.00 cm apart, what is the drift speed of the ion? Use h = 8.91 × 10−4 kg m−1 s−1.

8.25 Some enzymes are inhibited by high concentrations of their own

8.36 It is possible to estimate the isoelectric point of a protein from

substrates. (a) Show that when substrate inhibition is important, the reaction rate v is given by

its primary sequence. (a) A molecule of calf thymus histone contains one aspartic acid, one glutamic acid, 11 lysine, 15 arginine, and two histidine residues. Will the protein bear a net charge at pH = 7? If so, will the net charge be positive or negative? Is the isoelectric point of the protein less than, equal to, or greater than 7? Hint: See Exercise 4.45. (b) Each molecule of egg albumin has 51 acidic residues (aspartic and glutamic acid), 15 arginine, 20 lysine, and seven histidine residues. Is the isoelectric point of the protein less than, equal to, or greater than 7? (c) Can a mixture of calf thymus histone and egg albumin be separated by gel electrophoresis with the isoelectric focusing method?

v=

vmax 1 + KM/[S]0 + [S]0/KI

where KI is the equilibrium constant for dissociation of the inhibited enzyme–substrate complex. (b) What effect does substrate inhibition have on a plot of 1/v against 1/[S]0? 8.26 What is (a) the flux of nutrient molecules down a concentration gradient of 0.10 mol dm−3 m−1, (b) the amount of molecules (in moles) passing through an area of 5.0 mm2 in 1.0 min? Take for the diffusion coefficient the value for sucrose in water (5.22 × 10−10 m2 s−1). 8.27 How long does it take a sucrose molecule in water at 25°C to

8.37 We saw in Section 8.8 that to pass through a channel, the ion

must first lose its hydrating water molecules. To explore the motion of hydrated Na+ ions, we need to know that the diffusion coefficient D of an ion is related to its mobility u by the Einstein relation:

diffuse (a) 1 mm, (b) 1 cm, and (c) 1 m from its starting point? 8.28 The mobility of species through fluids is of the greatest

importance for nutritional processes. (a) Estimate the diffusion coefficient for a molecule that steps 150 pm each 1.8 ps. (b) What would be the diffusion coefficient if the molecule traveled only half as far on each step? 8.29 The diffusion coefficient of a particular kind of t-RNA molecule is D = 1.0 × 10−11 m2 s−1 in the medium of a cell interior at 37°C. How long does it take molecules produced in the cell nucleus to reach the walls of the cell at a distance 1.0 mm, corresponding to the radius of the cell? 8.30 The diffusion coefficients for a lipid in a plasma membrane and in a lipid bilayer are 1.0 × 10−10 m2 s−1 and 1.0 × 10−9 m2 s−1, respectively. How long will it take the lipid to diffuse 10 nm in a plasma membrane and a lipid bilayer? 8.31 Diffusion coefficients of proteins are often used as a measure of molar mass. For a spherical protein, D ∝ M−1/2. Considering only one-dimensional diffusion, compare the length of time it would take ribonuclease (M = 13.683 kg mol−1) to diffuse 10 nm to the length of time it would take the enzyme catalase (M = 250 kg mol−1) to diffuse the same distance. 8.32 Is diffusion important in lakes? How long would it take a small

pollutant molecule about the size of H2O to diffuse across a lake of width 100 m? 8.33 Pollutants spread through the environment by convection

(winds and currents) and by diffusion. How many steps must a molecule take to be 1000 step lengths away from its origin if it undergoes a one-dimensional random walk?

D=

uRT zF

where z is the charge number of the ion and F is Faraday’s constant. (a) Estimate the diffusion coefficient and the effective hydrodynamic radius a of the Na+ ion in water at 25°C. For water, h = 8.91 × 10−4 kg m−1 s−1. (b) Estimate the approximate number of water molecules that are dragged along by the cations. Ionic radii are given in Table 9.3. 8.38 For a pair of electron donor and acceptor, ket = 2.02 × 105 s−1 for DrG 3 = −0.665 eV. The standard reaction Gibbs energy changes to DrG 3 = −0.975 eV when a substituent is added to the electron acceptor and the rate constant for electron transfer changes to ket = 3.33 × 106 s−1. Assuming that the distance between donor and acceptor is the same in both experiments, estimate the value of the reorganization energy. 8.39 For a pair of electron donor and acceptor, ket = 2.02 × 105 s−1 when r = 1.11 nm and ket = 2.8 × 104 s−1 when r = 1.23 nm. (a) Assuming that DrG 3 and l are the same in both experiments, estimate the value of b. (b) Estimate the value of ket when r = 1.48 nm. 8.40 Azurin is a protein containing a copper ion that shuttles between

the +2 and +1 oxidation states, and cytochrome c is a protein in which a heme-bound iron ion shuttles between the +3 and +2 oxidation states. The rate constant for electron transfer from reduced azurin to oxidized cytochrome c is 1.6 × 103 dm3 mol−1 s−1. Estimate the electron self-exchange rate constant for azurin from the following data: Cytochrome c Azurin

kii /(dm3 mol−1 s−1) 1.5 × 102 ?

9 E cell /V 0.260 0.304

308

8 COMPLEX BIOCHEMICAL PROCESSES

Projects 8.41 Autocatalysis is the catalysis of a reaction by the products.

For example, for a reaction A → P it can be found that the rate law is v = k[A][P] and the reaction rate is proportional to the concentration of P. The reaction gets started because there are usually other reaction routes for the formation of some P initially, which then takes part in the autocatalytic reaction proper. Many biological and biochemical processes involve autocatalytic steps, and here we explore one case: the spread of infectious diseases. (a) Integrate the rate equation for an autocatalytic reaction of the form A → P, with rate law v = k[A][P], and show that

KES,a =

ESH+2 7 ESH + H+ KES,b =

[ES−][H+] [ESH] [ESH][H+] [ESH +2]

in which only the EH and ESH forms are active. (a) For the mechanism above, show that v=

v′max 1 + K M′ /[S]0

with

[P] e at = (1 + b) [P]0 1 + be at where a = ([A]0 + [P]0)k and b = [P]0/[A]0. Hint: Starting with the expression v = −d[A]/dt = k[A][P], write [A] = [A]0 − x, [P] = [P]0 + x and then write the expression for the rate of change of either species in terms of x. To integrate the resulting expression, the following relation will be useful: 1 1 1 D A 1 = + ([A]0 − x)([P]0 + x) [A]0 + [P]0 C [A]0 − x [P]0 + x F (b) Plot [P]/[P]0 against at for several values of b. Discuss the effect of autocatalysis on the shape of a plot of [P]/[P]0 against t by comparing your results with those for a first-order process, in which [P]/[P]0 = 1 − e−kt. (c) Show that for the autocatalytic process discussed in parts (a) and (b), the reaction rate reaches a maximum at tmax = −(1/a) ln b. (d) In the so-called SIR model of the spread and decline of infectious diseases, the population is divided into three classes: the susceptibles, S, who can catch the disease, the infectives, I, who have the disease and can transmit it, and the removed class, R, who have either had the disease and recovered, are dead, are immune, or are isolated. The model mechanism for this process implies the following rate laws: dS = −rSI dt

ESH 7 ES− + H+

dI = rSI − aI dt

dR = aI dt

(i) What are the autocatalytic steps of this mechanism? (ii) Find the conditions on the ratio a/r that decide whether the disease will spread (an epidemic) or die out. (iii) Show that a constant population is built into this system, namely that S + I + R = N, meaning that the timescales of births, deaths by other causes, and migration are assumed large compared to that of the spread of the disease. 8.42 In general, the catalytic efficiency of an enzyme depends on the

pH of the medium in which it operates. One way to account for this behavior is to propose that the enzyme and the enzyme–substrate complex are active only in specific protonation states. This proposition can be summarized by the following mechanism: EH + S 8 ESH

ka, ka′

ESH → E + P

kb

EH 7 E− + H+

KE,a =

[E−][H+] [EH]

EH+2 7 EH + H+

KE,b =

[EH][H+] [EH +2]

v′max =

vmax [H+] KES,a + 1+ KES,b [H+]

[H+] KE,a + KE,b [H+] K M′ = KM [H+] KES,a + 1+ KES,b [H+] 1+

where vmax and KM correspond to the form EH of the enzyme. (b) For pH values ranging from 0 to 14, plot v′max against pH for a hypothetical reaction for which vmax = 1.0 mmol dm−3 s−1, KES,b = 1.0 mmol dm−3, and KES,a = 10 nmol dm−3. Is there a pH at which vmax reaches a maximum value? If so, determine the pH. (c) Redraw the plot in part (b) by using the same value of vmax, but KES,b = 0.10 mmol dm−3 and KES,a = 0.10 nmol dm−3. Account for any differences between this plot and the plot from part (b). 8.43 Studies of biochemical reactions initiated by the absorption

of light have contributed significantly to our understanding of the kinetics of electron transfer processes. The experimental arrangement is that for time-resolved spectroscopy (In the laboratory 7.2) and relies on the observation that many substances become more efficient electron donors on absorbing energy from a light source, such as a laser. With judicious choice of electron acceptor, it is possible to set up an experimental system in which electron transfer will not occur in the dark (when only a poor electron donor is present) but will proceed after application of a laser pulse (when a better electron donor is generated). Nature makes use of this strategy to initiate the chain of electron transfer events that leads ultimately to the phosphorylation of ATP in photosynthetic organisms. (a) An elegant way to study electron transfer in proteins consists of attaching an electroactive species to the protein’s surface and then measuring ket between the attached species and an electroactive protein cofactor. J.W. Winkler and H.B. Gray, Chem. Rev. 92, 369 (1992), summarize data for cytochrome c modified by replacement of the heme iron by a Zn2+ ion, resulting in a zinc–porphyrin (ZnP) moiety in the interior of the protein, and by attachment of a ruthenium ion complex to a surface histidine amino acid. The edge-to-edge distance between the electroactive species was thus fixed at 1.23 nm. A variety of ruthenium ion complexes with different standard reduction potentials were used. For each rutheniummodified protein, either Ru2+ → ZnP+ or ZnP* → Ru3+, in which the zinc-porphyrin is excited by a laser pulse, was monitored.

PROJECTS

This arrangement leads to different standard reaction Gibbs energies because the redox couples ZnP +/ZnP and ZnP +/ZnP* have different standard potentials, with the electronically excited porphyrin being a more powerful reductant. Use the following data to estimate the reorganization energy for this system: DrG 9/eV ket /(106 s−1)

0.665 0.657

0.705 1.52

0.745 1.52

0.975 8.99

1.015 5.76

5.50 10.1

(b) The photosynthetic reaction center of the purple photosynthetic bacterium Rhodopseudomonas viridis is a protein complex containing a number of bound co-factors that participate in electron transfer reactions. The table below shows data compiled by Moser et al., Nature 355, 796 (1992), on the rate constants for electron transfer

309

between different co-factors and their edge-to-edge distances. (BChl, bacteriochlorophyll; BChl2, bacteriochlorophyll dimer, functionally distinct from BChl; BPh, bacteriopheophytin; Q A and Q B, quinone molecules bound to two distinct sites; cyt c559, a cytochrome bound to the reaction center complex.) Are these data in agreement with the behavior predicted by eqn 8.31? If so, evaluate the value of b. Reaction r/nm ket /s−1

BChl−→BPh 0.48 1.58 × 1012

BPh−→Chl +2 0.95 3.98 × 109

Reaction r/nm ket /s−1

Q A−→QB 1.35 3.98 × 107

Q A−→BChl+2 2.24 63.1

BPh−→QA 0.96 1.00 × 109

cyt c559→Chl +2 1.23 1.58 × 108

This page intentionally left blank

PART 3 Biomolecular

structure

We now begin our study of structural biology, the description of the molecular features that determine the structures of and the relationships between structure and function in biological macromolecules. In the following chapters, we shall see how concepts of physical chemistry can be used to establish some of the known ‘rules’ for the assembly of complex structures, such as proteins, nucleic acids, and biological membranes. However, not all the rules are known, so structural biology is a very active area of research that brings together biologists, chemists, physicists, and mathematicians.

This page intentionally left blank

9

Microscopic systems and quantization The first goal of our study of biological molecules and assemblies is to gain a firm understanding of their ultimate structural components, atoms. To make progress, we need to become familiar with the principal concepts of quantum mechanics, the most fundamental description of matter that we currently possess and the only way to account for the structures of atoms. Such knowledge is applied to rational drug design (see the Prolog) when computational chemists use quantum mechanical concepts to predict the structures and reactivities of drug molecules. Quantum mechanical phenomena also form the basis for virtually all the modes of spectroscopy and microscopy that are now so central to investigations of composition and structure in both chemistry and biology. Present-day techniques for studying biochemical reactions have progressed to the point where the information is so detailed that quantum mechanics has to be used in its interpretation. Atomic structure—the arrangement of electrons in atoms—is an essential part of chemistry and biology because it is the basis for the description of molecular structure and molecular interactions. Indeed, without intimate knowledge of the physical and chemical properties of elements, it is impossible to understand the molecular basis of biochemical processes, such as protein folding, the formation of cell membranes, and the storage and transmission of information by DNA.

Principles of quantum theory The role—indeed, the existence—of quantum mechanics was appreciated only during the twentieth century. Until then it was thought that the motion of atomic and subatomic particles could be expressed in terms of the laws of classical mechanics introduced in the seventeenth century by Isaac Newton (see Fundamentals F.3), for these laws were very successful at explaining the motion of planets and everyday objects such as pendulums and projectiles. Classical physics is based on three ‘obvious’ assumptions:

Principles of quantum theory

The emergence of the quantum theory In the laboratory 9.1 Electron microscopy 9.2 The Schrödinger equation 9.3 The uncertainty principle

313

9.1

Applications of quantum theory

314 317 318 321 323

Translation 324 Case study 9.1 The electronic structure of b-carotene 327 9.4

In the laboratory 9.2

Scanning probe microscopy 9.5 Rotation Case study 9.2 The electronic structure of phenylalanine 9.6 Vibration Case study 9.3 The vibration of the N–H bond of the peptide link

329 331

Hydrogenic atoms

337

9.7 9.8

334 335

336

The permitted energy levels of hydrogenic atoms 338 Atomic orbitals 339

The structures of manyelectron atoms

346

2. Any type of motion can be excited to a state of arbitrary energy.

The orbital approximation and the Pauli exclusion principle 9.10 Penetration and shielding 9.11 The building-up principle 9.12 Three important atomic properties Case study 9.4 The biological role of Zn2+

3. Waves and particles are distinct concepts.

Checklist of key concepts

357

Checklist of key equations

358

Further information 9.1: A justification of the Schrödinger equation

358

1. A particle travels in a trajectory, a path with a precise position and momentum at each instant.

These assumptions agree with everyday experience. For example, a pendulum swings with a precise oscillating motion and can be made to oscillate with any energy simply by pulling it back to an arbitrary angle and then letting it swing freely. Classical mechanics lets us predict the angle of the pendulum and the speed at which it is swinging at any instant.

9.9

346 348 349 352 356

Further information 9.2: The separation of variables procedure 359 Further information 9.3: The Pauli principle

359

Discussion questions

360

Exercises

360

Projects

363

314

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

Towards the end of the nineteenth century, experimental evidence accumulated showing that classical mechanics failed to explain all the experimental evidence on very small particles, such as individual atoms, nuclei, and electrons. It took until 1926 to identify the appropriate concepts and equations for describing them. We now know that classical mechanics is in fact only an approximate description of the motion of particles and the approximation is invalid when it is applied to molecules, atoms, and electrons. 9.1 The emergence of the quantum theory The structure of biological matter cannot be understood without understanding the nature of electrons. Moreover, because many of the experimental tools available to biochemists are based on interactions between light and matter, we also need to understand the nature of light. We shall see, in fact, that matter and light have a lot in common.

Fig. 9.1 A region of the spectrum of radiation emitted by excited iron atoms consists of radiation at a series of discrete wavelengths (or frequencies).

Quantum theory emerged from a series of observations made during the late nineteenth century, from which two important conclusions were drawn. The first conclusion, which countered what had been supposed for two centuries, is that energy can be transferred between systems only in discrete amounts. The second conclusion is that light and particles have properties in common: electromagnetic radiation (light), which had long been considered to be a wave, in fact behaves like a stream of particles, and electrons, which since their discovery in 1897 had been supposed to be particles, but in fact behave like waves. In this section we review the evidence that led to these conclusions, and establish the properties that a valid system of mechanics must accommodate. (a) Atomic and molecular spectra

A spectrum is a display of the frequencies or wavelengths (which are related by l = c/n; see Fundamentals F.3) of electromagnetic radiation that are absorbed or emitted by an atom or molecule. Figure 9.1 shows a typical atomic emission spectrum and Fig. 9.2 shows a typical molecular absorption spectrum. The obvious feature of both is that radiation is absorbed or emitted at a series of discrete frequencies. The emission or absorption of light at discrete frequencies can be understood if we suppose that • the energy of the atoms or molecules is confined to discrete values, for then energy can be discarded or absorbed only in packets as the atom or molecule jumps between its allowed states (Fig. 9.3) • the frequency of the radiation is related to the energy difference between the initial and final states. These assumptions are brought together in the Bohr frequency condition, which relates the frequency n (nu) of radiation to the difference in energy DE between two states of an atom or molecule: Fig. 9.2 When a molecule changes its state, it does so by absorbing radiation at definite frequencies. This spectrum of chlorophyll (Atlas R3) suggests that the molecule (and molecules in general) can possess only certain energies, not a continuously variable energy.

DE = hn

Bohr frequency relation

(9.1)

where h is the constant of proportionality. The additional evidence that we describe below confirms this simple relation and gives the value h = 6.626 × 10−34 J s. This constant is now known as Planck’s constant, for it arose in a context that had been suggested by the German physicist Max Planck. At this point we can conclude that one feature of nature that any system of mechanics must accommodate is that the internal modes of atoms and molecules

9.1 THE EMERGENCE OF THE QUANTUM THEORY

315

can possess only certain energies; that is, these modes are quantized. The limitation of energies to discrete values is called the quantization of energy. (b) Wave–particle duality

In Fundamentals F.3 we saw that classical physics describes light as electromagnetic radiation, an oscillating electromagnetic field that spreads as a harmonic wave through empty space, the vacuum, at a constant speed c. A new view of electromagnetic radiation began to emerge in 1900 when the German physicist Max Planck discovered that the energy of an electromagnetic oscillator is limited to discrete values and cannot be varied arbitrarily. This proposal is quite contrary to the viewpoint of classical physics, in which all possible energies are allowed. In particular, Planck found that the permitted energies of an electromagnetic oscillator of frequency n are integer multiples of hn: E = nhn

n = 0, 1, 2, . . .

Quantization of energy in electromagnetic oscillators

(9.2)

where h is Planck’s constant. This conclusion inspired Albert Einstein to conceive of radiation as consisting of a stream of particles, each particle having an energy hn. When there is only one such particle present, the energy of the radiation is hn, when there are two particles of that frequency, their total energy is 2hn, and so on. These particles of electromagnetic radiation are now called photons. According to the photon picture of radiation, an intense beam of monochromatic (singlefrequency) radiation consists of a dense stream of identical photons; a weak beam of radiation of the same frequency consists of a relatively small number of the same type of photons. Evidence that confirms the view that radiation can be interpreted as a stream of particles comes from the photoelectric effect, the ejection of electrons from metals when they are exposed to ultraviolet radiation (Fig. 9.4). Experiments show that no electrons are ejected, regardless of the intensity of the radiation, unless the frequency exceeds a threshold value characteristic of the metal. On the other hand, even at low light intensities, electrons are ejected immediately if the frequency is above the threshold value. These observations strongly suggest an interpretation of the photoelectric effect in which an electron is ejected in a collision with a particle-like projectile, the photon, provided the projectile carries enough energy to expel the electron from the metal. When the photon collides with an electron, it gives up all its energy, so we should expect electrons to appear as soon as the collisions begin, provided each photon carries sufficient energy. That is, through the principle of conservation of energy, the photon energy should be equal to the sum of the kinetic energy of the electron and the work function F (uppercase phi) of the metal, the energy required to remove the electron from the metal (Fig. 9.5). The photoelectric effect is strong evidence for the existence of photons and shows that light has certain properties of particles, a view that is contrary to the classical wave theory of light. A crucial experiment performed by the American physicists Clinton Davisson and Lester Germer in 1925 challenged another classical idea by showing that matter is wavelike: they observed the diffraction of electrons by a crystal (Fig. 9.6). Diffraction is the interference between waves caused by an object in their path and results in a series of bright and dark fringes where the waves are detected (Fig. 9.7). It is a typical characteristic of waves. The Davisson–Germer experiment, which has since been repeated with other particles (including molecular hydrogen), shows clearly that ‘particles’ have

Fig. 9.3 Spectral features can be accounted for if we assume that a molecule emits (or absorbs) a photon as it changes between discrete energy levels. Highfrequency radiation is emitted (or absorbed) when the two states involved in the transition are widely separated in energy; low-frequency radiation is emitted when the two states are close in energy. In absorption or emission, the change in the energy of the molecule, DE, is equal to hn, where n is the frequency of the radiation.

Fig. 9.4 The experimental arrangement to demonstrate the photoelectric effect. A beam of ultraviolet radiation is used to irradiate a patch of the surface of a metal, and electrons are ejected from the surface if the frequency of the radiation is above a threshold value that depends on the metal.

316

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

In the photoelectric effect, an incoming photon brings a definite quantity of energy, hn. It collides with an electron close to the surface of the metal target and transfers its energy to it. The difference between the work function, F, and the energy hn appears as the kinetic energy of the photoelectron, the electron ejected by the photon.

Fig. 9.5

Fig. 9.6 In the Davisson–Germer experiment, a beam of electrons was directed on a single crystal of nickel, and the scattered electrons showed a variation in intensity with angle that corresponded to the pattern that would be expected if the electrons had a wave character and were diffracted by the layers of atoms in the solid.

wavelike properties. We have also seen that ‘waves’ have particle-like properties. Thus we are brought to the heart of modern physics. When examined on an atomic scale, the concepts of particle and wave melt together, particles taking on the characteristics of waves and waves the characteristics of particles. This joint wave–particle character of matter and radiation is called wave–particle duality. You should keep this extraordinary, perplexing, and at the time revolutionary idea in mind whenever you are thinking about matter and radiation at an atomic scale. As these concepts emerged there was an understandable confusion—which continues to this day—about how to combine both aspects of matter into a single description. Some progress was made by Louis de Broglie when, in 1924, he suggested that any particle traveling with a linear momentum, p, should have (in some sense) a wavelength l given by the de Broglie relation: l=

Fig. 9.8 According to the de Broglie relation, a particle with low momentum has a long wavelength, whereas a particle with high momentum has a short wavelength. A high momentum can result either from a high mass or from a high velocity (because p = mv). Macroscopic objects have such large masses that, even if they are traveling very slowly, their wavelengths are undetectably short.

Fig. 9.7 When two waves (drawn as blue and orange lines) are in the same region of space they interfere (with the resulting wave drawn as a red line). Depending on the relative positions of peaks and troughs, they may interfere (a) constructively, to given an enhanced amplitude), or (b) destructively, to give a smaller amplitude.

h p

de Broglie relation

(9.3)

The wave corresponding to this wavelength, what de Broglie called a ‘matter wave’, has the mathematical form sin(2px/l). The de Broglie relation implies that the wavelength of a ‘matter wave’ should decrease as the particle’s speed increases (Fig. 9.8). The relation also implies that, for a given speed, heavy particles should be associated with waves of shorter wavelengths than those of lighter particles. Equation 9.3 was confirmed by the Davisson–Germer experiment, for the wavelength it predicts for the electrons they used in their experiment agrees with the details of the diffraction pattern they observed. We shall build on the relation, and understand it more, in the next section. Example 9.1

Estimating the de Broglie wavelength of electrons

The wave character of the electron is the key to imaging small samples by electron microscopy (see In the laboratory 9.1). Consider an electron microscope

9.1 THE EMERGENCE OF THE QUANTUM THEORY

in which electrons are accelerated from rest through a potential difference of 15.0 kV. Calculate the wavelength of the electrons. Strategy To use the de Broglie relation, we need to establish a relation between the kinetic energy Ek and the linear momentum p. With p = mv and Ek = 12 mv 2, it follows that Ek = 12 m(p/m)2 = p 2/2m, and therefore p = (2mEk )1/2. The kinetic energy acquired by an electron accelerated from rest by falling through a potential difference V is eV, where e = 1.602 × 10−19 C is the magnitude of its charge, so we can write Ek = eV and, after using me = 9.109 × 10−31 kg for the mass of the electron, p = (2meeV)1/2. Solution By using p = (2meeV)1/2 in de Broglie’s relation (eqn 9.3), we obtain

l=

h (2meeV)1/2

At this stage, all we need do is to substitute the data and use the relations 1 C V = 1 J and 1 J = 1 kg m2 s−2: l=

6.626 × 10−34 J s {2 × (9.109 × 10−31 kg) × (1.602 × 10−19 C) × (1.50 × 104 V)}1/2

= 1.00 × 10−11 m = 10.0 pm Calculate the wavelength of an electron accelerated from rest in an electric potential difference of 1.0 MV (1 MV = 106 V). Self-test 9.1

Answer: 1.2 pm

In the laboratory 9.1

Electron microscopy

The basic approach of illuminating a small area of a sample and collecting light with a microscope has been used for many years to image small specimens. However, the resolution of a microscope, the minimum distance between two objects that leads to two distinct images, is in the order of the wavelength of light being used. Therefore, conventional microscopes employing visible light have resolutions in the micrometer range and cannot resolve features on a scale of nanometers. There is great interest in the development of new experimental probes of very small specimens that cannot be studied by traditional light microscopy. For example, our understanding of biochemical processes, such as enzymatic catalysis, protein folding, and the insertion of DNA into the cell’s nucleus, will be enhanced if it becomes possible to image individual biopolymers—with dimensions much smaller than visible wavelengths—at work. The concept of wave–particle duality is directly relevant to biology because the observation that electrons can be diffracted led to the development of important techniques for the determination of the structures of biologically active matter. One technique that is often used to image nanometer-sized objects is electron microscopy, in which a beam of electrons with a well-defined de Broglie wavelength replaces the lamp found in traditional light microscopes. Instead of glass or quartz lenses, magnetic fields are used to focus the beam. In transmission electron microscopy (TEM), the electron beam passes through the specimen

317

318

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

and the image is collected on a screen. In scanning electron microscopy (SEM), electrons scattered back from a small irradiated area of the sample are detected and the electrical signal is sent to a video screen. An image of the surface is then obtained by scanning the electron beam across the sample.

Fig. 9.9 A TEM image of a cross-section of a plant cell showing chloroplasts, organelles responsible for the reactions of photosynthesis (Chapter 12). Chloroplasts are typically 5 mm long. (Dr Jeremy Burgess/ Science Photo Library.)

As in traditional light microscopy, the resolution of the microscope is governed by the wavelength (in this case, the de Broglie wavelength of the electrons in the beam) and the ability to focus the beam. Electron wavelengths in typical electron microscopes can be as short as 10 pm, but it is not possible to focus electrons well with magnetic lenses so, in the end, typical resolutions of TEM and SEM instruments are about 2 nm and 50 nm, respectively. It follows that electron microscopes cannot resolve individual atoms (which have diameters of about 0.2 nm). Furthermore, only certain samples can be observed under certain conditions. The measurements must be conducted under high vacuum. For TEM observations, the samples must be very thin cross-sections of a specimen and SEM observations must be made on dry samples. Bombardment with high-energy electrons can damage biological samples by excessive heating, ionization, and formation of radicals. These effects can lead to denaturation or more severe chemical transformation of biological molecules, such as the breaking of bonds and formation of new bonds not found in native structures. To minimize such damage, it has become common to cool samples to temperatures as low as 77 K or 4 K (by immersion in liquid N2 or liquid He, respectively) prior to and during examination with the microscope. This technique is known as electron cryomicroscopy.1 A consequence of these stringent experimental requirements is that electron microscopy cannot be used to study living cells. In spite of these limitations, the technique is very useful in studies of the internal structure of cells (Fig. 9.9).

9.2 The Schrödinger equation According to classical mechanics, a particle can have a well-defined trajectory, with a precisely specified position and momentum at each instant (as represented by the precise path in the diagram). According to quantum mechanics, a particle cannot have a precise trajectory; instead, there is only a probability that it may be found at a specific location at any instant. The wavefunction that determines its probability distribution is a kind of blurred version of the trajectory. Here, the wavefunction is represented by areas of shading: the darker the area, the greater the probability of finding the particle there.

Fig. 9.10

The surprising consequences of wave–particle duality led not only to powerful techniques in microscopy and medical diagnostics but also to new views of the mechanisms of biochemical reactions, particularly those involving the transfer of electrons and protons. To understand these applications, it is essential to know how electrons behave under the influence of various forces.

We take the de Broglie relation as our starting point for the formulation of a new mechanics and abandon the classical concept of particles moving along trajectories. From now on, we adopt the quantum mechanical view that a particle is spread through space like a wave. Like for a wave in water, where the water accumulates in some places but is low in others, there are regions where the particle is more likely to be found than others. To describe this distribution, we introduce the concept of wavefunction, y (psi), in place of the trajectory, and then set up a scheme for calculating and interpreting y. A ‘wavefunction’ is the modern term for de Broglie’s ‘matter wave’. To a very crude first approximation, we can visualize a wavefunction as a blurred version of a trajectory (Fig. 9.10); however, we shall refine this picture in the following sections. 1

The prefix ‘cryo’ originates from kryos, the Greek word for cold or frost.

9.2 THE SCHRÖDINGER EQUATION

319

(a) The formulation of the equation

In 1926, the Austrian physicist Erwin Schrödinger proposed an equation for calculating wavefunctions. The Schrödinger equation for a single particle of mass m moving with energy E in one dimension is −

ħ2 d2y + Vy = Ey 2m dx 2

Schrödinger equation

(9.4a)

Compact form of the Schrödinger equation

(9.4b)

You will often see eqn 9.4a written in the very compact form Ĥy = Ey

where Ĥy stands for everything on the left of eqn 9.4a. The quantity Ĥ is called the hamiltonian of the system after the mathematician William Hamilton, who had formulated a version of classical mechanics that used the concept. It is written with a caret (ˆ) to signify that it is an ‘operator’, something that acts in a particular way on y rather than just multiplying it (as E multiplies y in Ey). You should be aware that much of quantum theory is formulated in terms of various operators, but we shall encounter them only very rarely in this text.2 Technically, the Schrödinger equation is a second-order differential equation. In it, V, which may depend on the position x of the particle, is the potential energy; ħ (which is read h-bar) is a convenient modification of Planck’s constant: ħ=

h = 1.054 × 10−34 J s 2p

We provide a justification of the form of the equation in Further information 9.1. The rare cases where we need to see the explicit forms of its solution will involve very simple functions. For example (and to become familiar with the form of wavefunctions in three simple cases, but not putting in various constants): 1. The wavefunction for a freely moving particle is sin x (exactly as for de Broglie’s matter wave, sin(2px/l)). 2. The wavefunction for the lowest energy state of a particle free to oscillate to and fro near a point is e−x , where x is the displacement from the point (see Section 9.6), 2

3. The wavefunction for an electron in the lowest energy state of a hydrogen atom is e−r, where r is the distance from the nucleus (see Section 9.8). As can be seen, none of these wavefunctions is particularly complicated mathematically. One feature of the solution of any given Schrödinger equation, a feature common to all differential equations, is that an infinite number of possible solutions are allowed mathematically. For instance, if sin x is a solution of the equation, then so too is a sin bx, where a and b are arbitrary constants, with each solution corresponding to a particular value of E. However, it turns out that only some of these solutions are acceptable physically when the motion of a particle is constrained somehow (as in the case of an electron moving under the influence of the electric field of a proton in a hydrogen atom). In such instances, an acceptable solution must satisfy certain constraints called boundary conditions, which we describe shortly (Fig. 9.11). Suddenly, we are at the heart of quantum mechanics: 2

See, for instance, our Physical chemistry (2010).

Although an infinite number of solutions of the Schrödinger equation exist, not all of them are physically acceptable. Acceptable wavefunctions have to satisfy certain boundary conditions, which vary from system to system. In the example shown here, where the particle is confined between two impenetrable walls, the only acceptable wavefunctions are those that fit between the walls (like the vibrations of a stretched string). Because each wavefunction corresponds to a characteristic energy and the boundary conditions rule out many solutions, only certain energies are permissible.

Fig. 9.11

320

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

A note on good practice

The symbol d (see below, right) indicates a small (and, in the limit, infinitesimal) change in a parameter, as in x changing to x + dx. The symbol D indicates a finite (measurable) difference between two quantities, as in DX = Xfinal − Xinitial.

A brief comment

We are supposing throughout that y is a real function (that is, one that does not depend on i = (−1)1/2). In general, y is complex (has both real and imaginary components); in such cases y 2 is replaced by y*y, where y* is the complex conjugate of y. We do not consider complex functions in this text.3

the fact that only some solutions of the Schrödinger equation are acceptable, together with the fact that each solution corresponds to a characteristic value of E, implies that only certain values of the energy are acceptable. That is, when the Schrödinger equation is solved subject to the boundary conditions that the solutions must satisfy, we find that the energy of the system is quantized. Planck and his immediate successors had to postulate the quantization of energy for each system they considered: now we see that quantization is an automatic feature of a single equation, the Schrödinger equation, which is applicable to all systems. Later in this chapter and the next we shall see exactly which energies are allowed in a variety of systems, the most important of which (for chemistry) is an atom. (b) The interpretation of the wavefunction

Before going any further, it will be helpful to understand the physical significance of a wavefunction. The interpretation most widely used is based on a suggestion made by the German physicist Max Born. He made use of an analogy with the wave theory of light, in which the square of the amplitude of an electromagnetic wave is interpreted as its intensity and therefore (in quantum terms) as the number of photons present. The Born interpretation asserts: The probability of finding a particle in a small region of space of volume dV is proportional to y 2dV, where y is the value of the wavefunction in the region. In other words, y 2 is a probability density. As for other kinds of density, such as mass density (ordinary ‘density’), we get the probability itself by multiplying the probability density by the volume of the region of interest. The Born interpretation implies that wherever y 2 is large (‘high probability density’), there is a high probability of finding the particle. Wherever y2 is small (‘low probability density’), there is only a small chance of finding the particle. The density of shading in Fig. 9.12 represents this probabilistic interpretation, an interpretation that accepts that we can make predictions only about the probability of finding a particle somewhere. This interpretation is in contrast to classical physics, which claims to be able to predict precisely that a particle will be at a given point on its path at a given instant.

Example 9.2

Interpreting a wavefunction

The wavefunction of an electron in the lowest energy state of a hydrogen atom is proportional to e−r/a , with a0 = 52.9 pm and r the distance from the nucleus (Fig. 9.13). Calculate the relative probabilities of finding the electron inside a small volume located at (a) r = 0 (that is, at the nucleus) and (b) r = a0 away from the nucleus. 0

A wavefunction y does not have a direct physical interpretation. However, its square (its square modulus if it is complex), y2, tells us the probability of finding a particle at each point. The probability density implied by the wavefunction shown here is depicted by the density of shading in the band at the bottom of the figure. Fig. 9.12

Strategy The probability is proportional to y 2dV evaluated at the specified

location, with y ∝ e−r/a and y 2 ∝ e−2r/a . The volume of interest is so small (even on the scale of the atom) that we can ignore the variation of y within it and write 0

probability ∝ y 2dV with y evaluated at the point in question. 3 For the role, properties, and interpretation of complex wavefunctions, see our Physical chemistry (2010).

9.3 THE UNCERTAINTY PRINCIPLE

321

Solution (a) When r = 0, y 2 ∝ 1.0 (because e0 = 1) and the probability of find-

ing the electron at the nucleus is proportional to 1.0 × dV. (b) At a distance r = a0 in an arbitrary direction, y 2 ∝ e−2, so the probability of being found there is proportional to e−2 × dV = 0.14 × dV. Therefore, the ratio of probabilities is 1.0/0.14 = 7.1. It is more probable (by a factor of 7.1) that the electron will be found at the nucleus than in the same tiny volume located at a distance a0 from the nucleus. Self-test 9.2 The wavefunction for the lowest energy state in the ion He+ is proportional to e−2r/a . Calculate the ratio of probabilities as in Example 9.2, by comparing the cases for which r = 0 and r = a0. Any comment? 0

Answer: The ratio of probabilities is 55; a more compact wavefunction on account of the higher nuclear charge. The wavefunction for an electron in the ground state of a hydrogen atom is an exponentially decaying function of the form e−r/a , where a0 = 52.9 pm is the Bohr radius.

Fig. 9.13

9.3 The uncertainty principle Given that electrons behave like waves, we need to be able to reconcile the predictions of quantum mechanics with the existence of objects, such as biological cells and the organelles within them.

We have seen that, according to the de Broglie relation, a wave of constant wavelength, the wavefunction sin(2px/l), corresponds to a particle with a definite linear momentum p = h/l. However, a wave does not have a definite location at a single point in space, so we cannot speak of the precise position of the particle if it has a definite momentum. Indeed, because a sine wave spreads throughout the whole of space, we cannot say anything about the location of the particle: because the wave spreads everywhere, the particle may be found anywhere in the whole of space. This statement is one half of the uncertainty principle, proposed by Werner Heisenberg in 1927, in one of the most celebrated results of quantum mechanics: It is impossible to specify simultaneously, with arbitrary precision, both the momentum and the position of a particle. Before discussing the principle, we must establish the other half: that if we know the position of a particle exactly, then we can say nothing about its momentum. If the particle is at a definite location, then its wavefunction must be nonzero there and zero everywhere else (Fig. 9.14). We can simulate such a wavefunction by forming a superposition of many wavefunctions; that is, by adding together the amplitudes of a large number of sine functions (Fig. 9.15). This procedure is successful because the amplitudes of the waves add together at one location to give a nonzero total amplitude but cancel everywhere else. In other words, we can create a sharply localized wavefunction by adding together wavefunctions corresponding to many different wavelengths, and therefore, by the de Broglie relation, of many different linear momenta. The superposition of a few sine functions gives a broad, ill-defined wavefunction. As the number of functions used to form the superposition increases, the wavefunction becomes sharper because of the more complete interference between the positive and negative regions of the components. When an infinite number of components are used, the wavefunction is a sharp, infinitely narrow spike like that in Fig. 9.14, which corresponds to perfect localization of the

The wavefunction for a particle with a well-defined position is a sharply spiked function that has zero amplitude everywhere except at the particle’s position.

Fig. 9.14

322

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

The wavefunction for a particle with an ill-defined location can be regarded as the sum (superposition) of several wavefunctions of different wavelength that interfere constructively in one place but destructively elsewhere. As more waves are used in the superposition, the location becomes more precise at the expense of uncertainty in the particle’s momentum. An infinite number of waves are needed to construct the wavefunction of a perfectly localized particle. The numbers against each curve are the number of sine waves used in the superposition. (a) The wavefunctions; (b) the corresponding probability densities.

Fig. 9.15

A brief comment

Strictly, the uncertainty in momentum is the root mean square (r.m.s.) deviation of the momentum from its mean value, Dp = (〈p2〉 − 〈p〉2)1/2, where the angle brackets denote mean values. Likewise, the uncertainty in position is the r.m.s. deviation in the mean value of position, Dx = (〈 x 2〉 − 〈x〉2)1/2.

A representation of the content of the uncertainty principle. The range of locations of a particle is shown by the circles and the range of momenta by the arrows. In (a), the position is quite uncertain, and the range of momenta is small. In (b), the location is much better defined, and now the momentum of the particle is quite uncertain.

Fig. 9.16

particle. Now the particle is perfectly localized, but at the expense of discarding all information about its momentum. The exact, quantitative version of the position–momentum uncertainty relation is DpDx ≥ 12 ħ

Position–momentum uncertainty relation (in one dimension)

(9.5)

The quantity Dp is the ‘uncertainty’ in the linear momentum and Dx is the uncertainty in position (which is proportional to the width of the peak in Fig. 9.15). Equation 9.5 expresses quantitatively the fact that the more closely the location of a particle is specified (the smaller the value of Dx), then the greater the uncertainty in its momentum (the larger the value of Dp) parallel to that coordinate and vice versa (Fig. 9.16). The uncertainty principle applies to location and momentum along the same axis. It is silent on location on one axis and momentum along a perpendicular axis, such as location along the x-axis and momentum parallel to the y-axis. Example 9.3

Using the uncertainty principle

To gain some appreciation of the biological importance—or lack of it—of the uncertainty principle, estimate the minimum uncertainty in the position of

9.3 THE UNCERTAINTY PRINCIPLE each of the following, given that their speeds are known to within 1.0 mm s−1: (a) an electron in a hydrogen atom and (b) a mobile E. coli cell of mass 1.0 pg that can swim in a liquid or glide over surfaces by flexing tail-like structures, known as flagella. Comment on the importance of including quantum mechanical effects in the description of the motion of the electron and the cell. Strategy We can estimate Dp from mDv, where Dv is the uncertainty in the speed v; then we use eqn 9.5 to estimate the minimum uncertainty in position, Dx, where x is the direction in which the projectile is traveling. Solution From DpDx ≥ 12 ħ, the uncertainty in position is

(a) for the electron, with mass 9.109 × 10−31 kg: Dx ≥

ħ 1.054 × 10−34 J s = = 58 m 2Dp 2 × (9.109 × 10−31 kg) × (1.0 × 10−6 m s−1)

(b) for the E. coli cell (using 1 kg = 103 g): Dx ≥

ħ 1.054 × 10−34 J s = = 5.3 × 10−14 m 2Dp 2 × (1.0 × 10−15 kg) × (1.0 × 10−6 m s−1)

For the electron, the uncertainty in position is far larger than the diameter of the atom, which is about 100 pm. Therefore, the concept of a trajectory—the simultaneous possession of a precise position and momentum—is untenable. However, the degree of uncertainty is completely negligible for all practical purposes in the case of the bacterium. Indeed, the position of the cell can be known to within 0.05 per cent of the diameter of a hydrogen atom. It follows that the uncertainty principle plays no direct role in cell biology. However, it plays a major role in the description of the motion of electrons around nuclei in atoms and molecules and, as we shall see soon, the transfer of electrons between molecules and proteins during metabolism.

Self-test 9.3 Estimate the minimum uncertainty in the speed of an electron that can move along the carbon skeleton of a conjugated polyene (such as b-carotene) of length 2.0 nm.

Answer: 29 km s−1

The uncertainty principle epitomizes the difference between classical and quantum mechanics. Classical mechanics supposed, falsely as we now know, that the position and momentum of a particle can be specified simultaneously with arbitrary precision. However, quantum mechanics shows that position and momentum are complementary, that is, not simultaneously specifiable. Quantum mechanics requires us to make a choice: we can specify position at the expense of momentum or momentum at the expense of position.

Applications of quantum theory We shall now illustrate some of the concepts that have been introduced and gain some familiarity with the implications and interpretation of quantum mechanics, including applications to biochemistry. We shall encounter many

323

324

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

other illustrations in the following chapters, for quantum mechanics pervades the whole of chemistry. Just to set the scene, here we describe three basic types of motion: translation (motion in a straight line, like a beam of electrons in the electron microscope), rotation, and vibration. 9.4 Translation The three primitive types of motion—translation, rotation, and vibration—occur throughout science, and we need to be familiar with their quantum mechanical description before we can understand the motion of electrons in atoms and molecules.

In this section we shall see how quantization of energy arises when a particle is confined between two walls. When the potential energy of the particle within the walls is not infinite, the solutions of the Schrödinger equation reveal surprising features, especially the ability of particles to tunnel into and through regions where classical physics would forbid them to be found. A particle in a onedimensional region with impenetrable walls at either end. Its potential energy is zero between x = 0 and x = L and rises abruptly to infinity as soon as the particle touches either wall.

Fig. 9.17

(a) Motion in one dimension

Let’s consider the translational motion of a ‘particle in a box’, a particle of mass m that can travel in a straight line in one dimension (along the x-axis) but is confined between two walls separated by a distance L. The potential energy of the particle is zero inside the box but rises abruptly to infinity at the walls (Fig. 9.17). The particle might be an electron free to move along the linear arrangement of conjugated double bonds in a linear polyene, such as b-carotene (Atlas E1), the molecule responsible for the orange color of carrots and pumpkins. The boundary conditions for this system are the requirement that each acceptable wavefunction of the particle must fit inside the box exactly, like the vibrations of a violin string (as in Fig. 9.11). It follows that the wavelength, l, of the permitted wavefunctions must be one of the values l = 2L, L, 23 L, . . .

A brief comment

More precisely, the boundary conditions stem from the requirement that the wavefunction is continuous everywhere: because the wavefunction is zero outside the box, it must therefore be zero at its edges, at x = 0 and at x = L.

or

l=

2L , with n = 1, 2, 3, . . . n

(9.6)

Each wavefunction is a sine wave with one of these wavelengths; therefore, because a sine wave of wavelength l has the form sin(2px/l), the permitted wavefunctions are yn = N sin

npx L

n = 1, 2, . . .

Wavefunctions for a particle in a one-dimensional box

(9.7)

As shown in the following Justification, the normalization constant, N, a constant that ensures that the total probability of finding the particle anywhere is 1, is equal to (2/L)1/2.

Justification 9.1 The normalization constant

To calculate the constant N, we recall that the wavefunction y must have a form that is consistent with the interpretation of the quantity y(x)2dx as the probability of finding the particle in the infinitesimal region of length dx at the point x given that its wavefunction has the value y(x) at that point. Therefore, the total probability of finding the particle between x = 0 and x = L is the sum (integral) of all the probabilities of its being in each infinitesimal region.

9.4 TRANSLATION

That total probability is 1 (the particle is certainly in the range somewhere), so we know that

冮 y dx = 1 L

2

Substitution of eqn 9.7 turns this expression into

冮 sin npxL dx = 1 L

N2

2

Our task is to solve this equation for N. Because

冮 sin ax dx = x − sin4a2ax + constant 1 2

2

and sin bp = 0 (b = 0, 1, 2, . . .), it follows that, because the sine term is zero at x = 0 and x = L,

冮 sin npxL dx = L L

2

1 2

Therefore, N 2 × 12 L = 1 and hence N = (2/L)1/2. Note that, in this case but not in general, the same normalization factor applies to all the wavefunctions regardless of the value of n.

It is a simple matter to find the permitted energy levels because the only contribution to the energy is the kinetic energy of the particle: the potential energy is zero everywhere inside the box, and the particle is never outside the box. First, we note that it follows from the de Broglie relation, eqn 9.3, that the only acceptable values of the linear momentum are h nh p= = l 2L

n = 1, 2, . . .

(9.8)

Then, because the kinetic energy of a particle of momentum p and mass m is E = p 2/2m, it follows that the permitted energies of the particle are En =

n2h2 8mL2

n = 1, 2, . . .

Quantized energies of a particle in a one-dimensional box

(9.9)

As we see in eqns 9.7 and 9.9, the wavefunctions and energies of a particle in a box are labeled with the number n. A quantum number, of which n is an example, is an integer (in certain cases, as we shall see later, a half-integer) that labels the state of the system. As well as acting as a label, a quantum number specifies certain physical properties of the system: in the present example, n specifies the energy of the particle through eqn 9.9. The permitted energies of the particle are shown in Fig. 9.18 together with the shapes of the wavefunctions for n = 1 to 6. All the wavefunctions except the one of

325

326

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

The allowed energy levels and the corresponding (sine wave) wavefunctions for a particle in a box. Note that the energy levels increase as n2, and so their spacing increases as n increases. Each wavefunction is a standing wave, and successive functions possess one more halfwave and a correspondingly shorter wavelength. Fig. 9.18

lowest energy (n = 1) possess points called nodes where the function passes through zero. Passing through zero is an essential part of the definition: just becoming zero is not sufficient. The points at the edges of the box where y = 0 are not nodes because the wavefunction does not pass through zero there. The number of nodes in the wavefunctions shown in Fig. 9.18 increases from 0 (for n = 1) to 5 (for n = 6) and is n − 1 for a particle in a box in general. It is a general feature of quantum mechanics that the wavefunction corresponding to the state of lowest energy has no nodes, and as the number of nodes in the wavefunctions increases, the energy increases too. The solutions of a particle in a box introduce another important general feature of quantum mechanics. Because the quantum number n cannot be zero (for this system), the lowest energy that the particle may possess is not zero, as would be allowed by classical mechanics, but h2/8mL2 (the energy when n = 1). This lowest, irremovable energy is called the zero-point energy. The existence of a zero-point energy is consistent with the uncertainty principle. If a particle is confined to a finite region, its location is not completely indefinite; consequently its momentum cannot be specified precisely as zero, and therefore its kinetic energy cannot be precisely zero either. The zero-point energy is not a special, mysterious kind of energy. It is simply the last remnant of energy that a particle cannot give up.

9.4 TRANSLATION

327

For a particle in a box it can be interpreted as the energy arising from a ceaseless fluctuating motion of the particle between the two confining walls of the box. The energy difference between adjacent levels is DE = En+1 − En = (n + 1)2

h2 h2 h2 2 = (2n + 1) − n 8mL2 8mL2 8mL2

(9.10)

This expression shows that the difference decreases as the length L of the box increases and that it becomes zero when the walls are infinitely far apart (Fig. 9.19). Atoms and molecules free to move in laboratory-sized vessels may therefore be treated as though their translational energy is not quantized, because L is so large. The expression also shows that the separation decreases as the mass of the particle increases. Particles of macroscopic mass (like balls and planets and even minute specks of dust) behave as though their translational motion is unquantized. Both these conclusions are true in general: 1. The greater the size of the system, the less important are the effects of quantization. 2. The greater the mass of the particle, the less important are the effects of quantization. Case study 9.1

The electronic structure of b-carotene

Some linear polyenes, of which b-carotene is an example, are important biological co-factors that participate in processes as diverse as the absorption of solar energy in photosynthesis (Chapter 12) and protection against harmful biological oxidations. b-Carotene is a linear polyene in which 21 bonds, 10 single and 11 double, alternate along a chain of 22 carbon atoms. We already know from introductory chemistry that this bonding pattern results in conjugation, the sharing of p electrons among all the carbon atoms in the chain.4 Therefore, the particle in a one-dimensional box may be used as a simple model for the discussion of the distribution of p electrons in conjugated polyenes. If we take each C–C bond length to be about 140 pm, the length L of the molecular box in b-carotene is L = 21 × (1.40 × 10−10 m) = 2.94 × 10−9 m For reasons that will become clear in Sections 9.9 and 10.4, we assume that only one electron per carbon atom is allowed to move freely within the box and that, in the lowest energy state (called the ground state) of the molecule, each level is occupied by two electrons. Therefore, the levels up to n = 11 are occupied. From eqn 9.10 it follows that the separation in energy between the ground state and the state in which one electron is promoted from the n = 11 level to the n = 12 level is (6.626 × 10−34 J s)2 8 × (9.109 × 10−31 kg) × (2.94 × 10−9 m)2 −19 = 1.60 × 10 J

DE = E12 − E11 = (2 × 11 + 1)

We can relate this energy difference to the properties of the light that can bring about the transition. From the Bohr frequency condition (eqn 9.1), this energy separation corresponds to a frequency of 4

The quantum mechanical basis for conjugation is discussed in Chapter 10.

(a) A narrow box has widely spaced energy levels; (b) a wide box has closely spaced energy levels. (In each case, the separations depend on the mass of the particle too.)

Fig. 9.19

328

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

n=

DE 1.60 × 10−19 J = = 2.41 × 1014 Hz h 6.626 × 10−34 J s

(we have used 1 s−1 = 1 Hz) and a wavelength (l = c/n) of 1240 nm; the experimental value is 497 nm. This model of b-carotene is primitive and the agreement with experiment not very good, but the fact that the calculated and experimental values are of the same order of magnitude is encouraging as it suggests that the model is not ludicrously wrong. Moreover, the model gives us some insight into the origins of quantized energy levels in conjugated systems and predicts, for example, that the separation between adjacent energy levels decreases as the number of carbon atoms in the conjugated chain increases. In other words, the wavelength of the light absorbed by conjugated polyenes increases as the chain length increases. We shall develop better models in Chapter 10.

(b) Tunneling

A particle incident on a barrier from the left has an oscillating wavefunction, but inside the barrier there are no oscillations (for E < V ). If the barrier is not too thick, the wavefunction is nonzero at its opposite face, and so oscillation begins again there.

Fig. 9.20

We now need to consider the case in which the potential energy of a particle does not rise to infinity when it is in the walls of the container and E < V. If the walls are thin (so that the potential energy falls to zero again after a finite distance, as for a biological membrane) and the particle is very light (as for an electron or a proton), the wavefunction oscillates inside the box (eqn 9.7), varies smoothly inside the region representing the wall, and oscillates again on the other side of the wall outside the box (Fig. 9.20). Hence, the particle might be found on the outside of a container even though according to classical mechanics it has insufficient energy to escape. Such leakage by penetration through classically forbidden zones is called tunneling. Tunneling is a consequence of the wave character of matter. So, just as radio waves pass through walls and X-rays penetrate soft tissue, so can ‘matter waves’ tunnel through thin walls. The Schrödinger equation can be used to determine the probability of tunneling, the transmission probability, T, of a particle incident on a finite barrier. When the barrier is high (in the sense that V/E >> 1) and wide (in the sense that the wavefunction loses much of its amplitude inside the barrier), we may write5 T ≈ 16ε(1 − ε)e−2kL k =

The wavefunction of a heavy particle decays more rapidly inside a barrier than that of a light particle. Consequently, a light particle has a greater probability of tunneling through the barrier.

Fig. 9.21

{2m(V − E)}1/2 ħ

Transmission probability for a high and wide one-dimensional barrier

(9.11)

where ε = E/V and L is the thickness of the barrier. The transmission probability decreases exponentially with L and with m1/2. It follows that particles of low mass are more able to tunnel through barriers than heavy ones (Fig. 9.21). Hence, tunneling is very important for electrons, moderately important for protons, and negligible for most other heavier particles. The very rapid equilibration of proton transfer reactions (Chapter 4) is also a manifestation of the ability of protons to tunnel through barriers and transfer quickly from an acid to a base. Tunneling of protons between acidic and basic groups is also an important feature of the mechanism of some enzyme-catalyzed reactions. The process may be visualized as a proton passing through an activation barrier rather than having to acquire enough energy to travel over it (Fig. 9.22). Quantum mechanical tunneling can be the dominant process in reactions 5

For details of the calculation, see our Physical chemistry (2010).

9.4 TRANSLATION

involving hydrogen atom or proton transfer when the temperature is so low that very few reactant molecules can overcome the activation energy barrier. One indication that a proton transfer is taking place by tunneling is that an Arrhenius plot (Section 6.6) deviates from a straight line at low temperatures and the rate is higher than would be expected by extrapolation from room temperature. Equation 9.11 implies that the rates of electron transfer processes should decrease exponentially with distance between the electron donor and acceptor. This prediction is supported by the experimental evidence that we discussed in Section 8.11, where we showed that, when the temperature and Gibbs energy of activation are held constant, the rate constant ket of electron transfer is proportional to e−br, where r is the edge-to-edge distance between electron donor and acceptor and b is a constant with a value that depends on the medium through which the electron must travel from donor to acceptor. It follows that tunneling is an essential mechanistic feature of the electron transfer processes between proteins, such as those associated with oxidative phosphorylation.

In the laboratory 9.2

Scanning probe microscopy

Like electron microscopy, scanning probe microscopy (SPM) also opens a window into the world of nanometer-sized specimens and, in some cases, provides details at the atomic level. One version of SPM is scanning tunneling microscopy (STM), in which a platinum–rhodium or tungsten needle is scanned across the surface of a conducting solid. When the tip of the needle is brought very close to the surface, electrons tunnel across the intervening space (Fig. 9.23).

329

A proton can tunnel through the activation energy barrier that separates reactants from products, so the effective height of the barrier is reduced and the rate of the proton transfer reaction increases. The effect is represented by drawing the wavefunction of the proton near the barrier. Proton tunneling is important only at low temperatures, when most of the reactants are trapped on the left of the barrier. Fig. 9.22

In the constant-current mode of operation, the stylus moves up and down corresponding to the form of the surface, and the topography of the surface, including any adsorbates, can be mapped on an atomic scale. The vertical motion of the stylus is achieved by fixing it to a piezoelectric cylinder, which contracts or expands according to the potential difference it experiences. In the constant-z mode, the vertical position of the stylus is held constant and the current is monitored. Because the tunneling probability is very sensitive to the size of the gap (remember the exponential dependence of T on L), the microscope can detect tiny, atom-scale variations in the height of the surface (Fig. 9.24). It is difficult to observe individual atoms in large molecules, such as biopolymers. However, Fig. 9.25 shows that STM can reveal some details of the double helical structure of a DNA molecule on a surface. In atomic force microscopy (AFM), a sharpened tip attached to a cantilever is scanned across the surface. The force exerted by the surface and any molecules attached to it pushes or pulls on the tip and deflects the cantilever (Fig. 9.26). The deflection is monitored by using a laser beam. Because no current needs to pass between the sample and the probe, the technique can be applied to nonconducting surfaces and to liquid samples. Two modes of operation of AFM are common. In contact mode, or constantforce mode, the force between the tip and surface is held constant and the tip makes contact with the surface. This mode of operation can damage fragile samples on the surface. In noncontact, or tapping, mode, the tip bounces up and down with a specified frequency and never quite touches the surface. The amplitude of the tip’s oscillation changes when it passes over a species adsorbed on the surface.

A scanning tunneling microscope makes use of the current of electrons that tunnel between the surface and the tip of the stylus. That current is very sensitive to the height of the tip above the surface.

Fig. 9.23

330

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

An STM image of cesium atoms on a gallium arsenide surface.

Fig. 9.24

Image of a DNA molecule obtained by scanning tunneling microscopy, showing some features that are consistent with the double helical structure discussed in Fundamentals and Chapter 11. (Courtesy of J. Baldeschwieler, CIT.) Fig. 9.25

In atomic force microscopy, a laser beam is used to monitor the tiny changes in position of a probe as it is attracted to or repelled by atoms on a surface.

Fig. 9.26

Figure 9.27 demonstrates the power of AFM, which shows bacterial DNA plasmids on a solid surface. The technique also can visualize in real time processes occurring on the surface, such as the enzymatic degradation of DNA, and conformational changes in proteins. The tip may also be used to cleave biopolymers, achieving mechanically on a surface what enzymes do in solution or in organisms.

(c) Motion in two dimensions An atomic force microscopy image of bacterial DNA plasmids on a mica surface. (Courtesy of Veeco Instruments.)

Fig. 9.27

Now that we have described motion in one dimension, it is a simple matter to step into higher dimensions. The arrangement we consider is like a particle confined to a rectangular box of side LX in the x-direction and LY in the y-direction (Fig. 9.28). The wavefunction varies across the floor of the box, so it is a function of the variables x and y, written as y(x,y). We show in Further information 9.2 that, according to the separation of variables procedure, the wavefunction can be expressed as a product of wavefunctions for each direction y(x,y) = X(x)Y(y)

(9.12)

with each wavefunction satisfying a Schrödinger equation like that in eqn 9.4. The solutions are yn ,n (x,y) = Xn (x)Yn (y) X

Y

X

= A two-dimensional square well. The particle is confined to a rectangular plane bounded by impenetrable walls. As soon as the particle touches a wall, its potential energy rises to infinity.

Fig. 9.28

Y

A 4 D C LXLY F

1/2

sin

A nXpx D A n py D sin Y C LX F C LY F

Wavefunctions of a particle in a twodimensional box

(9.13a)

Figure 9.29 shows some examples of these wavefunctions. The energies are En ,n = En + En =

nX2 h2 nY2 h2 + 8mLX2 8mLY2

=

A nX2 nY2 D h2 + C LX2 LY2 F 8m

X

Y

X

Y

Energies of a particle in a two-dimensional box

(9.13b)

9.5 ROTATION

331

There are two quantum numbers, nX and nY, each allowed the values 1, 2, . . . independently. An especially interesting case arises when the region is a square, with LX = LY = L. The allowed energies are then En ,n = (nX2 + nY2) X

Y

h2 8mL2

(9.14)

This result shows that two different wavefunctions may correspond to the same energy. For example, the wavefunctions with nX = 1, nY = 2 and nX = 2, nY = 1 are different 2 A pxD A 2pyD y1,2(x,y) = sin sin C F C L F L L 2 A 2pxD A py D y2,1(x,y) = sin sin C L F C LF L

(9.15)

but both have the energy 5h2/8mL2. Different states with the same energy are said to be degenerate. Degeneracy occurs commonly in atoms, and is a feature that underlies the structure of the periodic table. The separation of variables procedure is very important because it tells us that energies of independent systems are additive and that their wavefunctions are products of simpler component wavefunctions. We shall encounter it several times in later chapters. Fig. 9.29 Three wavefunctions of a particle confined to a rectangular surface.

9.5 Rotation Rotational motion is the starting point for our discussion of the atom, in which electrons are free to circulate around a nucleus.

To describe rotational motion we need to focus on the angular momentum, J, a vector with a length proportional to the rate of circulation and a direction that indicates the axis of rotation (Fig. 9.30). The magnitude of the angular momentum of a particle that is traveling on a circular path of radius r is defined as J = pr

Magnitude of the angular momentum of a particle moving on a circular path

(9.16)

where p is the magnitude of its linear momentum (p = mv) at any instant. A particle that is traveling at high speed in a circle has a higher angular momentum than a particle of the same mass traveling more slowly. An object with a high angular momentum (such as a flywheel) requires a strong braking force (more precisely, a strong torque) to bring it to a standstill. (a) A particle on a ring

Consider a particle of mass m moving in a horizontal circular path of radius r. The energy of the particle is entirely kinetic because the potential energy is constant and can be set equal to zero everywhere. We can therefore write E = p 2/2m. By using eqn 9.16, we can express this energy in terms of the angular momentum as E=

J 2z 2mr 2

Kinetic energy of a particle moving on a circular path

(9.17)

where Jz is the angular momentum for rotation around the z-axis (the axis perpendicular to the plane). The quantity mr 2 is the moment of inertia of the particle

The angular momentum of a particle of mass m on a circular path of radius r in the xy-plane is represented by a vector J perpendicular to the plane and of magnitude pr.

Fig. 9.30

332

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

Mathematical toolkit 9.1 Vectors

A vector quantity has both magnitude and direction. The vector V shown in the figure has components on the x-, y-, and z-axes with magnitudes vx, vy , and vz , respectively. The direction of each of the components is denoted with a plus sign or minus sign. For example, if vx = −1.0, the x-component of the vector V has a magnitude of 1.0 and points in the −x direction. The magnitude of the vector is denoted v or | V | and is given by

Operations involving vectors are not as straightforward as those involving numbers. We describe the operations we need for this text in Mathematical toolkit 11.1.

v = (vx2 + v y2 + v 2z )1/2

about the z-axis and denoted I: a heavy particle in a path of large radius has a large moment of inertia (Fig. 9.31). It follows that the energy of the particle is E=

A particle traveling on a circular path has a moment of inertia I that is given by mr 2. (a) This heavy particle has a large moment of inertia about the central point; (b) this light particle is traveling on a path of the same radius, but it has a smaller moment of inertia. The moment of inertia plays a role in circular motion that is the analog of the mass for linear motion: a particle with a high moment of inertia is difficult to accelerate into a given state of rotation and requires a strong braking force to stop its rotation.

Fig. 9.31

J 2z 2I

Kinetic energy of a particle on a ring in terms of the moment of inertia

(9.18)

Now we use the de Broglie relation to see that the energy of rotation is quantized. To do so, we express the angular momentum in terms of the wavelength of the particle: Jz = pr =

hr l

The angular momentum in terms of the de Broglie wavelength

(9.19)

Suppose for the moment that l can take an arbitrary value. In that case, the amplitude of the wavefunction depends on the angle f as shown in Fig. 9.32. When the angle increases beyond 2p (that is, 360°), the wavefunction continues to change. For an arbitrary wavelength it gives rise to a different value at each point and the interference between the waves on successive circuits cancels the wave on its previous circuit. Thus, this arbitrarily selected wave cannot survive in the system. An acceptable solution is obtained only if the wavefunction reproduces itself on successive circuits: y(f + 2p) = y(f). We say that the wavefunction must satisfy cyclic boundary conditions. It follows that acceptable wavefunctions have wavelengths that are given by the expression l=

2pr n

n = 0, 1, . . .

(9.20)

where the value n = 0, which gives an infinite wavelength, corresponds to a uniform amplitude. It follows that the permitted energies are En =

(hr/l)2 (nh/2p)2 n2ħ2 = = 2I 2I 2I

(9.21)

with n = 0, ±1, ±2, . . . . It is conventional in the discussion of rotational motion to denote the quantum number by ml in place of n. Therefore, the final expression for the energy levels is Em = l

ml2ħ2 2I

ml = 0, ±1, . . .

Quantized energies of a particle on a ring

(9.22)

These energy levels are drawn in Fig. 9.33. The occurrence of ml2 in the expression for the energy means that two states of motion, such as those with ml = +1

9.5 ROTATION

Fig. 9.32 Two solutions of the Schrödinger equation for a particle on a ring. The circumference has been opened out into a straight line; the points at f = 0 and 2p are identical. The solution labeled (a) is unacceptable because it has different values after each circuit and so interferes destructively with itself. The solution labeled (b) is acceptable because it reproduces itself on successive circuits.

Fig. 9.33 The energy levels of a particle that can move on a circular path. Classical physics allowed the particle to travel with any energy; quantum mechanics, however, allows only discrete energies. Each energy level, other than the one with ml = 0, is doubly degenerate because the particle may rotate either clockwise or counterclockwise with the same energy.

and ml = −1, both correspond to the same energy. This degeneracy arises from the fact that the direction of rotation, represented by positive and negative values of ml, does not affect the energy of the particle. All the states with | ml | > 0 are doubly degenerate because two states correspond to the same energy for each value of | ml |. The state with ml = 0, the lowest energy state of the particle, is nondegenerate, meaning that only one state has a particular energy (in this case, zero). An important additional conclusion is that the angular momentum of a particle is quantized. We can use the relation between angular momentum and linear momentum (angular momentum J = pr), and between linear momentum and the allowed wavelengths of the particle (l = 2pr/ml), to conclude that the angular momentum of a particle around the z-axis is confined to the values Jz = pr =

hr hr h = = ml × l 2pr/ml 2p

(9.23)

That is, the angular momentum of the particle around the axis is confined to the values Jz = ml ħ

z-component of the angular momentum of a particle on a ring

(9.24)

with ml = 0, ±1, ±2, . . . . Positive values of ml correspond to clockwise rotation (as seen from below) and negative values correspond to counterclockwise rotation (Fig. 9.34). The quantized motion can be thought of in terms of the rotation of a bicycle wheel that can rotate only with a discrete series of angular momenta, so that as the wheel is accelerated, the angular momentum jerks from the values 0 (when the wheel is stationary) to ħ, 2ħ, . . . but can have no intermediate value.

333

Fig. 9.34 The significance of the sign of ml. When ml < 0, the particle travels in a counterclockwise direction as viewed from below; when ml > 0, the motion is clockwise.

334

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

A final point concerning the rotational motion of a particle is that it does not have a zero-point energy: ml may take the value 0, so E may be zero. This conclusion is also consistent with the uncertainty principle. Although the particle is certainly between the angles 0 and 360° on the ring, that range is equivalent to not knowing anything about where it is on the ring. Consequently, the angular momentum may be specified exactly, and a value of zero is possible. When the angular momentum is zero precisely, the energy of the particle is also zero precisely.

Case study 9.2

The electronic structure of phenylalanine

Just as the particle in a box gives us some understanding of the distribution and energies of p electrons in linear conjugated systems, the particle on a ring is a useful model for the distribution of p electrons around a cyclic conjugated system. Consider the p electrons of the phenyl group of the amino acid phenylalanine (Atlas A14). We may treat the group as a circular ring of radius 140 pm, with six electrons in the conjugated system moving along the perimeter of the ring. As in Case study 9.1, we assume that only one electron per carbon atom is allowed to move freely around the ring and that in the ground state of the molecule each level is occupied by two electrons. Therefore, only the ml = 0, +1, and −1 levels are occupied (with the last two states being degenerate). From eqn 9.22, the energy separation between the ml = ±1 and the ml = ±2 levels is DE = E ±2 − E ±1 = (4 − 1)

(1.054 × 10−34 J s)2 2 × (9.109 × 10−31 kg) × (1.40 × 10−10 m)2

= 9.33 × 10−19 J This energy separation corresponds to an absorption frequency of 1409 THz and a wavelength of 213 nm; the experimental value for a transition of this kind is 260 nm. Even though the model is primitive, it gives insight into the origin of the quantized p-electron energy levels in cyclic conjugated systems, such as the aromatic side chains of phenylalanine, tryptophan, and tyrosine, the purine and pyrimidine bases in nucleic acids, the heme group, and the chlorophylls.

(b) A particle on a sphere

The wavefunction of a particle on the surface of a sphere must satisfy two cyclic boundary conditions. The wavefunction must reproduce itself after the angles f and q are swept by 360° (or 2p radians). This requirement leads to two quantum numbers for its state of angular momentum.

Fig. 9.35

We now consider a particle of mass m free to move around a central point at a constant radius r. That is, it is free to travel anywhere on the surface of a sphere of radius r. To calculate the energy of the particle, we let—as we did for motion on a ring—the potential energy be zero wherever it is free to travel. Furthermore, when we take into account the requirement that the wavefunction should match as a path is traced over the poles as well as around the equator of the sphere surrounding the central point, we define two cyclic boundary conditions (Fig. 9.35). Solution of the Schrödinger equation leads to the following expression for the permitted energies of the particle: E = l(l + 1)

ħ2 2I

l = 0, 1, 2, . . .

Quantized energies of a particle on a sphere

(9.25)

9.6 VIBRATION

335

As before, the energy of the rotating particle is related classically to its angular momentum J by E = J 2/2I. Therefore, by comparing E = J 2/2I with eqn 9.25, we can deduce that because the energy is quantized, the magnitude of the angular momentum is also confined to the values J = {l(l + 1)}1/2ħ

l = 0, 1, 2 . . .

Magnitude of the angular momentum of a particle on a sphere

(9.26)

where l is the orbital angular momentum quantum number. For motion in three dimensions, the vector J has components Jx, Jy, and Jz along the x-, y-, and z-axes, respectively (Fig. 9.36). We have already seen (in the context of rotation in a plane) that the angular momentum about the z-axis is quantized and that it has the values Jz = ml ħ. However, it is a consequence of there being two cyclic boundary conditions that the values of ml are restricted, so the z-component of the angular momentum is given by Jz = mlħ

ml = l, l − 1, . . . , −l

Magnitude of the z-component of the angular momentum of a particle on a sphere

For motion in three dimensions, the angular momentum vector J has components Jx, Jy , and Jz on the x-, y-, and z-axes, respectively.

Fig. 9.36

(9.27)

and ml is now called the magnetic quantum number. We note that for a given value of l there are 2l + 1 permitted values of ml. Therefore, because the energy is independent of ml (because ml does not appear in the expression for the energy, eqn 9.25) a level with quantum number l is (2l + 1)-fold degenerate. 9.6 Vibration The atoms in a molecule vibrate about their equilibrium positions, and the following description of molecular vibrations sets the stage for a discussion of vibrational spectroscopy (Chapter 12), an important experimental technique for the structural characterization of biological molecules.

The simplest model that describes molecular vibrations is the harmonic oscillator, in which a particle is restrained by a spring that obeys Hooke’s law of force, that the restoring force is proportional to the displacement, x: restoring force = −k f x

Hooke’s law

(9.28a)

The constant of proportionality kf is called the force constant: a stiff spring has a high force constant and a weak spring has a low force constant. We show in the following Justification that the potential energy of a particle subjected to this force increases as the square of the displacement, and specifically V(x) = 12 k f x 2

Potential energy of a harmonic oscillator

(9.28b)

The variation of V with x is shown in Fig. 9.37: it has the shape of a parabola (a curve of the form y = ax 2), and we say that a particle undergoing harmonic motion has a ‘parabolic potential energy’.

Justification 9.2 Potential energy of a harmonic oscillator

Force is the negative slope of the potential energy: F = −dV/dx. Because the infinitesimal quantities may be treated as any other quantity in algebraic manipulations, we rearrange the expression into dV = −Fdx and then integrate

The parabolic potential energy characteristic of a harmonic oscillator. Positive displacements correspond to extension of the spring; negative displacements correspond to compression of the spring.

Fig. 9.37

336

9 MICROSCOPIC SYSTEMS AND QUANTIZATION both sides from x = 0, where the potential energy is V(0), to x, where the potential energy is V(x):

冮 F dx x

V(x) − V(0) = −

Now substitute F = −k f x: V(x) − V(0) = −

x

冮 x dx = k x x

(−k f x)dx = k f

1 2

f

2

We are free to choose V(0) = 0, which then gives eqn 9.28b.

Unlike the earlier cases we considered, the potential energy varies with position, so we have to use V(x) in the Schrödinger equation and solve it using the techniques for solving differential equations. Then we have to select the solutions that satisfy the boundary conditions, which in this case means that they must fit into the parabola representing the potential energy. More precisely, the wavefunctions must all go to zero for large displacements from x = 0: they do not have to go abruptly to zero at the edges of the parabola. The solutions of the Schrödinger equation for a harmonic oscillator are quite hard to find, but once found, they turn out to be very simple. For instance, the energies of the solutions that satisfy the boundary conditions are Ev = (v + 12)hn v = 0, 1, 2 . . . n =

1 A k f D 1/2 2π C mF

Quantized energies of a harmonic oscillator

(9.29)

where m is the mass of the particle and v is the vibrational quantum number.6 These energies form a uniform ladder of values separated by hn (Fig. 9.38). The separation is large for stiff springs and low masses. Figure 9.39 shows the shapes of the first few wavefunctions of a harmonic oscillator. The ground-state wavefunction (corresponding to v = 0 and having the zero-point energy 12 hn) is a bell-shaped curve, a curve of the form e−x (a Gaussian function; see Mathematical toolkit F.2), with no nodes. This shape shows that the particle is most likely to be found at x = 0 (zero displacement) but may be found at greater displacements with decreasing probability. The first excited wavefunction has a node at x = 0 and peaks on either side. Therefore, in this state, the particle will be found most probably with the ‘spring’ stretched or compressed to the same amount. In all the states of a harmonic oscillator the wavefunctions extend beyond the limits of motion of a classical oscillator (Fig. 9.40), but the extent decreases as v increases. This penetration into classically forbidden regions is another example of quantum mechanical tunneling, in this case tunneling into rather than through a barrier. 2

The array of energy levels of a harmonic oscillator. The separation depends on the mass and the force constant. Note the zero-point energy. Fig. 9.38

Case study 9.3

The vibration of the N–H bond of the peptide link

Atoms vibrate relative to one another in molecules with the bond acting like a spring. Therefore, eqn 9.29 describes the allowed vibrational energy levels of molecules. Here we consider the vibration of the N–H bond of the peptide link (1), making the approximation that the relatively heavy C, N, and O atoms 6

Be very careful to distinguish the quantum number v (italic vee) from the frequency n (Greek nu).

9.6 VIBRATION

Fig. 9.39 (a) The wavefunctions and (b) the probability densities of the first three states of a harmonic oscillator. Note how the probability of finding the oscillator at large displacements increases as the state of excitation increases. The wavefunctions and displacements are expressed in terms of the parameter a = (ħ2/mkf )1/4.

form a stationary anchor for the very light H atom. That is, only the H atom moves, vibrating as a simple harmonic oscillator. Because the force constant for an N–H bond can be set equal to 700 N m−1 and the mass of the 1H atom is mH = 1.67 × 10−27 kg, we write n=

337

A schematic illustration of the probability density for finding a harmonic oscillator at a given displacement. Classically, the oscillator cannot be found at displacements at which its total energy is less than its potential energy (because the kinetic energy cannot be negative). A quantum oscillator, however, can tunnel into regions that are classically forbidden. Fig. 9.40

1 A k f D 1/2 1 A 700 N m−1 D 1/2 = = 1.03 × 1014 Hz 2π C mF 2π C 1.67 × 10−27 kgF

or 103 THz. Therefore, we expect that radiation with a frequency of 103 THz, in the infrared range of the spectrum, induces a spectroscopic transition between v = 0 and the v = 1 levels of the oscillator. We shall see in Chapter 12 that the concepts just described represent the starting point for the interpretation of vibrational (infrared) spectroscopy, an important technique for the characterization of biopolymers both in solution and inside biological cells.

Hydrogenic atoms Quantum theory provides the foundation for the description of atomic structure. A hydrogenic atom is a one-electron atom or ion of general atomic number Z. Hydrogenic atoms include H, He+, Li2+, C5+, and even U91+. A many-electron atom is an atom or ion that has more than one electron. Many-electron atoms include all neutral atoms other than H. For instance, helium, with its two electrons, is a many-electron atom in this sense. Hydrogenic atoms, and H in particular, are important because the Schrödinger equation can be solved for them and their structures can be discussed exactly. Furthermore, the concepts learned from a study of hydrogenic atoms can be used to describe the structures of manyelectron atoms and of molecules too.

A note on good practice

To calculate the vibrational frequency precisely, we need to specify the nuclide. Also, the mass to use is the actual atomic mass in kilograms, not the element’s molar mass. In Section 12.3 we explain how to take into account the motion of both atoms in a bond by introducing the ‘effective mass’ of an oscillator.

338

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

Much of the material in the remainder of this chapter is a review of introductory chemistry. However, we provide some detail not commonly covered in introductory chemistry, with the goal of showing how core concepts of quantum mechanics can be applied to atoms. The material also sets the stage for the discussion of molecules in Chapter 10. 9.7 The permitted energy levels of hydrogenic atoms Hydrogenic atoms provide the starting point for the discussion of many-electron atoms and hence of the properties of all atoms and their abilities to form bonds and hence aggregate into molecules.

The quantum mechanical description of the structure of a hydrogenic atom is based on Rutherford’s nuclear model, in which the atom is pictured as consisting of an electron outside a central nucleus of charge +Ze, where Z is the atomic number. To derive the details of the structure of this type of atom, we have to set up and solve the Schrödinger equation in which the potential energy, V, is the Coulombic potential energy (Fundamentals F.3 and eqn F.13) for the interaction between the nucleus of charge Q1 = +Ze and the electron of charge Q2 = −e: V=−

Ze 2 4pε0r

(9.30)

where ε0 = 8.854 × 10−12 C2 J−1 m−1 is the vacuum permittivity. We also need to identify the appropriate boundary conditions that the wavefunctions must satisfy in order to be acceptable. For a hydrogenic atom, these conditions are that the wavefunction must not become infinite anywhere and that it must repeat itself (just like the particle on a sphere) as we circle the nucleus either over the poles or around the equator. With a lot of work, the Schrödinger equation with this potential energy and these boundary conditions can be solved, and we shall summarize the results. As usual, the need to satisfy boundary conditions leads to the conclusion that the electron can have only certain energies. Schrödinger himself found that for a hydrogenic atom of atomic number Z with a nucleus of mass mN, the allowed energy levels are given by the expression En = −hcR

Z2 n2

hcR =

me4 32p 2ε02ħ2

m=

memN me + mN

Energy levels of a hydrogenic atom

(9.31)

and n = 1, 2, . . . . The quantity R , the Rydberg constant, has the dimensions of a wavenumber and is commonly reported in units of reciprocal centimeters (cm−1). The quantity m is the reduced mass. For all except the most precise considerations, the mass of the nucleus is so much bigger than the mass of the electron that the latter may be neglected in the denominator of m, and then m ≈ me. Let’s unpack the significance of eqn 9.31: 1. The quantum number n is called the principal quantum number. It gives the energy of the electron in the atom by substituting its value into eqn 9.31. The energy levels of the hydrogen atom. The energies are relative to a proton and an infinitely distant, stationary electron.

Fig. 9.41

The resulting energy levels are depicted in Fig. 9.41. Note how they are widely separated at low values of n but then converge as n increases. At low values of n the electron is confined close to the nucleus by the pull between opposite charges and the energy levels are widely spaced like those of a particle in a narrow box. At high values of n, when the electron has such a high energy that it can travel out

9.8 ATOMIC ORBITALS

to large distances, the energy levels are close together, like those of a particle in a large box. 2. All the energies are negative, which signifies that an electron in an atom has a lower energy than when it is free. The zero of energy (which occurs at n = ∞) corresponds to the infinitely widely separated (so that the Coulomb potential energy is zero) and stationary (so that the kinetic energy is zero) electron and nucleus. The state of lowest, most negative energy, the ground state of the atom, is the one with n = 1 (the lowest permitted value of n and hence the most negative value of the energy). The energy of this state is E1 = −hcRZ 2 The negative sign means that the ground state lies at hcR Z 2 below the energy of the infinitely separated stationary electron and nucleus. The minimum energy needed to remove an electron completely from an atom is called the ionization energy, I. For a hydrogen atom, the ionization energy is the energy required to raise the electron from the ground state with energy E1 = −hcR to the state corresponding to complete removal of the electron (the state with n = ∞ and zero energy). Therefore, the energy that must be supplied is (using m ≈ me) IH =

mee 4 = 2.179 × 10−18 J 32p 2ε 02ħ2

or 2.179 aJ (1 aJ = 10−18 J). This energy corresponds to 13.60 eV and (after multiplication by NA, Avogadro’s constant) to 1312 kJ mol−1. 3. The energy of a given level, and therefore the separation of neighboring levels, is proportional to Z 2. This dependence on Z 2 stems from two effects. First, an electron at a given distance from a nucleus of charge +Ze has a potential energy that is Z times larger than that of an electron at the same distance from a proton (for which Z = 1). However, the electron is drawn into the vicinity of the nucleus by the greater nuclear charge, so it is more likely to be found closer to the nucleus of charge Z than the proton. This effect is also proportional to Z, so overall the energy of an electron can be expected to be proportional to the square of Z, one factor of Z representing the Z times greater strength of the nuclear field and the second factor of Z representing the fact that the electron is Z times more likely to be found closer to the nucleus. Predict the ionization energy of He+ given that the ionization energy of H is 13.60 eV. Hint: Decide how the energy of the ground state varies with Z.

Self-test 9.4

Answer: IHe+ = 4IH = 54.40 eV

9.8 Atomic orbitals The properties of elements and the formation of chemical bonds are consequences of the shapes and energies of the wavefunctions that describe the distribution of electrons in atoms. We need information about the shapes of these wavefunctions

339

340

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

to understand why compounds of carbon adopt the conformations that are responsible for the unique biological functions of such molecules as proteins, nucleic acids, and lipids.

The wavefunction of the electron in a hydrogenic atom is called an atomic orbital. The name is intended to express something less definite than the ‘orbit’ of classical mechanics. An electron that is described by a particular wavefunction is said to ‘occupy’ that orbital. So, in the ground state of the atom, the electron occupies the orbital of lowest energy (that with n = 1). (a) Shells and subshells

We have remarked that there are three boundary conditions on the orbitals: that the wavefunctions must not become infinite, that they must match as they encircle the equator, and that they must match as they encircle the poles. Each boundary condition gives rise to a quantum number, so each orbital is specified by three quantum numbers that act as a kind of ‘address’ of the electron in the atom. We can suspect that the values allowed to the three quantum numbers are linked because, for instance, to get the right shape on a polar journey, we also have to note how the wavefunction changes shape as it wraps around the equator. The quantum numbers are: • The principal quantum number n, which determines the energy of the orbital through eqn 9.31 and has values n = 1, 2, . . . (without limit)

Principal quantum number

• The orbital angular momentum quantum number l,7 which is restricted to the values l = 0, 1, 2, . . . , n − 1

Orbital angular momentum quantum number

For a given value of n, there are n allowed values of l: all the values are positive (for example, if n = 3, then l may be 0, 1, or 2). • The magnetic quantum number, ml, which is confined to the values ml = l, l − 1, l − 2, . . . . −l

Magnetic quantum number

For a given value of l, there are 2l + 1 values of ml (for example, when l = 3, ml may have any of the seven values +3, +2, +1, 0, −1, −2, −3). A note on good practice

Always give the sign of ml, even when it is positive. So, write ml = +1, not ml = 1.

It follows from the restrictions on the values of the quantum numbers that there is only one orbital with n = 1, because when n = 1 the only value that l can have is 0, and that in turn implies that ml can have only the value 0. Likewise, there are four orbitals with n = 2, because l can take the values 0 and 1, and in the latter case ml can have the three values +1, 0, and −1. In general, there are n2 orbitals with a given value of n. Because the energy of a hydrogenic atom depends only on the principal quantum number n, orbitals of the same value of n but different values of l and ml have the same energy. It follows that all orbitals with the same value of n are degenerate. But be careful: this statement applies only to hydrogenic atoms. A second point is that the average distance of an electron from the nucleus of a hydrogenic atom of atomic number Z increases as n increases. As Z increases, the average distance is reduced because the increasing nuclear charge draws the electron closer in. 7

This quantum number is also called by its older name, the azimuthal quantum number.

9.8 ATOMIC ORBITALS

341

The degeneracy of all orbitals with the same value of n (remember that there are n2 of them) and their similar mean radii is the basis of saying that they all belong to the same shell of the atom. It is common to refer to successive shells by letters: n

1 K

2 L

3 M

4... N...

Thus, all four orbitals of the shell with n = 2 form the L shell of the atom. Orbitals with the same value of n but different values of l belong to different subshells of a given shell. These subshells are denoted by the letters s, p, . . . using the following correspondence: l

0 s

1 p

2 d

3... f...

For the shell with n = 1, there is only one subshell, the one with l = 0. For the shell with n = 2 (which allows l = 0, 1), there are two subshells, namely the 2s subshell (with l = 0) and the 2p subshell (with l = 1). The general pattern of the first three shells and their subshells is shown in Fig. 9.42. In a hydrogenic atom, all the subshells of a given shell correspond to the same energy (because, as we have seen, the energy depends on n and not on l). We have seen that if the orbital angular momentum quantum number is l, then ml can take the 2l + 1 values ml = 0, ±1, . . . , ±l. Therefore, each subshell contains 2l + 1 individual orbitals (corresponding to the 2l + 1 values of ml for each value of l). It follows that in any given subshell, the number of orbitals is s 1

p 3

d 5

f... 7...

An orbital with l = 0 (and necessarily ml = 0) is called an s orbital. A p subshell (l = 1) consists of three p orbitals (corresponding to ml = +1, 0, −1). An electron that occupies an s orbital is called an s electron. Similarly, we can speak of p, d, . . . electrons according to the orbitals they occupy. How many orbitals are there in a shell with n = 5 and what is their designation?

Self-test 9.5

Answer: 25; one s, three p, five d, seven f, nine g

(b) The shapes of s orbitals

We saw in Section 9.4c that in certain cases a wavefunction can be separated into factors that depend on different coordinates and that the Schrödinger equation separates into simpler versions for each variable. Application of this separation of variables procedure to the hydrogen atom leads to a Schrödinger equation that separates into one equation for the electron moving around the nucleus (the analog of the particle on a sphere) and an equation for the radial dependence. The wavefunction is written as yn,l,m (r,q,f) = Yl,m (q,f)Rn,l(r) 1

l

Wavefunctions of hydrogenic atoms

(9.32)

The factor R(r) is a function of the distance r from the nucleus and is known as the radial wavefunction. Its form depends on the values of n and l but is

Fig. 9.42 The structures of atoms are described in terms of shells of electrons that are labeled by the principal quantum number n and a series of n subshells of these shells, with each subshell of a shell being labeled by the quantum number l. Each subshell consists of 2l + 1 orbitals.

342

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

independent of ml: that is, all orbitals of the same subshell of a given shell have the same radial wavefunction. In other words, all p orbitals of a shell have the same radial wavefunction, all d orbitals of a shell likewise (but different from that of the p orbitals), and so on. The other factor, Y(q,f), is called the angular wavefunction; it is independent of the distance from the nucleus but varies with the angles q and f. This factor depends on the quantum numbers l and ml. Therefore, regardless of the value of n, orbitals with the same value of l and ml have the same angular wavefunction. In other words, for a given value of ml, a d orbital has the same angular shape regardless of the shell to which it belongs. The mathematical form of a 1s orbital (the wavefunction with n = 1, l = 0, and ml = 0) for a hydrogen atom is

The radial dependence of the wavefunction of a 1s orbital (n = 1, l = 0) and the corresponding probability density. The quantity a0 is the Bohr radius (52.92 pm). Fig. 9.43

Representations of the first two hydrogenic s orbitals, (a) 1s and (b) 2s, in terms of the electron densities (as represented by the density of shading).

y=

1 A 4 D 1/2 −r/a 1 e = e−r/a (4p)1/2 C a 30F (pa 30)1/2

a0 =

4pε0 ħ2 mee 2

Wavefunction of a 1s electron in a hydrogen atom

(9.33)

In this case the angular wavefunction, Y0,0 = 1/(4p)1/2, is a constant, independent of the angles q and f. You should recall that in Section 9.2 we anticipated that a wavefunction for an electron in the ground state of a hydrogen atom has a wavefunction proportional to e−r: eqn 9.33 is its precise form. The constant a0 is called the Bohr radius (because it occurred in the equations based on an early model of the structure of the hydrogen atom proposed by the Danish physicist Niels Bohr) and has the value 52.92 pm. The amplitude of a 1s orbital depends only on the radius, r, of the point of interest and is independent of angle (the latitude and longitude of the point). Therefore, the orbital has the same amplitude at all points at the same distance from the nucleus regardless of direction. Because, according to the Born interpretation (Section 9.2b), the probability density of the electron is proportional to the square of the wavefunction, we now know that the electron will be found with the same probability in any direction (for a given distance from the nucleus). We summarize this angular independence by saying that a 1s orbital is spherically symmetrical. Because the same factor Y occurs in all orbitals with l = 0, all s orbitals have the same spherical symmetry (but different radial dependences). The wavefunction in eqn 9.33 decays exponentially toward zero from a maximum value at the nucleus (Fig. 9.43). It follows that the most probable point at which the electron will be found is at the nucleus itself. A method of depicting the probability of finding the electron at each point in space is to represent y2 by the density of shading in a diagram (Fig. 9.44). A simpler procedure is to show only the boundary surface, the shape that captures about 90 per cent of the electron probability. For the 1s orbital, the boundary surface is a sphere centered on the nucleus (Fig. 9.45). We often need to know the total probability that an electron will be found in the range r to r + dr from a nucleus regardless of its angular position (Fig. 9.46). We can calculate this probability by combining the wavefunction in eqn 9.33 with the Born interpretation and find that for s orbitals, the answer can be expressed as

Fig. 9.44

probability = P(r)dr with P(r) = 4pr 2y2

Radial distribution function of an s orbital

The function P is called the radial distribution function.

(9.34)

9.8 ATOMIC ORBITALS

343

Justification 9.3 The radial distribution function

Consider two spherical shells centered on the nucleus, one of radius r and the other of radius r + dr. The probability of finding the electron at a radius r regardless of its direction is equal to the probability of finding it between these two spherical surfaces. The volume of the region of space between the surfaces is equal to the surface area of the inner shell, 4pr 2, multiplied by the thickness, dr, of the region and is therefore 4pr 2dr. According to the Born interpretation, the probability of finding an electron inside a small volume of magnitude dV is given, for a normalized wavefunction, by the value of y 2dV. Therefore, interpreting V as the volume of the shell, we obtain probability = y 2 × (4pr 2dr)

Fig. 9.45 The boundary surface of an s orbital within which there is a high probability of finding the electron.

as in eqn 9.34. The result we have derived is for any s orbital. For orbitals that depend on angle, the more general form is P(r) = r 2R(r)2, where R(r) is the radial wavefunction. Self-test 9.6 Calculate the probability that an electron in a 1s orbital will be found between a shell of radius a0 and a shell of radius 1.0 pm greater. Hint: Use r = a0 in the expression for the probability density and dr = 1.0 pm in eqn 9.34.

Answer: 0.010

The radial distribution function tells us the total probability of finding an electron at a distance r from the nucleus regardless of its direction. Because r2 increases from 0 as r increases but y2 decreases toward 0 exponentially, P starts at 0, goes through a maximum, and declines to 0 again. The location of the maximum marks the most probable radius (not point) at which the electron will be found. For a 1s orbital of hydrogen, the maximum occurs at a0, the Bohr radius. An analogy that might help to fix the significance of the radial distribution function for an electron is the corresponding distribution for the population of the Earth regarded as a perfect sphere. The radial distribution function is zero at the center of the Earth and for the next 6400 km (to the surface of the planet), when it peaks sharply and then rapidly decays again to zero. It remains virtually zero for all radii more than about 10 km above the surface. Almost all the population will be found very close to r = 6400 km, and it is not relevant that people are dispersed non-uniformly over a very wide range of latitudes and longitudes. The small probabilities of finding people above and below 6400 km anywhere in the world corresponds to the population that happens to be down mines or living in places as high as Denver or Tibet at the time. A 2s orbital (an orbital with n = 2, l = 0, and ml = 0) is also spherical, so its boundary surface is a sphere. Because a 2s orbital spreads farther out from the nucleus than a 1s orbital—because the electron it describes has more energy to climb away from the nucleus—its boundary surface is a sphere of larger radius. The orbital also differs from a 1s orbital in its radial dependence (Fig. 9.47), for although the wavefunction has a nonzero value at the nucleus (like all s orbitals), it passes through zero before commencing its exponential decay toward zero at large distances. We summarize the fact that the wavefunction passes through zero everywhere at a certain radius by saying that the orbital has a radial node. A 3s

Fig. 9.46 The radial distribution function gives the probability that the electron will be found anywhere in a shell of radius r and thickness dr regardless of angle. The graph shows the output from an imaginary shell-like detector of variable radius and fixed thickness dr.

344

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

The radial wavefunctions of the hydrogenic 1s, 2s, 3s, 2p, 3p, and 3d orbitals. Note that the s orbitals have a nonzero and finite value at the nucleus. The vertical scales are different in each case. Fig. 9.47

orbital has two radial nodes; a 4s orbital has three radial nodes. In general, an ns orbital has n − 1 radial nodes. (c) The shapes of p orbitals

Now we turn our attention to the p orbitals (orbitals with l = 1), which have a double-lobed appearance like that shown in Fig. 9.48. The two lobes are separated by a nodal plane that cuts through the nucleus. There is zero probability density for an electron on this plane. Here, for instance, is the explicit form of the 2pz orbital: y=

A brief comment

The radial wavefunction is zero at r = 0, but because r does not take negative values that is not a radial node: the wavefunction does not pass through zero there. A 2p orbital has an angular node, not a radial node.

=

A 3 D 1/2 A 1 D 1/2 r −r/2a cos q × 12 e C 4p F C 6a 30F a 0

A 1 D 1/2 r cos q e−r/2a C 32pa 50F

Wavefunction associated with a 2pz orbital

(9.35)

Note that because y is proportional to r, it is zero at the nucleus, so there is zero probability of finding the electron in a small volume centered on the nucleus. The orbital is also zero everywhere on the plane with cos q = 0, corresponding to q = 90°. The px and py orbitals are similar but have nodal planes perpendicular to the x- and y-axes, respectively.

9.8 ATOMIC ORBITALS

345

The boundary surfaces of p orbitals. A nodal plane passes through the nucleus and separates the two lobes of each orbital.

Fig. 9.48

The exclusion of the electron from the region of the nucleus is a common feature of all atomic orbitals except s orbitals. To understand its origin, we need to recall from Section 9.5 that the value of the quantum number l tells us the magnitude of the angular momentum of the electron around the nucleus (eqn 9.26, J = {l(l + 1)}1/2ħ). For an s orbital, the orbital angular momentum is zero (because l = 0), and in classical terms the electron does not circulate around the nucleus. Because l = 1 for a p orbital, the magnitude of the angular momentum of a p electron is 21/2ħ. As a result, a p electron is flung away from the nucleus by the centrifugal force arising from its motion, but an s electron is not. The same centrifugal effect appears in all orbitals with angular momentum (those for which l > 0), such as d orbitals and f orbitals, and all such orbitals have nodal planes that cut through the nucleus. Each p subshell consists of three orbitals (ml = +1, 0, −1). The three orbitals are normally represented by their boundary surfaces, as depicted in Fig. 9.48. The px orbital has a symmetrical double-lobed shape directed along the x-axis, and similarly the py and pz orbitals are directed along the y- and z-axes, respectively. As n increases, the p orbitals become bigger (for the same reason as s orbitals) and have n − 2 radial nodes. However, their boundary surfaces retain the double-lobed shape shown in the illustration. We can now explain the physical significance of the quantum number ml. It indicates the component of the electron’s orbital angular momentum around an arbitrary axis passing through the nucleus. Positive values of ml correspond to clockwise motion seen from below and negative values correspond to counterclockwise motion. The larger the value of | ml |, the higher is the angular momentum around the arbitrary axis. Specifically: component of angular momentum = ml ħ An s electron (an electron described by an s orbital) has ml = 0 and has no angular momentum about any axis. A p electron can circulate clockwise about an axis as seen from below (ml = +1). Of its total angular momentum of 21/2ħ = 1.414ħ, an amount ħ is due to motion around the selected axis (the rest is due to motion around the other two axes). A p electron can also circulate counterclockwise as seen from below (ml = −1) or not at all (ml = 0) about that selected axis. Except for orbitals with ml = 0, there is not a one-to-one correspondence between the value of ml and the orbitals shown in the illustrations: we cannot say, for instance, that a px orbital has ml = +1. For technical reasons, the orbitals we draw are combinations of orbitals with equal but opposite values of ml (px, for instance, is a combination of the orbitals with ml = +1 and −1). (d) The shapes of d orbitals

When n = 3, l can be 0, 1, or 2. As a result, this shell consists of one 3s orbital, three 3p orbitals, and five 3d orbitals, corresponding to five different values of the magnetic quantum number (ml = +2, +1, 0, −1, −2) for the value l = 2 of the orbital angular momentum quantum number. That is, an electron in the d

346

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

The boundary surfaces of d orbitals. Two nodal planes in each orbital intersect at the nucleus and separate the four lobes of each orbital.

Fig. 9.49

subshell can circulate with five different amounts of angular momentum about an arbitrary axis (+2ħ, +ħ, 0, −ħ, −2ħ). As for the p orbitals, d orbitals with opposite values of ml (and hence opposite senses of motion around an arbitrary axis) may be combined in pairs to give orbitals designated as dxy , dyz , dzx , dx −y , and dz and having the shapes shown in Fig. 9.49. 2

2

2

The structures of many-electron atoms The Schrödinger equation for a many-electron atom is highly complicated because all the electrons interact with one another. Even for a He atom, with its two electrons, no mathematical expression for the orbitals and energies can be given and we are forced to make approximations. Modern computational techniques, however, are able to refine the approximations we are about to make and permit highly accurate numerical calculations of energies and wavefunctions. The periodic recurrence of analogous ground state electron configurations as the atomic number increases accounts for the periodic variation in the properties of atoms. Here we concentrate on two aspects of atomic periodicity—atomic radius and ionization energy—and see how they can help to explain the different biological roles played by different elements. 9.9 The orbital approximation and the Pauli

exclusion principle Here we begin to develop the rules by which electrons occupy orbitals of different energies and shapes. We shall see that our study of hydrogenic atoms was a crucial step toward our goal of ‘building’ many-electron atoms and associating atomic structure with biological function.

In the orbital approximation we suppose that a reasonable first approximation to the exact wavefunction is obtained by letting each electron occupy (that is, have a wavefunction corresponding to) its ‘own’ orbital and writing y = y(1)y(2) . . .

Orbital approximation

(9.36)

where y(1) is the wavefunction of electron 1, y(2) that of electron 2, and so on. We can think of the individual orbitals as resembling the hydrogenic orbitals. For example, consider a model of the helium atom in which both electrons occupy the same 1s orbital, so the wavefunction for each electron is y = (8/pa 30)1/2e−2r/a (because Z = 2). If electron 1 is at a radius r1 and electron 2 is at a radius r2 (and at any angle), then the overall wavefunction for the two-electron atom is 0

y = y(1)y(2) =

A 8 D C pa 30F

1/2

e−2r /a × 1

A 8 D C pa 30F

1/2

e−2r /a = 2

A 8 D −2(r +r )/a e C pa 30F 1

2

9.9 THE ORBITAL APPROXIMATION AND THE PAULI EXCLUSION PRINCIPLE

This description is only approximate because it neglects repulsions between electrons and does not take into account the fact that the nuclear charge is modified by the presence of all the other electrons in the atom. The orbital approximation allows us to express the electronic structure of an atom by reporting its configuration, the list of occupied orbitals (usually, but not necessarily, in its ground state). For example, because the ground state of a hydrogen atom consists of a single electron in a 1s orbital, we report its configuration as 1s1 (read ‘one s one’). A helium atom has two electrons. We can imagine forming the atom by adding the electrons in succession to the orbitals of the bare nucleus (of charge +2e). The first electron occupies a hydrogenic 1s orbital, but because Z = 2, the orbital is more compact than in H itself. The second electron joins the first in the same 1s orbital, and so the electron configuration of the ground state of He is 1s2 (read ‘one s two’). To continue our description, we need to introduce the concept of spin, an intrinsic angular momentum that every electron possesses and that cannot be changed or eliminated (just like its mass or its charge). The name ‘spin’ is evocative of a ball spinning on its axis, and this classical interpretation can be used to help to visualize the motion. However, spin is a purely quantum mechanical phenomenon and has no classical counterpart, so the analogy must be used with care. We shall make use of two properties of electron spin: 1. Electron spin is described by a spin quantum number, s (the analog of l for orbital angular momentum), with s fixed at the single (positive) value of 12 for all electrons at all times. 2. The spin can be clockwise or counterclockwise; these two states are distinguished by the spin magnetic quantum number, ms, which can take the values + 12 or − 12 but no other values (Fig. 9.50). An electron with ms = + 12 is called an a electron and commonly denoted a or ↑; an electron with ms = − 12 is called a b electron and denoted b or ↓. When an atom contains more than one electron, we need to consider the interactions between the electron spin states. Consider lithium (Z = 3), which has three electrons. Two of its electrons occupy a 1s orbital drawn even more closely than in He around the more highly charged nucleus. The third electron, however, does not join the first two in the 1s orbital because a 1s3 configuration is forbidden by a fundamental feature of nature summarized by the Austrian physicist Wolfgang Pauli in the Pauli exclusion principle: No more than two electrons may occupy any given orbital, and if two electrons do occupy one orbital, then their spins must be paired. Electrons with paired spins, denoted ↑↓, have zero net spin angular momentum because the spin angular momentum of one electron is canceled by the spin of the other. In Further information 9.3 we see that the exclusion principle is a consequence of an even deeper statement about wavefunctions. Lithium’s third electron cannot enter the 1s orbital because that orbital is already full: we say that the K shell is complete and that the two electrons form a closed shell. Because a similar closed shell occurs in the He atom, we denote it [He]. The third electron is excluded from the K shell (n = 1) and must occupy the next available orbital, which is one with n = 2 and hence belonging to the L shell. However, we now have to decide whether the next available orbital is the 2s orbital or a 2p orbital and therefore whether the lowest energy configuration of the atom is [He]2s1 or [He]2p1.

347

A classical representation of the two allowed spin states of an electron. The magnitude of the spin angular momentum is (31/2/2)ħ in each case, but the directions of spin are opposite. Fig. 9.50

A note on good practice

The quantum number s should not be confused with or used in place of ms. The spin quantum number s has a single, positive value (12 ; there is no need to write a + sign). Use ms to denote the orientation of the spin (ms = + 12 or − 12), and always include the + sign in ms = + 12.

348

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

9.10 Penetration and shielding Penetration and shielding account for the general form of the periodic table and the physical and chemical properties of the elements. The two effects underlie all the varied properties of the elements and hence their contributions to biological systems.

An electron at a distance r from the nucleus experiences a Coulombic repulsion from all the electrons within a sphere of radius r that is equivalent to a point negative charge located on the nucleus. The effect of the point charge is to reduce the apparent nuclear charge of the nucleus from Ze to Zeffe. Fig. 9.51

An electron in an s orbital (here a 3s orbital) is more likely to be found close to the nucleus than an electron in a p orbital of the same shell. Hence it experiences less shielding and is more tightly bound.

Fig. 9.52

An electron in a many-electron atom experiences a Coulombic repulsion from all the other electrons present. When the electron is at a distance r from the nucleus, the repulsion it experiences from the other electrons can be modeled by a point negative charge located on the nucleus and having a magnitude equal to the charge of the electrons within a sphere of radius r (Fig. 9.51). The effect of the point negative charge is to lower the full charge of the nucleus from Ze to Zeff e, the effective nuclear charge.8 To express the fact that an electron experiences a nuclear charge that has been modified by the other electrons present, we say that the electron experiences a shielded nuclear charge. The electrons do not actually ‘block’ the full Coulombic attraction of the nucleus: the effective charge is simply a way of expressing the net outcome of the nuclear attraction and the electronic repulsions in terms of a single equivalent charge at the center of the atom. The effective nuclear charges experienced by s and p electrons are different because the electrons have different wavefunctions and therefore different distributions around the nucleus (Fig. 9.52). An s electron has a greater penetration through inner shells than a p electron of the same shell in the sense that an s electron is more likely to be found close to the nucleus than a p electron of the same shell (a p orbital, remember, is proportional to r and hence has zero probability density at the nucleus). As a result of this greater penetration, an s electron experiences less shielding than a p electron of the same shell and therefore experiences a larger Zeff. Consequently, by the combined effects of penetration and shielding, an s electron is more tightly bound than a p electron of the same shell. Similarly, a d electron (which has a wavefunction proportional to r 2) penetrates less than a p electron of the same shell, and it therefore experiences more shielding and an even smaller Zeff. As a consequence of penetration and shielding, the energies of orbitals in the same shell of a many-electron atom lie in the order s < p < d < f. The individual orbitals of a given subshell (such as the three p orbitals of the p subshell) remain degenerate because they all have the same radial characteristics and so experience the same effective nuclear charge. We can now complete the Li story. Because the shell with n = 2 has two nondegenerate subshells, with the 2s orbital lower in energy than the three 2p orbitals, the third electron occupies the 2s orbital. This arrangement results in the ground state configuration 1s22s1, or [He]2s1. It follows that we can think of the structure of the atom as consisting of a central nucleus surrounded by a complete heliumlike shell of two 1s electrons and around that a more diffuse 2s electron. The electrons in the outermost shell of an atom in its ground state are called the valence electrons because they are largely responsible for the chemical bonds that the atom forms (and, as we shall see, the extent to which an atom can form bonds is called its ‘valence’). Thus, the valence electron in Li is a 2s electron, and lithium’s other two electrons belong to its core, where they take little part in bond formation. 8 Commonly, Zeff itself is referred to as the ‘effective nuclear charge,’ although strictly that quantity is Zeff e.

9.11 THE BUILDING-UP PRINCIPLE

9.11 The building-up principle The exclusion principle and the consequences of shielding are our keys to understanding the structures of complex atoms and ions, chemical periodicity, and molecular structure.

The extension of the procedure used for H, He, and Li to other atoms is called the building-up principle.9 The building-up principle specifies an order of occupation of atomic orbitals that in most cases reproduces the experimentally determined ground state configurations of atoms and ions. (a) Neutral atoms

We imagine the bare nucleus of atomic number Z and then feed into the available orbitals Z electrons one after the other. The first two rules of the building-up principle are: 1. The order of occupation of orbitals is 1s 2s 2p 3s 3p 4s 3d 4p 5s 4d 5p 6s 5d 4f 6p . . . 2. According to the Pauli exclusion principle, each orbital may accommodate up to two electrons. The order of occupation is approximately the order of energies of the individual orbitals because in general the lower the energy of the orbital, the lower the total energy of the atom as a whole when that orbital is occupied. An s subshell is complete as soon as two electrons are present in it. Each of the three p orbitals of a shell can accommodate two electrons, so a p subshell is complete as soon as six electrons are present in it. A d subshell, which consists of five orbitals, can accommodate up to 10 electrons. As an example, consider a carbon atom. Because Z = 6 for carbon, there are six electrons to accommodate. Two enter and fill the 1s orbital, two enter and fill the 2s orbital, leaving two electrons to occupy the orbitals of the 2p subshell. Hence its ground configuration is 1s22s22p2, or more succinctly [He]2s22p2, with [He] the helium-like 1s2 core. On electrostatic grounds, we can expect the last two electrons to occupy different 2p orbitals, for they will then be farther apart on average and repel each other less than if they were in the same orbital. Thus, one electron can be thought of as occupying the 2px orbital and the other the 2py orbital, and the lowest energy configuration of the atom is [He]2s22px1 2p1y . The same rule applies whenever degenerate orbitals of a subshell are available for occupation. Therefore, another rule of the building-up principle is: 3. Electrons occupy different orbitals of a given subshell before doubly occupying any one of them. It follows that a nitrogen atom (Z = 7) has the configuration [He]2s22px1 2p1y 2p1z. Only when we get to oxygen (Z = 8) is a 2p orbital doubly occupied, giving the configuration [He]2s22p2x 2p1y 2p1z. An additional point arises when electrons occupy degenerate orbitals (such as the three 2p orbitals) singly, as they do in C, N, and O, for there is then no requirement that their spins should be paired. We need to know whether the lowest energy is achieved when the electron spins are the same (both ↑, for instance, 9 The building-up principle is still widely called the Aufbau principle, from the German word for ‘building up’.

349

350

9 MICROSCOPIC SYSTEMS AND QUANTIZATION denoted ↑↑, if there are two electrons in question, as in C) or when they are paired (↑↓). This question is resolved by Hund’s rule: 4. In its ground state, an atom adopts a configuration with the greatest number of unpaired electrons. The explanation of Hund’s rule is complicated, but it reflects the quantum mechanical property of spin correlation, that electrons in different orbitals with parallel spins have a quantum mechanical tendency to stay well apart (a tendency that has nothing to do with their charge: even two ‘uncharged electrons’ would behave in the same way). Their mutual avoidance allows the atom to shrink slightly, so the electron–nucleus interaction is improved when the spins are parallel. We can now conclude that in the ground state of a C atom, the two 2p electrons have the same spin, that all three 2p electrons in an N atom have the same spin, and that the two electrons that singly occupy different 2p orbitals in an O atom have the same spin (the two in the 2px orbital are necessarily paired). Neon, with Z = 10, has the configuration [He]2s22p6, which completes the L shell. This closed-shell configuration is denoted [Ne] and acts as a core for subsequent elements. The next electron must enter the 3s orbital and begin a new shell, and so an Na atom, with Z = 11, has the configuration [Ne]3s1. Like lithium with the configuration [He]2s1, sodium has a single s electron outside a complete core.

Self-test 9.7

Predict the ground state electron configuration of sulfur. Answer: [Ne]3s23p 2x 3p1y 3p1z

This analysis has brought us to the origin of chemical periodicity. The L shell is completed by eight electrons, and so the element with Z = 3 (Li) should have similar properties to the element with Z = 11 (Na). Likewise, Be (Z = 4) should be similar to Mg (Z = 12), and so on up to the noble gases He (Z = 2), Ne (Z = 10), and Ar (Z = 18). Argon has complete 3s and 3p subshells, and as the 3d orbitals are high in energy, the atom effectively has a closed-shell configuration. Indeed, the 4s orbitals are so lowered in energy by their ability to penetrate close to the nucleus that the next electron (for potassium) occupies a 4s orbital rather than a 3d orbital and the K atom resembles an Na atom. The same is true of a Ca atom, which has the configuration [Ar]4s2, resembling that of its congener Mg, which is [Ne]3s2. Ten electrons can be accommodated in the five 3d orbitals, which accounts for the electron configurations of scandium to zinc. The building-up principle has less clear-cut predictions about the ground-state configurations of these elements, and a simple analysis no longer works. Calculations show that for these atoms the energies of the 3d orbitals are always lower than the energy of the 4s orbital. However, experiments show that Sc has the configuration [Ar]3d14s2 instead of [Ar]3d3 or [Ar]3d24s1. To understand this observation, we have to consider the nature of electron–electron repulsions in 3d and 4s orbitals. The most probable distance of a 3d electron from the nucleus is less than that for a 4s electron, so two 3d electrons repel each other more strongly than two 4s electrons. As a result, Sc has the configuration [Ar]3d14s2 rather than the two alternatives, for then the strong electron–electron repulsions in the 3d orbitals are minimized. The total

9.11 THE BUILDING-UP PRINCIPLE

energy of the atom is least despite the cost of allowing electrons to populate the high-energy 4s orbital (Fig. 9.53). The effect just described is generally true for Sc through Zn, so the electron configurations of these atoms are of the form [Ar]3dn4s2, where n = 1 for Sc and n = 10 for Zn. Experiments show that there are two notable exceptions: Cr, with electron configuration [Ar]3d54s1, and Cu, with electron configuration [Ar]3d104s1. At gallium, the energy of the 3d orbitals has fallen so far below those of the 4s and 4p orbitals that they (the full 3d orbitals) can be largely ignored, and the building-up principle can be used in the same way as in preceding periods. Now the 4s and 4p subshells constitute the valence shell, and the period terminates with krypton. Because 18 electrons have intervened since argon, this period is the first long period of the periodic table. The existence of the d block (the ‘transition metals’) reflects the stepwise occupation of the 3d orbitals, and the subtle shades of energy differences along this series give rise to the rich complexity of inorganic (and bioinorganic) d-metal chemistry (Case study 9.4 and Section 10.8). A similar intrusion of the f orbitals in Periods 6 and 7 accounts for the existence of the f block of the periodic table (the lanthanoids and actinoids; still commonly the lanthanides and actinides). (b) Cations and anions

The configurations of cations of elements in the s, p, and d blocks of the periodic table are derived by removing electrons from the ground state configuration of the neutral atom in a specific order. First, we remove any valence p electrons, then the valence s electrons, and then as many d electrons as are necessary to achieve the stated charge. We consider a few examples below. Calcium, an essential constituent of bone and a key player in a number of biochemical processes (such as muscle contraction, cell division, blood clotting, and the conduction of nerve impulses), is taken up by and functions in the cell as the Ca2+ ion. Because the configuration of Ca is [Ar]4s2, the Ca2+ cation has the same configuration, [Ar], as the argon atom. Iron, copper, and manganese can shuttle between different cationic forms and participate in electron transfer reactions that form the core of bioenergetics. For instance, because the configuration of Fe is [Ar]3d64s2, the Fe2+ and Fe3+ cations have the configurations [Ar]3d6 and [Ar]3d5, respectively. These are the oxidation states adopted by the iron ions bound to the protein cytochrome c as it transfers electrons between complexes II and IV in the mitochondrial electron transport chain (Section 5.10). The configurations of anions are derived by continuing the building-up procedure and adding electrons to the neutral atom until the configuration of the next noble gas has been reached. It is the chloride ion, and not elemental chlorine, that works together with Na+ and K+ ions to establish membrane potentials (Section 5.3) and to maintain osmotic pressure (Section 3.10) and charge balance in the cell. The configuration of a Cl− ion is achieved by adding an electron to [Ne]3s23p5, giving the configuration of Ar.

Self-test 9.8

Predict the electron configurations of (a) a Cu2+ ion and

2−

(b) an O ion. Answer: (a) [Ar]3d9, (b) [He]2s22p6

351

Strong electron–electron repulsions in the 3d orbitals are minimized in the ground state of a scandium atom if the atom has the configuration [Ar]3d14s2 (shown on the left) instead of [Ar]3d24s1 (shown on the right). The total energy of the atom is lower when it has the configuration [Ar]3d14s2 despite the cost of populating the high-energy 4s orbital. Fig. 9.53

352

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

9.12 Three important atomic properties The fitness of an element for a biological role is a consequence of electronic structure. We now need to understand how electronic structure affects atomic and ionic radii, and the thermodynamic ability of an atom to release or acquire electrons to form ions or chemical bonds.

We now explore three important atomic properties: the atomic (and ionic) radius, the ionization energy, and the electron affinity. These properties are of great significance in chemistry and biology, for they are controls on the number and types of chemical bonds the atom can form. Indeed, we can use these properties to reveal an important reason for the unique role of carbon in biology. (a) Atomic and ionic radii

Table 9.1 Atomic radii of main-group elements, r/pm

Li

Be

157 112 Na

B 88

Mg Al

C 77 Si

N 74 P

O

F

66 S

64 Cl

191 160 143 118 110 104 99 K

Ca

Ga

Ge

As

Se

Br

235 197 153 122 121 117 114 Rb

Sr

In

Sn

Sb

Te

I

250 215 167 158 141 137 133 Cs

Ba

Tl

Pb

Bi

Po

272 224 171 175 182 167

The variation of atomic radius through the periodic table. Note the contraction of radius following the lanthanoids in Period 6 (following lutetium, Z = 71).

Fig. 9.54

The atomic radius of an element is half the distance between the centers of neighboring atoms in a solid (such as Cu) or, for nonmetals, in a hom*onuclear molecule (such as H2 or S8). If there is one single attribute of an element that determines its chemical properties (either directly, or indirectly through the variation of other properties), then it is atomic radius. In general, atomic radii decrease from left to right across a period and increase down each group (Table 9.1 and Fig. 9.54). The decrease across a period can be traced to the increase in nuclear charge, which draws the electrons in closer to the nucleus. The increase in nuclear charge is partly canceled by the increase in the number of electrons, but because electrons are spread over a region of space, one electron does not fully shield one nuclear charge, so the increase in nuclear charge dominates. The increase in atomic radius down a group (despite the increase in nuclear charge) is explained by the fact that the valence shells of successive periods correspond to higher principal quantum numbers. That is, successive periods correspond to the start and then completion of successive (and more distant) shells of the atom that surround each other like the successive layers of an onion. The need to occupy a more distant shell leads to a larger atom despite the increased nuclear charge. A modification of the increase down a group is encountered in Period 6, for the radii of the atoms late in the d block and in the following regions of the p block are not as large as would be expected by simple extrapolation down the group. The reason can be traced to the fact that in Period 6 the f orbitals are in the process of being occupied to form the 14 lanthanoids, cerium (Ce) to lutetium (Lu). An f electron is a very inefficient shielder of nuclear charge (for reasons connected

9.12 THREE IMPORTANT ATOMIC PROPERTIES

Table 9.2

Ionic radii of selected main-group elements*

Ion

Main biochemical function

Mg 2+

Binds to ATP, constituent of chlorophyll, control of protein folding and muscle contraction

Ca2+

Component of bone and teeth, control of protein folding, hormonal action, blood clotting, and cell division

100

Control of osmotic pressure, charge balance, and membrane potentials

138

Na+ 5 K+ Cl−

6 7

r/pm 72

102 167

*The values are for ions surrounded by six counter-ions in a crystal.

with its radial extension), and as the atomic number increases from Ce to Lu, there is a considerable contraction in radius. By the time the d block resumes (at hafnium, Hf ), the poorly shielded but considerably increased nuclear charge has drawn in the surrounding electrons, and the atoms are compact. They are so compact that the metals in this region of the periodic table (iridium to lead) are very dense. The reduction in radius below that expected by extrapolation from preceding periods is called the lanthanide contraction. The ionic radius of an element is its share of the distance between neighboring ions in an ionic solid (2). That is, the distance between the centers of a neighboring cation and anion is the sum of the two ionic radii. Table 9.2 lists the radii of some ions that play important roles in biochemical processes. When an atom loses one or more valence electrons to form a cation, the remaining atomic core is generally much smaller than the parent atom. Therefore, a cation is often smaller than its parent atom. For example, the atomic radius of Na, with the configuration [Ne]3s1, is 191 pm, but the ionic radius of Na+, with the configuration [Ne], is only 102 pm. Like atomic radii, cationic radii increase down each group because electrons are occupying shells with higher principal quantum numbers. An anion is larger than its parent atom because the electrons added to the valence shell repel one another. Without a compensating increase in the nuclear charge, which would draw the electrons closer to the nucleus and each other, the ion expands. The variation in anionic radii shows the same trend as that for atoms and cations, with the smallest anions at the upper right of the periodic table, close to fluorine. Atoms and ions with the same number of electrons are called isoelectronic. For example, Ca2+, K+, and Cl− have the configuration [Ar] and are isoelectronic. However, their radii differ because they have different nuclear charges. The Ca2+ ion has the largest nuclear charge, so it has the strongest attraction for the electrons and the smallest radius. The Cl− ion has the lowest nuclear charge of the three isoelectronic ions and, as a result, the largest radius. (b) Ionization energy

The minimum energy necessary to remove an electron from a many-electron atom is its first ionization energy, I1. The second ionization energy, I2, is the minimum energy needed to remove a second electron (from the singly charged cation):

353

354

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

The periodic variation of the first ionization energies of the elements.

Fig. 9.55

Table 9.3

First ionization energies of main-group elements, I/eV*

H

He

13.60 Li 5.32 Na 5.14 K 4.34 Rb 4.18 Cs 3.89

24.59 Be

B

C

N

O

F

Ne

9.32

8.30

11.26

14.53

13.62

17.42

21.56

Mg

Al

Si

7.65

5.98

Ca

Ga

6.11

6.00

Sr

In

5.70

5.79

Ba

Tl

5.21

6.11

8.15 Ge 7.90 Sn 7.34 Pb 7.42

P

S

Cl

Ar

10.49

10.36

12.97

15.76

As

Se

Br

Kr

11.81

14.00

9.81 Sb 8.64 Bi 7.29

9.75 Te 9.01 Po 8.42

I

Xe

11.81

14.00

At

Rn

9.64

10.78

*1 eV = 96.485 kJ mol−1.

A note on good practice

E(g) → E+(g) + e−(g)

I1 = E(E+) − E(E)

The physical state of the electron is given because ionization (and electron attachment; see below) is an actual process, unlike in electrochemistry, where the half-reaction, such as E(s) → E+(aq) + e−, is hypothetical and the electron is stateless.

E+(g) → E2+(g) + e−(g)

I2 = E(E2+) − E(E+)

(9.37)

The variation of the first ionization energy through the periodic table is shown in Fig. 9.55, and some numerical values are given in Table 9.3. The ionization energy of an element plays a central role in determining the ability of its atoms to participate in bond formation (for bond formation, as we shall see in Chapter 10, is a consequence of the relocation of electrons from one atom to another). After atomic radius, it is the most important property for determining an element’s chemical characteristics. Lithium has a low first ionization energy: its outermost electron is well shielded from the weakly charged nucleus by the core (Zeff = 1.3 compared with Z = 3) and it is easily removed. Beryllium has a higher nuclear charge than Li, and its outermost electron (one of the two 2s electrons) is more difficult to remove: its ionization energy is larger. The ionization energy decreases between Be and B because in the latter the outermost electron occupies a 2p orbital and is less strongly bound than if it had been a 2s electron. The ionization energy increases between B and C because the latter’s outermost electron is also 2p and the nuclear

9.12 THREE IMPORTANT ATOMIC PROPERTIES

355

charge has increased. Nitrogen has a still higher ionization energy because of the further increase in nuclear charge. There is now a kink in the curve because the ionization energy of O is lower than would be expected by simple extrapolation. At O a 2p orbital must become doubly occupied, and the electron–electron repulsions are increased above what would be expected by simple extrapolation along the row. (The kink is less pronounced in the next row, between P and S, because their orbitals are more diffuse.) The values for O, F, and Ne fall roughly on the same line, the increase of their ionization energies reflecting the increasing attraction of the nucleus for the outermost electrons. The outermost electron in Na is 3s. It is far from the nucleus, and the latter’s charge is shielded by the compact, complete neon-like core. As a result, the ionization energy of Na is substantially lower than that of Ne. The periodic cycle starts again along this row, and the variation of the ionization energy can be traced to similar reasons. (c) Electron affinity

The electron affinity, Eea, is the difference in energy between a neutral atom and its anion. It is the energy released in the process E(g) + e−(g) → E−(g)

Eea = E(E) − E(E−)

(9.38)

The electron affinity is positive if the anion has a lower energy than the neutral atom. Electron affinities (Table 9.4) vary much less systematically through the periodic table than ionization energies. Broadly speaking, however, the highest electron affinities are found close to F. In the halogens, the incoming electron enters the valence shell and experiences a strong attraction from the nucleus. The electron affinities of the noble gases are negative—which means that the anion has a higher energy than the neutral atom—because the incoming electron occupies an orbital outside the closed valence shell. It is then far from the nucleus and repelled by the electrons of the closed shells. The first electron affinity of O is

Table 9.4

Electron affinities of main-group elements, Eea/eV*

H

He

+0.75

0 signifies a ‘positive affinity’ for the added electron. Distinguish the electron affinity from the electron-gain enthalpy, which is negative for such an exothermic process (that is, has the opposite sign to the electron affinity, and differs very slightly in value).

356

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

positive for the same reason as for the halogens. However, the second electron affinity (for the formation of O2− from O−) is strongly negative because although the incoming electron enters the valence shell, it experiences a strong repulsion from the net negative charge of the O− ion. Further analysis of ionization energies and electron affinities can begin to tell us why carbon is an essential building block of complex biological structures. Among the elements in Period 2, C has intermediate values of the ionization energy and electron affinity, so it can share electrons (that is, form covalent bonds) with many other elements, such as H, N, O, S, and, more importantly, other C atoms. As a consequence, such networks as long carbon–carbon chains (as in lipids) and chains of peptide links can form readily. Because the ionization energy and electron affinity of C are neither too high nor too low, the bonds in these covalent networks are neither too strong nor too weak. As a result, biological molecules are sufficiently stable to form viable organisms but are still susceptible to dissociation (essential to catabolism) and rearrangement (essential to anabolism). In Chapter 10 we shall develop additional concepts that will complete this aspect of carbon’s biological role. Case study 9.4

The biological role of Zn 2+

Here we explore how the ionic radius and charge can work together to impart unique chemical properties to an ion, leading to unique biochemical function. Consider the Zn2+ ion, which is found in the active sites of many enzymes. An example is carbonic anhydrase (Atlas P2), which catalyzes the hydration of CO2 in red blood cells to give bicarbonate (hydrogencarbonate) ion: CO2 + H2O → HCO 3− + H+ To understand the catalytic role played by the Zn2+ ion, we need to know that a ‘Lewis acid’ is an electron-poor species that forms a complex with a ‘Lewis base,’ an electron-rich species. Metal cations are good Lewis acids, and molecules with lone pairs of electrons, such as H2O, are good Lewis bases. The Lewis acidity of a metal cation increases with its effective nuclear charge, Zeff (defined here as the charge experienced by a Lewis base on the ‘surface’ of the cation), and decreases with the ionic radius, rion. Among the M2+ d-metal ions found in the active sites of enzymes, Cu2+ and Zn2+ are the best Lewis acids because they have the largest Zeff /rion ratios. Thermodynamically, organisms make use of the Cu2+/Cu+ redox couple for electron transport processes (Chapters 5 and 8) and, generally, the Cu2+ ion does not act as a Lewis acid in biochemical processes. On the other hand, the Zn2+ ion is not used in biological redox reactions but is a ubiquitous biological Lewis acid. To illustrate the consequences of the Lewis acidity of the Zn2+ ion, we consider the mechanism of the hydration of CO2 by carbonic anhydrase (Fig. 9.56). In the first two steps, a Lewis acid–base complex forms between the proteinbound Zn2+ ion and a water molecule, which is then deprotonated. The Zn2+ ion has a large Zeff /rion ratio and gives rise to a strong electric field in its vicinity, so it stabilizes the negative charge on the bound OH− ion, thus effectively lowering the pKw of water from 14 to about 7. Thermodynamically, the Zn2+ ion facilitates the generation of a strong nucleophile, the OH− ion, which can attack CO2 more effectively than H2O. In the next steps, CO2 binds to the active site and then reacts with the bound OH− ion, forming a hydrogencarbonate ion. Release of the bicarbonate ion poises the enzyme for another catalytic cycle.

CHECKLIST OF KEY CONCEPTS

357

The mechanism of the hydration of CO2 by carbonic anhydrase. In the first two steps, a Lewis acid–base complex forms between the protein-bound Zn2+ ion and a water molecule, which is then deprotonated. In the next steps, CO2 binds to the active site and then reacts with the bound OH− ion, forming a bicarbonate ion. Release of the bicarbonate ion poises the enzyme for another catalytic cycle. Fig. 9.56

Checklist of key concepts 1. Atoms and molecules have discrete energy levels, which are revealed by their absorption or emission spectra. 2. Planck proposed that electromagnetic oscillators of frequency n could acquire or discard energy in quanta of magnitude hn. 3. The photoelectric effect is the ejection of electrons when radiation of greater than a threshold frequency is incident on a metal. The photon energy is equal to the sum of the kinetic energy of the electron and the work function F of the metal, the energy required to remove the electron from the metal. 4. The wavelike character of electrons was demonstrated by the Davisson–Germer diffraction experiment. 5. The joint wave–particle character of matter and radiation is called wave–particle duality. 6. A wavefunction, y, contains all the dynamical information about a system and is found by solving the appropriate Schrödinger equation subject to the constraints on the solutions known as boundary conditions. 7. According to the Born interpretation, the probability of finding a particle in a small region of space of volume dV is proportional to y2dV, where y is the value of the wavefunction in the region. 8. According to the Heisenberg uncertainty principle, it is impossible to specify simultaneously, with arbitrary precision, both the momentum and the position of a particle. 9. Because wavefunctions do not, in general, decay abruptly to zero, particles may tunnel into and through classically forbidden regions. 10. A particle undergoes harmonic motion if it is subjected to a Hooke’s-law restoring force (a force proportional to the displacement).

11. Hydrogenic atoms are atoms with a single electron. 12. The wavefunctions of hydrogenic atoms are labeled with three quantum numbers: the principal quantum number n = 1, 2, . . . , the orbital angular momentum quantum number l = 0, 1, . . . , n − 1, and the magnetic quantum number ml = l, l − 1, . . . , −l. 13. s Orbitals are spherically symmetrical and have nonzero amplitude at the nucleus. 14. The p and d orbitals of a shell are shown in Figs. 9.48 and 9.49, respectively. 15. A radial distribution function, P(r), is the probability density for finding an electron between r and r + dr regardless of orientation. 16. An electron possesses an intrinsic angular momentum, its spin, which is described by the quantum numbers s = 12 and ms = ± 12. 17. In the orbital approximation, each electron in a many-electron atom is supposed to occupy its own orbital. 18. The Pauli exclusion principle states that no more than two electrons may occupy any given orbital and if two electrons do occupy one orbital, then their spins must be paired. 19. In a many-electron atom, the orbitals of a given shell lie in the order s < p < d < f as a result of the effects of penetration and shielding. 20. Atomic radii typically decrease from left to right across a period and increase down a group. 21. Ionization energies typically increase from left to right across a period and decrease down a group. 22. Electron affinities are highest toward the top right of the periodic table (near fluorine).

358

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

Checklist of key equations Property

Equation

Bohr frequency relation

DE = hn

Comment

de Broglie relation

l = h/p

Schrödinger equation

−(ħ2/2m)(d2y/dx 2) + Vy = Ey

Motion in one dimension

Heisenberg uncertainty relation

DpDx ≥ 12 ħ

Motion in one dimension

Particle in a box: energy levels zero-point energy wavefunctions

En = n2h2/8mL2 E1 = h2/8mL2 yn(x) = (2/L)1/2 sin(npx/L)

Motion in one dimension n = 1, 2, . . .

T ≈ 16ε(1 − ε)e−2kL, k = {2m(V − E)}1/2/ħ

Motion in one dimension; V/E >> 1

Particle on a ring: energy levels z-component of the angular momentum

Eml = ml2ħ2/2I, I = mr 2 Jz = mlħ

ml = 0, ±1, ±2, . . .

Particle on a sphere energy levels angular momentum z-component of the angular momentum

E = l(l + 1)(ħ2/2I ) J = {l(l + 1)}1/2ħ Jz = mlħ

l = 0, 1, 2, . . .

Harmonic oscillator: potential energy energy levels

V(x) = 12 k f x 2 E v = (v + 12 )hn, n = (1/2p)(kf /m)1/2

Motion in one dimension v = 0, 1, 2, . . .

En = −hcR (Z 2/n2), hcR = me4/(32p2ε 20ħ2), m = memN /(me + mN) yn,l,m (r,q,f) = Yl,m (q,f)Rn,l(r)

n = 1, 2, . . .

P(r) = 4pr 2y 2

s orbitals

Transmission probability

Hydrogenic atoms: energy levels wavefunctions

l

Radial distribution function

ml = l, l − 1, . . . , −l

l = 0, 1, 2, . . . , n − 1 ml = l, l − 1, . . . , −l

l

Further information Further information 9.1 A justification of the Schrödinger equation

The form of the Schrödinger equation can be justified to a certain extent by showing that it implies the de Broglie relation for a freely moving particle. Free motion means motion in a region where the potential energy is zero (V = 0 everywhere). Then Ĥ=−

ħ2 d2 2m dx 2

ħ2 d2y = Ey 2m dx 2

d cos kx = −k sin kx dx

Thus −

A solution is y = sin kx

d sin kx = k cos kx dx

d2 sin(kx) = −k 2 sin(kx) dx 2

and eqn 9.4 simplifies to −

The function sin kx is a wave of wavelength l = 2p/k, as we can see by comparing sin kx with sin(2px/l), the standard form of a harmonic wave with wavelength l. To verify that sin kx is indeed a solution, we insert y = sin kx into both sides of the differential equation and use

k = (2mE)1/2/ħ

ħ2 d2y ħ2 d2 sin (kx) ħ2 k 2ħ2 =− =− (−k 2 sin(kx)) = y 2 2 2m dx 2m 2m 2m dx

According to the Schrödinger equation, the final term of this expression is equal to Ey, so it follows that E = k 2ħ2/2m and k = (2mE)1/2/ħ.

FURTHER INFORMATION

Next, we note that the energy of the particle is entirely kinetic (because V = 0 everywhere), so the total energy of the particle is just its kinetic energy: E = Ek = p2/2m Because E is related to k by E = k 2ħ2/2m, it follows from a comparison of the two equations that p = kħ. Therefore, the linear momentum is related to the wavelength of the wavefunction by p=

2p h h × = l 2p l

Further information 9.2 The separation of variables procedure

We illustrate the separation of variables procedure with motion in a rectangular box as example. The Schrödinger equation for the problem described in Section 9.4c is ħ2 ∂ 2y(x,y) ħ2 ∂ 2y(x,y) − = Ey(x,y) 2m ∂x 2 2m ∂y 2

where we have noted that for a function of two variables the derivatives to be calculated are partial derivatives (see Mathematical toolkit 8.1). For simplicity, we can write this expression as ĤXy(x,y) + ĤYy(x,y) = Ey(x,y) where ĤX operates only on functions of x and ĤY operates only on functions of y. To see if y(x,y) = X(x)Y(y) is indeed a solution, we substitute this product on both sides of the last equation, ĤXX(x)Y(y) + ĤYX(x)Y(y) = EX(x)Y(y) and note that ĤX acts only on X(x), with Y(y) being treated as a constant, and ĤY acts only on Y(y), with X(x) being treated as a constant. Therefore, this equation becomes Y(y)ĤXX(x) + X(x)ĤYY(y) = EX(x)Y(y) When we divide both sides by X(x)Y(y), we obtain 1 1 ĤX X(x) + ĤYY(y) = E X(x) Y(y) Now we come to the crucial part of the argument. The first term on the left depends only on x and the second term depends only on y. Therefore, if x changes, only the first term can change. But its sum with the unchanging second term is the constant E. Therefore, the first term cannot in fact change when x changes. That is, the first term is equal to a constant, which we write as EX. The same argument applies to the second term when y is changed, so it too is equal to a constant, which we write as EY. That is, we have shown that 1 ĤXX(x) = EX X(x)

with EX + EY = E. These two equations are easily turned into ĤX X(x) = EX X(x)

ĤYY(y) = EYY(y)

which we should recognize as the Schrödinger equation for one-dimensional motion, one along the x-axis and the other along the y-axis. Thus, the variables have been separated, and because the boundary conditions are essentially the same for each axis (the only difference being the actual values of the lengths LX and LY), the individual wavefunctions are essentially the same as those already found for the one-dimensional case. Further information 9.3 The Pauli principle

which is the de Broglie relation. We see, in the case of a freely moving particle, that the Schrödinger equation has led to an experimentally verified conclusion.

359

1 ĤYY(y) = EY Y(y)

Some elementary particles have s = 1 and therefore have a higher intrinsic angular momentum than an electron. For our purposes the most important spin-1 particle is the photon. It is a very deep feature of nature that the fundamental particles from which matter is built have half-integral spin (such as electrons and quarks, all of which have s = 12). The particles that transmit forces between these particles, so binding them together into entities such as nuclei, atoms, and planets, all have integral spin (such as s = 1 for the photon, which transmits the electromagnetic interaction between charged particles). Fundamental particles with half-integral spin are called fermions; those with integral spin are called bosons. Matter therefore consists of fermions bound together by bosons. The Pauli exclusion principle is a special case of a general statement called the Pauli principle: When the labels of any two identical fermions are exchanged, the total wavefunction changes sign. When the labels of any two identical bosons are exchanged, the total wavefunction retains the same sign. The Pauli exclusion principle applies only to fermions. ‘Total wavefunction’ means the entire wavefunction, including the spin of the particles. Consider the wavefunction for two electrons y(1,2). The Pauli principle implies that it is a fact of nature that the wavefunction must change sign if we interchange the labels 1 and 2 wherever they occur in the function: y(2,1) = −y(1,2). Suppose the two electrons in an atom occupy an orbital y; then in the orbital approximation the overall wavefunction is y(1)y(2). To apply the Pauli principle, we must deal with the total wavefunction, the wavefunction including spin. There are several possibilities for two spins: the state a(1)a(2) corresponds to parallel spins, whereas (for technical reasons related to the cancelation of each spin’s angular momentum by the other) the combination a(1)b(2) − b(1)a(2) corresponds to paired spins. The total wavefunction of the system is one of the following: Parallel spins: y(1)y(2)a(1)a(2) Paired spins: y(1)y(2){a(1)b(2) − b(1)a(2)} The Pauli principle, however, asserts that for a wavefunction to be acceptable (for electrons), it must change sign when the electrons are exchanged. In each case, exchanging the labels 1

360

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

and 2 converts the factor y(1)y(2) into y(2)y(1), which is the same because the order of multiplying the functions does not change the value of the product. The same is true of a(1)a(2). Therefore, the first combination is not allowed because it does not change sign. The second combination, however, changes to y(2)y(1){a(2)b(1) − b(2)a(1)} = −y(1)y(2){a(1)b(2) − b(1)a(2)} This combination does change sign (it is ‘antisymmetric’) and is therefore acceptable.

Now we see that the only possible state of two electrons in the same orbital allowed by the Pauli principle is the one that has paired spins. This is the content of the Pauli exclusion principle. The exclusion principle is irrelevant when the orbitals occupied by the electrons are different and both electrons may then have (but need not have) the same spin state. Nevertheless, even then the overall wavefunction must still be antisymmetric overall and must still satisfy the Pauli principle itself.

Discussion questions 9.1 Summarize the evidence that led to the introduction of quantum theory. 9.2 Consult texts or online sources to establish the size range for the following particles: a plant cell, an animal cell, a bacterium, a ribosome, a protein (such as chymotrypsin), a small molecule (such as an amino acid), and an atom. Choose among light microscopy (which uses visible light as a probe), electron microscopy, AFM, and STM as suitable techniques for the study of the size and general shape (but not the internal structure) of these particles. 9.3 Discuss the physical origin of the quantization of energy of a particle confined to moving inside a one-dimensional box or on a ring. 9.4 Define, justify, and provide examples of zero-point energy. 9.5 Discuss the physical origins of quantum mechanical tunneling. Why is tunneling more likely to contribute to the mechanisms of electron transfer and proton transfer processes than to mechanisms of group transfer reactions, such as AB + C → A + BC (where A, B, and C are large molecular groups)?

9.6 Explain how the technique of separation of variables is used to simplify the discussion of multi-dimensional problems. When can it not be used? 9.7 List and describe the significance of the quantum numbers needed to specify the internal state of a hydrogenic atom. 9.8 Explain the significance of (a) a boundary surface and (b) the

radial distribution function for hydrogenic orbitals. 9.9 Describe the orbital approximation for the wavefunction of a many-electron atom. What are the limitations of the approximation? 9.10 The d metals iron, copper, and manganese form cations with different oxidation states. For this reason they are found in many oxidoreductases and in several proteins of oxidative phosphorylation and photosynthesis (Section 5.10). Explain why many d metals form cations with different oxidation states.

Exercises 9.11 Calculate the size of the quantum involved in the excitation of

(a) an electronic motion of frequency 1.0 × 1015 Hz, (b) a molecular vibration of period 20 fs, and (c) a pendulum of period 0.50 s. Express the results in joules and in kilojoules per mole. 9.12 Calculate the average power output of a photodetector that

collects 8.0 × 107 photons in 3.8 ms from monochromatic light of wavelength (a) 470 nm, the wavelength produced by some commercially available light-emitting diodes (LED), and (b) 780 nm, a wavelength produced by lasers that are commonly used in compact disc (CD) players. Hint: The total energy emitted by a source or collected by a detector in a given interval is its power multiplied by the time interval of interest (1 J = 1 W s). 9.13 Calculate the de Broglie wavelength of (a) a mass of 1.0 g

traveling at 1.0 m s−1, (b) the same, traveling at 1.00 × 105 km s−1, (c) an He atom traveling at 1000 m s−1 (a typical speed at room temperature), (d) yourself traveling at 8 km h−1, and (e) yourself at rest. 9.14 Calculate the linear momentum per photon, energy per photon,

and the energy per mole of photons for radiation of wavelength

(a) 600 nm (red), (b) 550 nm (yellow), (c) 400 nm (violet), (d) 200 nm (ultraviolet), (e) 150 pm (X-ray), and (f) 1.0 cm (microwave). 9.15 Electron microscopes can obtain images with several hundred-

fold higher resolution than optical microscopes because of the short wavelength obtainable from a beam of electrons. For electrons moving at speeds close to c, the speed of light, the expression for the de Broglie wavelength (eqn 9.3) needs to be corrected for relativistic effects: l=

h

! 2m eV A 1 + eV D # @ e C 2mec 2F $

1/2

where c is the speed of light in a vacuum and V is the potential difference through which the electrons are accelerated. (a) Calculate the de Broglie wavelength of electrons accelerated through 50 kV. (b) Is the relativistic correction important? 9.16 Suppose that you designed a spacecraft to work by photon

pressure. The sail was a completely absorbing fabric of area 1.0 km2 and you directed a red laser beam of wavelength 650 nm onto it at a

EXERCISES

rate of NA photons per second from a base on the Moon. What are (a) the force and (b) the pressure exerted by the radiation on the sail? (c) Suppose the mass of the spacecraft was 1.0 kg. Given that, after a period of acceleration from standstill, speed = (force/mass) × time, how long would it take for the craft to accelerate to a speed of 1.0 m s−1?

361

ground state of the molecule each quantized level is occupied by two electrons. (a) Calculate the energy of an electron in the highest occupied level. (b) Calculate the frequency of radiation that can induce a transition between the highest occupied and lowest unoccupied levels.

9.17 The speed of a certain proton is 350 km s−1. If the uncertainty in its momentum is 0.0100 per cent, what uncertainty in its location must be tolerated? 9.18 An electron is confined to a linear region with a length of the same order as the diameter of an atom (about 100 pm). Calculate the minimum uncertainties in its position and speed. 9.19 Calculate the probability that an electron will be found

(a) between x = 0.1 and 0.2 nm, and (b) between 4.9 and 5.2 nm in a box of length L = 10 nm when its wavefunction is y = (2/L)1/2 sin(2px/L). Hint: Treat the wavefunction as a constant in the small region of interest and interpret dV as dx. 9.20 Repeat Exercise 9.19, but allow for the variation of the

wavefunction in the region of interest. What are the percentage errors in the procedure used in Exercise 9.19? Hint: You will need to integrate y2dx between the limits of interest. The indefinite integral you require is given in Justification 9.1.

9.27 The particle on a ring is a useful model for the motion of electrons around the porphine ring (4), the conjugated macrocycle that forms the structural basis of the heme group and the chlorophylls. We may treat the group as a circular ring of radius 440 pm, with 20 electrons in the conjugated system moving along the perimeter of the ring. As in Exercise 9.26, assume that in the ground state of the molecule quantized each level is occupied by two electrons. (a) Calculate the energy and angular momentum of an electron in the highest occupied level. (b) Calculate the frequency of radiation that can induce a transition between the highest occupied and lowest unoccupied levels.

9.21 What is the probability of finding a particle of mass m in (a) the

left-hand one-third, (b) the central one-third, and (c) the right-hand one-third of a box of length L when it is in the state with n = 1? 9.22 A certain wavefunction is zero everywhere except between

x = 0 and x = L, where it has the constant value A. Normalize the wavefunction. 9.23 The conjugated system of retinal consists of 11 carbon atoms and one oxygen atom. In the ground state of retinal, each level up to n = 6 is occupied by two electrons. Assuming an average internuclear distance of 140 pm, calculate (a) the separation in energy between the ground state and the first excited state in which one electron occupies the state with n = 7 and (b) the frequency of the radiation required to produce a transition between these two states. 9.24 Many biological electron transfer reactions, such as those

associated with biological energy conversion, may be visualized as arising from electron tunneling between protein-bound cofactors, such as cytochromes, quinones, flavins, and chlorophylls. This tunneling occurs over distances that are often greater than 1.0 nm, with sections of protein separating electron donor from acceptor. For a specific combination of electron donor and acceptor, the rate of electron tunneling is proportional to the transmission probability, with k ≈ 7 nm−1 (eqn 9.11). By what factor does the rate of electron tunneling between two co-factors increase as the distance between them changes from 2.0 nm to 1.0 nm? 9.25 The rate, v, at which electrons tunnel through a potential barrier of height 2 eV, like that in a scanning tunneling microscope, and thickness d can be expressed as v = Ae−d/l, with A = 5 × 1014 s−1 and l = 70 pm. (a) Calculate the rate at which electrons tunnel across a barrier of width 750 pm. (b) By what factor is the current reduced when the probe is moved away by a further 100 pm? 9.26 The particle in a two-dimensional well is a useful model for the motion of electrons around the indole ring (3), the conjugated cycle found in the side chain of tryptophan. We may regard indole as a rectangle with sides of length 280 pm and 450 pm, with 10 electrons in the conjugated p system. As in Case study 9.1, we assume that in the

9.28 Use mathematical software or an electronic spreadsheet to plot the wavefunctions y1,1, y1,2, y2,1, and y2,2, and the corresponding probability densities, for a particle in a square well. 9.29 (a) Use the separation of variables procedure to write expressions

for the wavefunctions and energies of a particle trapped in a three-dimensional box with sides LX, LY, and LZ. (b) Using results from part (a), write expressions for the wavefunctions and energies of a particle in a cubic box with sides L. Investigate the existence of degeneracy in this system. 9.30 The HI molecule may be treated as a stationary I atom around which an H atom moves. Assuming that the H atom circulates in a plane at a distance of 161 pm from the I atom, calculate (a) the moment of inertia of the molecule and (b) the greatest wavelength of the radiation that can excite the molecule into rotation. 9.31 Consider again the HI molecule as you did in Exercise 9.30.

Assuming that the H atom oscillates toward and away from the I atom and that the force constant of the HI bond is 314 N m−1, calculate (a) the vibrational frequency of the molecule and (b) the wavelength required to excite the molecule into vibration. (c) Assuming that the force constant of the bond does not change upon isotopic substitution, by what factor will the vibrational frequency of HI change when H is replaced by deuterium? 9.32 The ground state wavefunction of a harmonic oscillator is proportional to e−ax /2, where a depends on the mass and force constant. (a) Normalize this wavefunction. (b) At what displacement 2

362

9 MICROSCOPIC SYSTEMS AND QUANTIZATION

is the oscillator most likely to be found in its ground state? Hint: +∞ For part (a), you will need the integral 2−∝ e−ax dx = (p/a)1/2. For part (b), recall that the maximum (or minimum) of a function f(x) occurs at the value of x for which df/dx = 0. 2

9.33 The solutions of the Schrödinger equation for a harmonic oscillator also apply to diatomic molecules. The only complication is that both atoms joined by the bond move, so the ‘mass’ of the oscillator has to be interpreted carefully. Detailed calculation shows that for two atoms of masses mA and mB joined by a bond of force constant kf, the energy levels are given by eqn 9.29, but the vibrational frequency is

1 A kf D 1/2 n= 2p C m F

m m m= A B mA + mB

and m is called the effective mass of the molecule. Consider the vibration of carbon monoxide, a poison that prevents the transport and storage of O2 (see Exercise 9.48). The bond in a 12C16O molecule has a force constant of 1860 N m−1. (a) Calculate the vibrational frequency, n, of the molecule. (b) In infrared spectroscopy it is common to convert the vibrational frequency of a molecule to its vibrational wavenumber, 6, given by 6 = n/c. What is the vibrational wavenumber of a 12C16O molecule? (c) Assuming that isotopic substitution does not affect the force constant of the C≡O bond, calculate the vibrational wavenumbers of the following molecules: 12 16 C O, 13C16O, 12C18O, 13C18O. 9.34 Predict the ionization energy of Li2+ given that the ionization

energy of He+ is 54.40 eV.

9.36 Consider the ground state of the H atom. (a) At what radius does the probability of finding an electron in a small volume located at a point fall to 25 per cent of its maximum value? (b) At what radius does the radial distribution function have 25 per cent of its maximum value? (c) What is the most probable distance of an electron from the nucleus? Hint: Look for a maximum in the radial distribution function. 9.37 What is the probability of finding an electron anywhere in one

lobe of a p orbital given that it occupies the orbital? 9.38 The (normalized) wavefunction for a 2s orbital in a hydrogen

of an H atom. 9.40 The wavefunction of one of the d orbitals is proportional to sin q cos q. At what angles does it have nodal planes? 9.41 What is the orbital angular momentum (as multiples of ħ) of

an electron in the orbitals (a) 1s, (b) 3s, (c) 3d, (d) 2p, and (e) 3p? Give the numbers of angular and radial nodes in each case. 9.42 How many electrons can occupy subshells with the following

values of l: (a) 0, (b) 3, (c) 5? 9.43 If we lived in a four-dimensional world, there would be one s

orbital, four p orbitals, and nine d orbitals in their respective subshells. (a) Suggest what form the periodic table might take for the first 24 elements. (b) Which elements (using their current names) would be noble gases? (c) On what element would life be likely to be based? 9.44 The central iron ion of cytochrome c changes between the +2 and +3 oxidation states as the protein shuttles electrons between complex III and complex IV of the respiratory chain (Section 5.10). Which do you expect to be larger: Fe2+ or Fe3+? Why?

Group 13 of the periodic table and is most often found in the +1 oxidation state. Aluminum, which causes anemia and dementia, is also a member of the group, but its chemical properties are dominated by the +3 oxidation state. Examine this issue by plotting the first, second, and third ionization energies for the Group 13 elements against atomic number. Explain the trends you observe. Hints: The third ionization energy, I3, is the minimum energy needed to remove an electron from the doubly charged cation: E2+(g) → E3+(g) + e−(g), I3 = E(E3+) − E(E2+). For data, see the links to databases of atomic properties provided in the text’s website. 9.46 How is the ionization energy of an anion related to the electron

affinity of the parent atom? 9.47 To perform many of their biological functions, the Lewis acids

atom is A 1 D C 32pa 03F

9.39 Locate the radial nodes in (a) the 3s orbital and (b) the 4s orbital

9.45 Thallium, a neurotoxin, is currently the heaviest member of

9.35 How many orbitals are present in the N shell of an atom?

y=

function of a hydrogenic 2s electron and plot the function against r. What is the most probable radius at which the electron will be found? (c) For a more accurate determination of the most probable radius at which an electron will be found in an H2s orbital, differentiate the radial distribution function to find where it is a maximum.

1/2

r D −r/2a A 2− e C a0F

where a0 is the Bohr radius. (a) Calculate the probability of finding an electron that is described by this wavefunction in a volume of 1.0 pm3 (i) centered on the nucleus, (ii) at the Bohr radius, and (iii) at twice the Bohr radius. (b) Construct an expression for the radial distribution

Mg 2+ and Ca2+ must be bound to Lewis bases, such as nucleotides (with ATP4− as an example) or the side chains of amino acids in proteins. The equilibrium constant for the association of a doubly charged cation M2+ to a Lewis base increases in the order: Ba2+ < Sr 2+ < Ca2+ < Mg 2+. Provide a molecular interpretation for this trend, which does not depend on the nature of the Lewis base. Hint: Consider the effect of ionic radius.

PROJECTS

363

Projects 9.48 Here we see how infrared spectroscopy can be used to study the

binding of diatomic molecules to heme proteins. We focus on carbon monoxide, which is poisonous because it binds strongly to the Fe2+ ion of the heme group of hemoglobin and myoglobin and interferes with the transport and storage of O2 (Case study 4.1). (a) Estimate the vibrational frequency and wavenumber of CO bound to myoglobin by using the data in Exercise 9.33 and by making the following assumptions: the atom that binds to the heme group is immobilized, the protein is infinitely more massive than either the C or O atom, the C atom binds to the Fe2+ ion, and binding of CO to the protein does not alter the force constant of the C≡O bond. (b) Of the four assumptions made in part (a), the last two are questionable. Suppose that the first two assumptions are still reasonable and that you have at your disposal a supply of myoglobin, a suitable buffer in which to suspend the protein, 12C16O, 13C16O, 12 18 C O, 13C18O, and an infrared spectrometer, an instrument used for the determination of vibrational frequencies. Describe a set of experiments that: (i) proves which atom, C or O, binds to the heme group of myoglobin and (ii) allows for the determination of the force constant of the C≡O bond for myoglobin-bound carbon monoxide. 9.49 The postulation of a plausible reaction mechanism requires careful analysis of many experiments designed to determine the fate of atoms during the formation of products. Observation of the kinetic isotope effect, a decrease in the rate of a chemical reaction on replacement of one atom in a reactant by a heavier isotope, facilitates the identification of bond-breaking events in the rate-determining step. A primary kinetic isotope effect is observed when the ratedetermining step requires the scission of a bond involving the isotope. A secondary kinetic isotope effect is the reduction in reaction rate even though the bond involving the isotope is not broken to form product. In both cases, the effect arises from the change in activation energy that accompanies the replacement of an atom by a heavier isotope on account of changes in the zero-point vibrational energies. We now explore the primary kinetic isotope effect in some detail. Consider a reaction, such as the rearrangements catalyzed by vitamin B12, in which a C–H bond is cleaved. If scission of this bond is the rate-determining step, then the reaction coordinate corresponds to the stretching of the C–H bond and the potential energy profile is shown in Fig. 9.57. On deuteration, the dominant change is the reduction of the zero-point energy of the bond (because the deuterium atom is heavier). The whole reaction profile is not lowered, however, because the relevant vibration in the activated complex has a very low force constant, so there is little zero-point energy associated with the reaction coordinate in either form of the activated complex.

(a) Assume that the change in the activation energy arises only from the change in zero-point energy of the stretching vibration and show that 1 A m D 1/2 Ea(C–D) − Ea(C–H) = NAhc6(C–H) ! 1 − C–H # @ C mC–D F $ 2

Changes in the reaction profile when a C–H bond undergoing cleavage is deuterated. In this illustration, the C–H and C–D bonds are modeled as simple harmonic oscillators. The only significant change is in the zero-point energy of the reactants, which is lower for C–D than for C–H. As a result, the activation energy is greater for C–D cleavage than for C–H cleavage.

Fig. 9.57

where 6 is the relevant vibrational wavenumber and m is the relevant effective mass (Exercise 9.33). (b) Now consider the effect of deuteration on the rate constant, kr, of the reaction. (i) Starting with the Arrhenius equation (eqn 6.19) and assuming that the pre-exponential factor does not change on deuteration, show that the rate constants for the two species should be in the ratio kr(C–D) −l =e kr(C–H)

with

l=

hc6(C–H) ! Am D 1 − C–H C mC–D F 2kT @

1/2

# $

(ii) Does kr(C–D)/kr(C–H) increase or decrease with decreasing temperature? (c) From infrared spectroscopy, the fundamental vibrational wavenumber for stretching of a C–H bond is about 3000 cm−1. Predict the value of the ratio kr(C–D)/kr(C–H) at 298 K. (d) In some cases (including several enzyme-catalyzed reactions), substitution of deuterium for hydrogen results in values of kr(C–D)/ kr(C–H) that are too low to be accounted for by the model described above. Explain this effect.

10 Valence bond theory

Diatomic molecules 10.2 Polyatomic molecules

The chemical bond

365

10.1

Molecular orbital theory

365 367 373

Linear combinations of atomic orbitals 373 10.4 hom*onuclear diatomic molecules 375 Case study 10.1 The biochemical reactivity of O2 and N2 382 10.5 Heteronuclear diatomic molecules 384 Case study 10.2 The biochemistry of NO 386 10.6 The structures of polyatomic molecules 387 10.7 Hückel theory 387 Case study 10.3 The unique role of carbon in biochemistry 391 10.8 d-Metal complexes 392 Case study 10.4 Ligandfield theory and the binding of O2 to hemoglobin 397 10.3

Computational biochemistry

398

Computational techniques 10.10 Graphical output 10.11 The prediction of molecular properties

400

Checklist of key concepts Checklist of key equations Discussion questions Exercises

402 403 403 404

Projects

406

10.9

398 400

The chemical bond, a link between atoms, is central to all aspects of chemistry and biochemistry. The theory of the origin of the numbers, strengths, and three-dimensional arrangements of chemical bonds between atoms is called valence theory. Valence theory is an attempt to explain the properties of molecules ranging from the smallest to the largest. For instance, it explains why N2 is so inert that it acts as a diluent for the aggressive oxidizing power of atmospheric oxygen. At the other end of the scale, valence theory deals with the structural origins of the function of protein molecules and the molecular biology of DNA. Certain ideas of valence theory will be familiar from introductory chemistry. We know that chemical bonds may be classified on the basis of the degree of redistribution of electron density among interacting atomic nuclei: • An ionic bond is formed by the transfer of electrons from one atom to another and the consequent attraction between the ions so formed. • A covalent bond is formed when two atoms share a pair of electrons. The character of a covalent bond, the main focus of this chapter, was identified by G.N. Lewis in 1916, before quantum mechanics was fully developed. Lewis’s original theory was unable to account for the shapes adopted by molecules. The most elementary (but qualitatively quite successful) explanation of the shapes adopted by molecules is the valence-shell electron pair repulsion model (VSEPR model). In this model, which should be familiar from introductory chemistry courses, the shape of a molecule is ascribed to the repulsions between electron pairs in the valence shell. The purpose of this chapter is to extend these elementary arguments and to indicate some of the contributions that quantum theory has made to understanding why atoms form bonds and molecules adopt characteristic shapes. There are two major approaches to the calculation of molecular structure, valence bond theory (VB theory) and molecular orbital theory (MO theory). Almost all modern computational work makes use of MO theory, and we concentrate on that theory in this chapter. Valence bond theory, however, has left its imprint on the language of chemistry, and it is important to know the significance of terms that chemists use every day. The structure of this chapter is therefore as follows. First, we present VB theory and the terms it introduces. Next, we present in more detail the basic ideas of MO theory. Finally, we see how computational techniques based on MO theory pervade all current discussions of molecular structure, including the prediction of the physiological properties of therapeutic agents. Both theories of molecular structure adopt the Born–Oppenheimer approximation in which it is supposed that the nuclei, being so much heavier than an electron, move relatively slowly and may be treated as stationary while the electrons move around them. We can therefore think of the nuclei as being fixed at arbitrary locations and then

10.1 DIATOMIC MOLECULES

solve the Schrödinger equation for the electrons alone. The approximation is quite good for molecules in their electronic ground states, for calculations suggest that (in classical terms) the nuclei in H2 move through only about 1 pm while the electron speeds through 1000 pm. By invoking the Born–Oppenheimer approximation, we can select an internuclear separation in a diatomic molecule and solve the Schrödinger equation for the electrons for that nuclear separation. Then we can choose a different separation and repeat the calculation, and so on. In this way we can explore how the energy of the molecule varies with bond length and obtain a molecular potential energy curve, a graph showing how the energy of the molecule depends on the internuclear separation (Fig. 10.1). The graph is called a potential energy curve because the nuclei are stationary and contribute no kinetic energy. Once the curve has been calculated, we can identify the equilibrium bond length, Re, the internuclear separation at the minimum of the curve, and De, the depth of the minimum below the energy of the infinitely widely separated atoms. In Chapter 12 we shall also see that the narrowness of the potential well is an indication of the stiffness of the bond. Similar considerations apply to polyatomic molecules, where bond angles may be varied as well as bond lengths.

Valence bond theory In VB theory, a bond is regarded as forming when an electron in an atomic orbital on one atom pairs its spin with that of an electron in an atomic orbital on another atom. To understand why this pairing leads to bonding, we have to examine the wavefunction for the two electrons that form the bond. 10.1 Diatomic molecules There are many diatomic molecules of biological importance, including O2 (the source of oxidizing power for catabolism), N2 (the ultimate source of nitrogen for the synthesis of a host of biomolecules, including proteins and nucleic acids), and NO (a versatile carrier of biochemical messages). We need to know how bonding in these molecules determines their physical and chemical properties and hence their biological function.

We begin by considering the simplest possible chemical bond, the one in molecular hydrogen, H–H, and then see how the concepts it introduces can be extended to other diatomic molecules. (a) Formulation of the VB wavefunction

When two ground-state H atoms are far apart, we can be confident that electron 1 is in the 1s orbital of atom A, which we denote yA(1), and that electron 2 is in the 1s orbital of atom B, which we denote yB(2). It is a general rule in quantum mechanics that the wavefunction for several noninteracting particles is the product of the wavefunctions for each particle, so we can write y(1,2) = yA(1)yB(2). When the two atoms are at their bonding distance, it may still be true that electron 1 is on A and electron 2 is on B. However, an equally likely arrangement is for electron 1 to escape from A and be found on B and for electron 2 to be on A. In this case the wavefunction is y(1,2) = yA(2)yB(1). Whenever two outcomes are equally likely, the rules of quantum mechanics tell us to add together the two corresponding wavefunctions. Therefore, the (unnormalized) wavefunction for the two electrons in a hydrogen molecule is

365

Fig. 10.1 A molecular potential energy curve. The equilibrium bond length Re corresponds to the energy minimum De.

366

10 THE CHEMICAL BOND

yH–H(1,2) = yA(1)yB(2) + yA(2)yB(1)

A valence-bond wavefunction

(10.1)

This expression is the VB wavefunction for the bond in molecular hydrogen. For technical reasons related to the Pauli exclusion principle (see the following Justification), this wavefunction can exist only if the two electrons it describes have opposite spins. Bonds do not form because electrons tend to pair their spins: bonds are allowed to form when the electrons pair their spins. Justification 10.1 The role of spin pairing in VB theory

The spatial wavefunction in eqn 10.1 does not change sign when the labels 1 and 2 are interchanged: yH–H(2,1) = yA(2)yB(1) + yA(1)yB(2) = yA(1)yB(2) + yA(2)yB(1) = yH–H(1,2) According to the Pauli principle (Further information 9.3), the overall wavefunction of the molecule (the wavefunction including spin) must change sign when we interchange the labels 1 and 2. Therefore, we must multiply yA–B(2,1) by an antisymmetric spin function of the form shown in Further information 9.3. There is only one choice: yA–B(1,2) = {yA(1)yB(2) + yA(2)yB(1)} × {a(1)b(2) − b(1)a(2)} For this combination, yA–B(2,1) = −yA–B(1,2) as required. Because the spin state a(1)b(2) − b(1)a(2) corresponds to paired electron spins, we conclude that the two electron spins in the bond must be paired in order for the bond to form.

(b) The energy of interaction

The electron density in H2 according to the valence-bond model of the chemical bond and the electron densities corresponding to the contributing atomic orbitals. The nuclei are denoted by large dots on the horizontal line. Note the accumulation of electron density in the internuclear region.

Fig. 10.2

Why, though, does the VB wavefunction result in bonding? As can be seen from Fig. 10.2, as the two atoms approach each other, there is an accumulation of electron density between the two nuclei where the two atomic orbitals overlap and their amplitudes add together. The electrons that have accumulated between the nuclei attract them and the potential energy is lowered. However, this decrease in energy is counteracted by an increase in energy from the Coulombic repulsion between the two positively charged nuclei. At intermediate internuclear separations the attraction dominates the internuclear repulsion, but at very short distances the repulsion dominates the attraction and the total energy rises above that of the widely separated atoms. Qualitatively at least, we see that this description leads to a molecular potential energy curve like that depicted in Fig. 10.1 and hence accounts for the existence of a bond. To test the model quantitatively we calculate the energy of a molecule for a series of internuclear separations by substituting the VB wavefunction into the Schrödinger equation for the molecule and calculate the corresponding values of the energy. When this energy is plotted against R, we do indeed get a curve very much like that shown in Fig. 10.1, although numerically the agreement between the calculated and experimental bond length and depth of the well is not very good. (c) s and p bonds

Because the wavefunction in eqn 10.1 is built from two H1s orbitals we can expect the overall distribution of the electrons in the molecule to be sausage shaped

10.2 POLYATOMIC MOLECULES

(Fig. 10.3). A VB wavefunction with cylindrical symmetry around the internuclear axis is called a s bond. It is so called because, when viewed along the bond, it resembles a pair of electrons in an s orbital (and s, sigma, is the Greek equivalent of s). All VB wavefunctions are constructed in a similar way, by using the atomic orbitals available on the participating atoms. In general, therefore, the (unnormalized) VB wavefunction for an A–B bond has the form given in eqn 10.1 with the two contributing wavefunctions the atomic orbitals that are being used to form the bond (for instance, the 2p orbitals of carbon atoms). We can use a similar description for molecules built from atoms that contribute more than one electron to the bonding. For example, to construct the VB description of N2, we consider the valence-electron configuration of each atom, which is 2s22p1x 2p1y 2p1z. It is conventional to take the z-axis to be the internuclear axis, so we can imagine each atom as having a 2pz orbital pointing toward a 2pz orbital on the other atom, with the 2px and 2py orbitals perpendicular to the axis (Fig. 10.4). Each of these p orbitals is occupied by one electron, so we can think of bonds as being formed by the merging of matching orbitals on neighbouring atoms and the pairing of the electrons that occupy them. We get a cylindrically symmetric s bond from the merging of the two 2pz orbitals and the pairing of the electrons they contain. The remaining N2p orbitals cannot merge to give s bonds because they do not have cylindrical symmetry around the internuclear axis. Instead, the 2px orbitals merge and the two electrons pair to form a p bond, so called because, viewed along the internuclear axis, it resembles a pair of electrons in a p orbital (and p is the Greek equivalent of p). Similarly, the 2py orbitals merge and their electrons pair to form another p bond. In general, a p bond arises from the merging of two p orbitals that approach side by side and the pairing of the electrons that they contain. It follows that the overall bonding pattern in N2 is a s bond plus two p bonds (Fig. 10.5), which is consistent with the Lewis structure :N≡N: in which the atoms are linked by a triple bond.

367

In the valence bond theory, a s bond is formed when two electrons in orbitals on neighboring atoms, as in (a), pair and the orbitals merge to form a cylindrical electron cloud, as in (b).

Fig. 10.3

Self-test 10.1 Describe the VB ground state of an O2 molecule.

Answer: One s(O2pz,O2pz) bond and one p(O2px,O2px) bond. See Case study 10.1 for an important comment.

10.2 Polyatomic molecules To understand the role of molecules in the processes of life, including self-assembly, metabolism, and self-replication, we need to extend the discussion to include the electronic structures and shapes of polyatomic molecules, ranging in size from H2O to DNA.

The ideas we have introduced so far are easily extended to polyatomic molecules. Each s bond in a polyatomic molecule is formed by the merging of orbitals with cylindrical symmetry about the internuclear axis and the pairing of the spins of the electrons they contain. Likewise, each p bond (if there is one) is formed by pairing electrons that occupy atomic orbitals of the appropriate symmetry. The description of the electronic structure of H2O will make this clear, but also bring to light a deficiency of the theory. The valence electron configuration of an O atom is 2s22p2x 2p1y 2p1z. The two unpaired electrons in the O2p orbitals can each pair with an electron in a H1s orbital, and each combination results in the formation of a s bond (each bond

Fig. 10.4 The bonds in N2 are built by allowing the electrons in the N2p orbitals to pair. However, only one orbital on each atom can form a s bond: the orbitals perpendicular to the axis form p bonds.

368

10 THE CHEMICAL BOND

has cylindrical symmetry about the respective O–H internuclear axis, Fig. 10.6). Because the 2py and 2pz orbitals lie at 90° to each other, the two s bonds they form also lie at 90° to each other. We predict, therefore, that H2O should be an angular (‘bent’) molecule, which it is. However, the model predicts a bond angle of 90°, whereas the actual bond angle is 104°. Clearly, the VB model needs to be improved. A further major deficiency becomes apparent as soon as we apply these arguments to carbon. The ground state valence configuration of a carbon atom is 2s22p1x 2p1y , which suggests that it should be capable of forming only two bonds, not the four bonds that are so characteristic of this element. Self-test 10.2 Give a VB description of NH3, and predict the bond angle of the molecule on the basis of this description. The experimental bond angle is 107°.

Answer: Three s(N2p,H1s) bonds; 90°.

(a) Promotion

Fig. 10.5 The electrons in the 2p orbitals of two neighboring N atoms merge to form s and p bonds. The electrons in the N2pz orbitals pair to form a bond of cylindrical symmetry. Electrons in the N2p orbitals that lie perpendicular to the axis also pair to form two p bonds.

Two modifications solve both deficiencies of VB theory. They are both based on the fact that it might be appropriate to invest energy initially in order to achieve a greater overall lowering of energy by allowing bond angles to change and stronger and perhaps more bonds to form. Suppose we imagine that a valence electron is promoted from a full atomic orbital to an empty atomic orbital. In carbon, with ground state configuration 2s22p1x 2p1y , for example, the promotion of a 2s electron to a 2p orbital leads to the configuration 2s12p1x 2p1y 2p1z, with four unpaired electrons in separate orbitals. These electrons may pair with four electrons in orbitals provided by four other atoms (such as four H1s orbitals if the molecule is CH4), and as a result the atom can form four s bonds. Promotion is worthwhile if the energy it requires can be more than recovered in the greater strength or greater number of bonds that can be formed. We should not think of the atom as making an initial transition to an excited state: promotion is just a way of analyzing the electron rearrangement that takes place as bonds form and achieve the lowest possible energy. We can now see why tetravalent carbon is so common. The promotion energy of carbon is small because the promoted electron leaves a doubly occupied 2s orbital and enters a vacant 2p orbital, hence significantly relieving the electron– electron repulsion it experiences in the former. Furthermore, the energy required for promotion is more than recovered by the atom’s ability to form four bonds in place of the two bonds of the unpromoted atom. (b) Hybridization

Promotion, however, appears to imply the presence of three s bonds of one type (in CH4, from the merging of H1s and C2p orbitals) and a fourth s bond of a distinctly different type (formed from the merging of H1s and C2s). It is well known, however, that all four bonds in methane are exactly equivalent in terms of both their chemical and their physical properties (their lengths, strengths, and stiffness). This problem is overcome in VB theory by drawing on another feature of quantum mechanics that allows the same electron distribution to be described in different ways. In this case, we can describe the electron distribution in the promoted atom either as arising from four electrons in one s and three p orbitals or as

10.2 POLYATOMIC MOLECULES

369

arising from four electrons in four different mixtures of these orbitals. Mixtures (more formally, linear combinations) of atomic orbitals on the same atom are called hybrid orbitals. We can picture them by thinking of the four original atomic orbitals, which are waves centered on a nucleus, as being like ripples spreading from a single point on the surface of a lake. These waves interfere destructively (where their amplitudes cancel) or constructively (where their amplitudes add) in different regions and give rise to four new shapes. The specific linear combinations that give rise to four equivalent hybrid orbitals are h1 = s + px + py + pz h3 = s − px + py − pz

h2 = s − px − py + pz h4 = s + px − py − pz

sp3 hybrid orbitals

(10.2a)

As a result of the constructive and destructive interference between the positive and negative regions of the component orbitals, each hybrid orbital has a large lobe pointing toward one corner of a regular tetrahedron (Fig. 10.7). Because each hybrid is built from one s orbital and three p orbitals, it is called an sp3 hybrid orbital. We can now see how the VB description of CH4 leads to a tetrahedral molecule containing four equivalent C–H bonds. It is energetically favorable (in the end, after bonding has been taken into account) for the C atom to undergo promotion. The promoted configuration has a distribution of electrons that is equivalent to one electron occupying each of four tetrahedral hybrid orbitals. Each hybrid orbital of the promoted atom contains a single unpaired electron; an H1s electron can pair with each one, giving rise to a s bond pointing in a tetrahedral direction. Because each sp3 hybrid orbital has the same composition, all four s bonds are identical apart from their orientation in space (Fig. 10.8). Hybridization is also used in the VB description of alkenes. Consider ethene (ethylene), which is not only an important industrial gas but also a hormone associated with the ripening of fruit. An ethene molecule is planar, with HCH and HCC bond angles close to 120°. To reproduce this s-bonding structure, we think of each C atom as being promoted to a 2s12p1x 2p1y 2p1z configuration. However, instead of using all four orbitals to form hybrids, we form sp2 hybrid orbitals by allowing the s orbital and two of the p orbitals to interfere. As shown in Fig. 10.9a, the three hybrid orbitals

Fig. 10.6 The bonding in an H2O molecule can be pictured in terms of the pairing of an electron belonging to one H atom with an electron in an O2p orbital; the other bond is formed likewise, but using a perpendicular O2p orbital. The predicted bond angle is 90°, which is in poor agreement with the experimental bond angle (104°).

h1 = s + 21/2py h2 = s + (32)1/2px − (12)1/2 py h3 = s − (32)1/2px − (12)1/2 py

sp2 hybrid orbitals

(10.2b)

lie in a plane and point toward the corners of an equilateral triangle. The third 2p orbital (2pz) is not included in the hybridization, and its axis is perpendicular to the plane in which the hybrids lie (Fig. 10.9b). The coefficients 21/2, etc., in the hybrids have been chosen to give the correct directional properties of the hybrids. The sp2-hybridized C atoms each form three s bonds with either the h1 hybrid of the other C atom or with the H1s orbitals. The s framework therefore consists of bonds at 120° to each other. Moreover, provided the two CH2 groups lie in the same plane, the two electrons in the unhybridized C2pz orbitals can pair and form a p bond (Fig. 10.10). The formation of this p bond locks the framework into the planar arrangement, for any rotation of one CH2 group relative to the other leads to a weakening of the p bond (and consequently an increase in energy of the molecule).

The 2s and three 2p orbitals of a carbon atom hybridize, and the resulting hybrid orbitals point toward the corners of a regular tetrahedron.

Fig. 10.7

370

10 THE CHEMICAL BOND

The valence bond description of the structure of CH4. Each s bond is formed by the pairing of an electron in an H1s orbital with an electron in one of the hybrid orbitals shown in Fig. 10.7. The resulting molecule is regular tetrahedral. Fig. 10.8

(a) Trigonal planar hybridization is obtained when an s and two p orbitals are hybridized. The three lobes lie in a plane and make an angle of 120° to each other. (b) The remaining p orbital in the valence shell of an sp2-hybridized atom lies perpendicular to the plane of the three hybrids. Fig. 10.9

Fig. 10.10 The valence bond description of the structure of a carbon–carbon double bond, as in ethene. The electrons in the two sp2 hybrids that point toward each other pair and form a s bond. Electrons in the two p orbitals that are perpendicular to the plane of the hybrids pair and form a p bond. The electrons in the remaining hybrid orbitals are used to form bonds to other atoms (in ethene itself, to H atoms).

A similar description applies to a linear ethyne (acetylene) molecule, H–C≡C–H. Now the carbon atoms are sp hybridized, and the s bonds are built from hybrid atomic orbitals of the form h1 = s + pz

Fig. 10.11 The electronic structure of ethyne (acetylene). The electrons in the two sp hybrids on each atom pair to form s bonds either with the other C atom or with an H atom. The remaining two unhybridized 2p orbitals on each atom are perpendicular to the axis: the electrons in corresponding orbitals on each atom pair to form two p bonds. The overall electron distribution is cylindrical.

h2 = s − p z

sp hybrid orbitals

(10.2c)

The two hybrids lie along the z-axis. The electrons in them pair either with an electron in the corresponding hybrid orbital on the other C atom or with an electron in an H1s orbital. Electrons in the two remaining p orbitals on each atom, which are perpendicular to the molecular axis, pair to form two perpendicular p bonds (as in Fig. 10.11). It is possible to form hybrid orbitals with intermediate proportions of atomic orbitals. For example, as more p-orbital character is included in an sp-hybridization scheme, the hybridization changes toward sp2 and the angle between the hybrids changes from 180° for pure sp hybridization to 120° for pure sp2 hybridization. If the proportion of p character continues to be increased (by reducing the proportion of s orbital), then the hybrids eventually become pure p orbitals at an angle of 90° to each other (Fig. 10.12). Hybridization schemes involving d orbitals (Table 10.1) are often invoked to account for (or at least be consistent with) other molecular geometries but are not commonly invoked in biology. Regardless of the types of orbitals used, an important point is that: The hybridization of N atomic orbitals always results in the formation of N hybrid orbitals.

10.2 POLYATOMIC MOLECULES

Table 10.1

371

Hybrid orbitals

Number

Shape

Hybridization*

2

Linear

sp

3

Trigonal planar

sp2

4

Tetrahedral

sp3

5

Trigonal bipyramidal

sp3d

6

Octahedral

sp3d2

*Other combinations are possible.

Hybridization accounts for—or at least is consistent with—the structure of H2O, with its bond angle of 104°. Each O–H s bond is formed from an O atom hybrid orbital with a composition that lies between pure p (which would lead to a bond angle of 90°) and pure sp2 (which would lead to a bond angle of 120°). The actual bond angle and hybridization adopted are found by calculating the energy of the molecule as the bond angle is varied and looking for the angle at which the energy is a minimum. Example 10.1

Bonding in the peptide group

Use VB theory to describe the CO, CN, and NH bonds of the peptide group based on the structure shown in (1). Strategy To calculate the number of hybrid orbitals, we note that each orbital can hold either one or two electrons. If it contains one electron, the orbital is ready to make a s bond with an orbital on another atom. If it contains a pair of electrons, then it does not participate in bonding but acts as a lone pair. It follows that the number of hybrid orbitals on an atom is equal to the sum of the number of s bonds to the atom and the number of lone pairs on the atom. Unhybridized p orbitals can participate in p bonds, as described in Section 10.4. As noted in Section 10.2, a double bond consists of a s and a p bond. Solution The O atom is sp2 hybridized because it has two lone pairs and makes

a s bond with the C atom. The C atom is sp2 hybridized because it makes three s bonds: one with the O atom, one with the Ca1 atom, and one with the N atom. The N atom is sp3 hybridized because it has one lone pair and makes three s bonds: one with the H atom, one with the C atom, and one with the Ca2 atom. We can infer that the CO group has a s bond between Csp2 and Osp2 hybrid orbitals and a p bond between unhybridized C2pz and O2pz orbitals (where again we have taken the z-axis to be perpendicular to the plane containing the hybrid orbitals). The CN group has a s bond between Csp2 and Nsp3 hybrid orbitals. Finally, the NH group has a s bond between a Nsp3 hybrid orbital and a H1s atomic orbital. This pattern of hybridizations is summarized in Fig. 10.13; but read on! Self-test 10.3 Estimate the values of the Ca1CN and CNCa2 bond angles for

the structure shown in (1). Answer: 120°, cB:

Bonding orbitals c A2 > c B2

Antibonding orbitals c B2 > c A2

Figure 10.33 shows a schematic representation of this point. These features of polar bonds can be illustrated by considering the N–H bond in the peptide group (1). The electronegativity of N is greater than that of H, so we expect a polar bond with the charge distribution d−N–Hd+. For the purposes of illustrating concepts and expressing this polarity in terms of molecular orbitals, we treat the NH fragment in isolation, disregarding its interactions with other atoms in the group. The general form of the molecular orbitals of the NH fragment is y = cHyH + cNyN, where yH is an H1s orbital and yN is an N2pz orbital. Because the ionization energy of a hydrogen atom is 13.6 eV, we know that the energy of the H1s orbital is −13.6 eV. As usual, the zero of energy is the infinitely separated electron and proton (Fig. 10.34). Similarly, from the ionization energy of nitrogen, which is 14.5 eV, we know that the energy of the N2pz orbital is −14.5 eV, about 0.9 eV lower than the H1s orbital. It follows that the bonding s orbital in NH is mainly N2pz and the antibonding s orbital is mainly H1s orbital in character. The two electrons in the bonding orbital are most likely to be found in the N2pz orbital, so there is a partial negative charge on the N atom and a partial positive charge on the H atom. A systematic way of finding the coefficients in the linear combinations is to solve the Schrödinger equation and to look for the values of the coefficients that

A schematic representation of the relative contributions of atoms of different electronegativities to bonding and antibonding molecular orbitals. In the bonding orbital, the more electronegative atom makes the greater contribution (represented by the larger sphere), and the electrons of the bond are more likely to be found on that atom. The opposite is true of an antibonding orbital. A part of the reason why an antibonding orbital is of high energy is that the electrons that occupy it are likely to be found on the more electropositive atom.

Fig. 10.33

Fig. 10.34 The atomic orbital energy levels of H and N atoms and the molecular orbitals they form. The bonding orbital has predominantly N atom character and the antibonding orbital has predominantly H atom character. Energies are in electronvolts.

386

10 THE CHEMICAL BOND

result in the lowest energy. For NH, the lowest energy is obtained for the orbital1 y = 0.54yH + 0.84yN. We see that indeed the N2pz orbital does make the greater contribution to the bonding s orbital. Self-test 10.10 What percentage of its time does a electron in the NH frag-

ment spend in a N2pz orbital? Answer: 71 per cent (= (0.84)2 × 100%)

Case study 10.2

The biochemistry of NO

Nitric oxide (nitrogen monoxide, NO) is a small molecule that diffuses quickly between cells, carrying chemical messages that help initiate a variety of processes, such as regulation of blood pressure, inhibition of platelet aggregation, and defense against inflammation and attacks to the immune system. Initially there was much opposition to the suggestion that such a small reactive molecule could be biologically relevant, but in due course the proposal was recognized by the award of the Nobel Prize for Medicine in 1998 (to R.F. Furchgott, L.J. Ignarro, and F. Murad). The molecule is synthesized from the amino acid arginine in a series of reactions catalyzed by nitric oxide synthase and requiring O2 and NADPH. To gain insight into the biochemistry of NO, we need to consider its electronic structure. Figure 10.35 shows the bonding scheme in NO and illustrates a number of points we have made about heteronuclear diatomic molecules. The ground configuration is 1s22s23s21p42p1. (The g,u designation is not applicable because the molecule is heteronuclear, and we are numbering each species of orbital in sequence or increasing energy.) The 3s and 1p orbitals are predominantly of O character because that is the more electronegative element. The highest occupied molecular orbital (hom*o) is 2p, contains one electron, and has more N character than O character. It follows that NO is a radical with an unpaired electron that can be regarded as localized more on the N atom than on the O atom. The lowest unoccupied molecular orbital (LUMO) is 4s, which is also localized predominantly on N. Because NO is a radical, we expect it to be reactive. Its half-life is estimated at approximately 1–5 s, so it needs to be synthesized often in the cell. As we saw in Case study 10.1, there is a biochemical price to be paid for the reactivity of biological radicals. Like O2, NO participates in some reactions that are not beneficial to the cell. Indeed, the radicals O2− and NO combine to form the peroxynitrite ion (5): NO + O2− → ONOO− The peroxynitrite ion is a reactive oxygen species that damages proteins, DNA, and lipids, possibly leading to heart disease, amyotrophic lateral sclerosis (Lou Gehrig’s disease), Alzheimer’s disease, and multiple sclerosis. We note that the structure of the ion is consistent with the bonding scheme of Fig. 10.35: because the unpaired electron in NO is slightly more localized on the N atom, we expect that atom to form a bond with an O atom from the O 2− ion.

The molecular orbital energy level diagram for NO.

Fig. 10.35

1

The values of the coefficients are best found by using software of the kind described in Section 10.9. For the purpose of this illustration, we are ignoring overlap between the atomic orbitals.

10.7 HÜCKEL THEORY

387

10.6 The structures of polyatomic molecules Polyatomic molecules are the building blocks of living organisms, and to understand their electronic structures we need to use MO theory; by doing so, we shall come to understand the unique role of carbon.

The bonds in polyatomic molecules are built in the same way as in diatomic molecules, the only differences being that we use more atomic orbitals to construct the molecular orbitals and these molecular orbitals spread over the entire molecule, not just the adjacent atoms of the bond. In general, a molecular orbital is a linear combination of all the atomic orbitals of all the atoms in the molecule. In H2O, for instance, the atomic orbitals are the two H1s orbitals, the O2s orbital, and the three O2p orbitals (if we consider only the valence shell). From these six atomic orbitals we can construct six molecular orbitals that differ in energy. The lowest-energy, most strongly bonding orbital has the least number of nodes between adjacent atoms. The highest-energy, most strongly antibonding orbital has the greatest numbers of nodes between neighboring atoms (Fig. 10.36). According to MO theory, the bonding influence of a single electron pair is distributed over all the atoms, and each electron pair (the maximum number of electrons that can occupy any single molecular orbital) helps to bind all the atoms together. In the LCAO approximation, each molecular orbital is modeled as a linear combination of atomic orbitals of matching symmetry, with atomic orbitals contributed by all the atoms in the molecule. Thus, a typical molecular orbital in H2O constructed from H1s orbitals (denoted yA and yB) and O2s and O2py and O2pz orbitals will have the composition y = c1yA + c2yB + c3yO2s + c4yO2p + c5yO2p y

z

(10.10)

The O2px orbital (with x perpendicular to the molecular frame) does not contribute because it has the wrong symmetry to overlap with the H1s orbitals. Because five atomic orbitals are being used to form the LCAO, there are five molecular orbitals of this kind: the lowest-energy (most bonding) orbital will have no internuclear nodes and the highest-energy (most antibonding) orbital will have a node between each pair of neighboring nuclei. 10.7 Hückel theory Many biological systems, such as those responsible for photosynthesis, vision, and the colors of vegetation, consist of molecules with conjugated p-electron systems. We need a simple way to construct their molecular orbitals and assess their energies.

An important example of the application of MO theory is to the orbitals that may be formed from the p orbitals perpendicular to a molecular plane, such as that of the phenyl ring of the amino acid phenylalanine. A computational scheme was proposed by Erich Hückel and provides a simple way of establishing the molecular orbitals of p-electron systems, especially hydrocarbons such as ethene, benzene, and their derivatives. A common procedure is to treat the s-bonding framework using the language of VB theory, and to treat the p-electron system separately by MO theory. We use that approach here. (a) Ethene

Each carbon atom in ethene, CH2=CH2, is regarded as sp2 hybridized and forming C–C and C–H s-bonds at 120° to each other by spin-pairing and either (Csp2,Csp2)or (Csp2,H1s)-orbital overlap (note the VB language). The unhybridized C2pz

Schematic form of the molecular orbitals of H2O and their energies. Fig. 10.36

388

10 THE CHEMICAL BOND

orbitals perpendicular to the s-framework (yA and yB), each of which is occupied by a single electron, are then used to construct molecular orbitals (Fig. 10.37): y = cAyA + cByB

(10.11)

We show in the following Justification that to find the energies and coefficients of the two molecular orbitals that can be formed from these two atomic orbitals we need to solve the following simultaneous equations: (HAA − ESAA)cA + (HAB − ESAB)cB = 0

Secular equations for ethene

(HBA − ESBA)cA + (HBB − ESBB)cB = 0

(10.12)

In the context of MO theory, these simultaneous equations are called secular equations. The HJK are expressions that include various contributions to the energy, including the repulsion between electrons and their attractions to the nuclei; the SJK are the overlap integrals between orbitals on atoms J and K. Justification 10.3 The secular equations

The bonding and antibonding p molecular orbitals of ethene and their energies. Fig. 10.37

We begin by substituting eqn 10.11 into the Schrödinger equation written in the form Ĥy = Ey: cAĤyA + cBĤyB = cAEyA + cBEyB Then we multiply both sides by yA cAyAĤyA + cByAĤyB = cAyAEyA + cByAEyB and integrate over all space (with the term dt denoting an infinitesimal volume element in three dimensions; in cartesian coordinates, dt = dxdydz):

cA yAĤyAdt + cB yAĤyBdt = cAE yAyAdt + cBE yAyBdt E is a constant, so we have been able to take it outside the integral. Now we write

HAA = yAĤyAdt

HAB = yAĤyBdt

SAA = yAyAdt

SAB = yAyBdt

The preceding equation then becomes cAHAA + cBHAB = cAESAA + cB ESAB which is easy to rearrange into the first of eqn 10.12. If instead of multiplying through by yA we multiply by yB, we obtain the second of eqn 10.12. To simplify the solution of the secular equations Hückel introduced the following drastic approximations: • All HJJ are set equal to a single quantity a called the Coulomb integral. • All HJK are set equal to zero unless atoms J and K are adjacent, when it is set equal to a single quantity b (a negative quantity) called the resonance integral. • All SJJ are set equal to 1 and all SJK are set equal to 0 whether or not J and K are adjacent.

10.7 HÜCKEL THEORY

389

Mathematical toolbox 10.1 Simultaneous equations and determinants

dx + ey + fz = 0

Two simultaneous equations of the form

gx + hy + iz = 0

ax + by = 0 cx + dy = 0

have a solution only if

have solutions only if the ‘determinant’ of the coefficients is equal to zero. In this case we write 4a 4c

b4 =0 d4

b e h

c4 f 4=0 i4

This 3 × 3 determinant expands as follows:

where the term on the left is the determinant and has the following meaning: 4a 4c

4a 4d 4g

b4 = ad − bc d4

4a 4d 4g

b e h

c4 e f 4 = a44 h i4

ax + by + cz = 0

With these ‘Hückel approximations’ the secular equations become bcA + (a − E)cB = 0

Hückel approximation for ethene

(10.13a)

As set out in Mathematical toolbox 10.1, these two simultaneous equations have a solution only if the secular determinant vanishes: 4a − E 4 b

b 4 = (a − E)2 − b2 = 0 a − E4

Hückel secular determinant for ethene

(10.13b)

Hückel energies for ethene

(10.13c)

This condition is satisfied if E=a±b

When each value is substituted into eqn 10.13a, we find: For E = a + b

cA = cB, so y = cA(yA + yB)

For E = a − b

cA = −cB, so y = cA(yA − yB)

f 4 4d +c i 4 4g

e4 h4

Note the alternation in signs for successive columns. The 2 × 2 determinants then expand like the one above.

Three simultaneous equations of the form

(a − E)cA + bcB = 0

f 4 4d −b i 4 4g

(Remember that b < 0, so E = a + b is the lower energy of the two.) These energies and orbitals are represented in Fig. 10.37: they will be recognized as the bonding and antibonding combinations of the C2pz atomic orbitals. The value of the one unknown, cA, is found by ensuring that each orbital is normalized, but we do not need its explicit value. Because there are two electrons to be accommodated, both enter the lower energy orbital and contribute 2a + 2b to the energy of the molecule. We can also infer that the energy needed to excite a p electron to the antibonding combination is 2| b |. A typical value of b in hydrocarbons is about −2.4 eV, or −230 kJ mol−1.

390

10 THE CHEMICAL BOND

Self-test 10.11 Write down the Hückel secular determinant for butadiene.

4a − E 4 b Answer: 4 0 4 0

b a−E b 0

0 b a−E b

0 4 0 4 b 4 a − E4

(b) Benzene

Exactly the same procedure can be used for benzene, C6H6. Each C atom is regarded as sp2 hybridized (note the VB language again) and forms a planar hexagonal framework of s bonds (Fig. 10.38). There is an unhybridized C2pz orbital on each atom perpendicular to the ring from which we form molecular orbitals. From these six atomic orbitals we construct six molecular orbitals of the form y = cAyA + cByB + cCyC + cDyD + cEyE + cF yF The orbitals used to construct the molecular orbitals of benzene. Fig. 10.38

Fig. 10.39 The p orbitals of benzene and their energies. The lowest-energy orbital is fully bonding between neighboring atoms, but the uppermost orbital is fully antibonding. The two pairs of doubly degenerate molecular orbitals have an intermediate number of internuclear nodes. As usual, different colors represent different signs of the wavefunction.

(10.14)

Then we set up the six simultaneous equations for the coefficients and the corresponding 6 × 6 secular determinant and apply the Hückel approximations. Its form resembles that for cyclobutadiene in Exercise 10.33, but with six rows and six columns. Full-frontal attack on it to determine the six values of E is rather tedious, especially as there are procedures that make use of symmetry that greatly simplifies the solution. As should be verified, the energies and the corresponding (unnormalized) molecular orbitals obtained are as follows (Fig. 10.39):

10.7 HÜCKEL THEORY

Energy Highest (most antibonding) a − 2b a−b a−b a+b a+b a + 2b Lowest (most bonding)

Orbital y = yA − yB + yC − yD + yE − yF y = 21/2yA − yB − yC + 21/2yD − yE − yF y = yB − yC + yE − yF y = 21/2yA + yB + yC − 21/2yD − yE − yF y = yB + yC − yE − yF y = yA + yB + yC + yD + yE + yF

Note that the lowest-energy, most bonding orbital has no internuclear nodes. It is strongly bonding because the constructive interference between neighboring p orbitals results in a good accumulation of electron density between the nuclei (but slightly off the internuclear axis, as in the p bonds of diatomic molecules). In the most antibonding orbital the alternation of signs in the linear combination results in destructive interference between neighbors, and the molecular orbital has a nodal plane between each pair of neighbors, as shown in the illustration. The four intermediate orbitals form two doubly degenerate pairs, one net bonding and the other net antibonding. There are six electrons to be accommodated (one is supplied by each C atom), and they occupy the lowest three orbitals in Fig. 10.39. The resulting electron distribution is like a double donut. It is an important feature of the configuration that the only molecular orbitals occupied have a net bonding character, for this is one contribution to the stability (in the sense of low energy) of the benzene molecule. It may be helpful to note the similarity between the molecular orbital energy level diagram for benzene and that for N2 (see Fig. 10.31): the strong bonding, and hence the stability, of benzene and of the phenyl ring in aromatic amino acids is an echo of the strong bonding in the nitrogen molecule. A feature of the molecular orbital description of benzene is that each molecular orbital spreads either all around or partially around the C6 ring. That is, p bonding is delocalized, and each electron pair helps to bind together several or all of the C atoms. The delocalization of bonding influence is a primary feature of MO theory that we shall use time and again when discussing conjugated systems, such as those found in selected amino acid side chains (phenylalanine, tyrosine, histidine, and tryptophan), the purine and pyrimidine bases in nucleic acids, the heme group, and the pigments involved in photosynthesis and vision. The stabilization of the benzene molecule due to delocalization can be expressed quantitatively. If the six p electrons occupied three localized ethene-like orbitals, then their energy would be 3 × (2a + 2b) = 6a + 6b. However, their energy in benzene is 2(a + 2b) + 4(a + b) = 6a + 8b. The delocalization energy, the difference of these two energies, is therefore 2b, or about −460 kJ mol−1. Case study 10.3

The unique role of carbon in biochemistry

Now we can take stock of our knowledge of chemical bonding and continue the discussion in Section 9.12 of the properties of carbon that make it uniquely suitable for building complex biological structures. Among the elements of Period 2, carbon has an intermediate electronegativity, so it can form covalent bonds with many other elements, such as hydrogen, nitrogen, oxygen, sulfur, and, more importantly, other carbon atoms. Furthermore, because it has four valence electrons, carbon atoms can form chains and

391

392

10 THE CHEMICAL BOND

rings containing single, double, or triple C–C bonds. Such a variety of bonding options leads to the intricate molecular architectures of proteins, mucleic acids, and cell membranes. Bonds need to be sufficiently strong to maintain the structure of the cell yet need to be susceptible to dissociation and rearrangement during chemical reactions. To get a sense of the uniqueness of the C–C bond, consider the energetics of the N–N and Si–Si bonds. The comparison is useful because nitrogen and silicon are neighbors of carbon in the periodic table and are abundant elements on Earth. The atomic radius of silicon is greater than that of carbon, so we expect an Si–Si bond to be longer than a C–C bond and the orbital overlap to be weaker. The atomic radius of nitrogen is smaller than that of carbon, but the length and energy of an N–N bond, such as that in hydrazine (H2N–NH2), are influenced by the fact that sp3 hybridization leaves lone pairs on the nitrogen atoms. These lone pairs repel each other, making an N–N bond weaker than a C–C bond. A C–C bond is sufficiently strong that it can be used as a motif for the formation of robust cellular components. Weaker bonds, such as C–N and C–O, are more reactive, breaking during catabolism and re-forming during anabolism.

10.8 d-Metal complexes Ions of the d metals participate in biological electron transfer (Chapter 8), the binding and transport of O2, and the mechanisms of action of many enzymes. To understand the biochemical function of d metal atoms, we need to develop a theory for the formation of bonds between them and biological molecules.

In Chapter 9 we saw that the d-metal ions typically have an incomplete shell of d electrons. These electrons play a special role in d-metal complexes, giving rise to their biochemical activity, their colors, and their magnetic properties. There are two approaches: one, crystal field theory, is a simple approach that accounts for the general structures of complexes; the other, ligand field theory, is an adaptation of MO theory and is much more powerful. (a) Crystal field theory

In an octahedral d-metal complex six identical ions or molecules, the ligands, are at the vertices of a regular octahedron, with the metal atom at its center. An example of this arrangement is the complex [Fe(OH2)6]2+ (6), in which the Fe2+ ion, a good Lewis acid, is surrounded by six H2O molecules, which are good Lewis bases. In crystal field theory, each ligand is regarded as a point negative charge that repels the d electrons of the central ion. The energy of the entire system decreases when the six ligands approach the central metal cation on account of the favorable Coulomb interactions between its positive charge and the lone electron pairs of the ligands. However, because the point charges representing the ligands repel the d electrons present on the atom, there is also a relatively small modification of that overall decrease in energy. Figure 10.40 shows that the five d orbitals of the central metal ion fall into two groups: dx −y and dz point directly toward the ligand positions, whereas dxy, dyz, and dzx point between them. According to crystal-field theory, an electron occupying an orbital of the former group has a less favorable potential energy than when 2

2

2

10.8 D-METAL COMPLEXES

it occupies any of the three orbitals of the other group, and so the d orbitals split into two sets (7): a triply degenerate set comprising the dxy, dyz, and dzx orbitals and labeled t2g and a doubly degenerate set comprising the dx −y and dz orbitals and labeled eg. (The notation is derived from group theory, the mathematical theory of symmetry.) The energy difference between the two sets of orbitals is called the crystal-field splitting and denoted DO. The splitting is about 10 per cent of the total energy of interaction of the ligands with the central metal ion. If we know the number of electrons supplied by the central ion, then we can use the building-up principle to arrive at its electronic configuration by letting the electrons occupy the d orbitals so as to achieve the lowest possible energy bearing in mind, as usual, the Pauli exclusion principle. If the ion has one d electron, as in the case of Ti3+, the configuration of the complex is t12g. For two and three d elec2 3 trons, the configurations are, respectively, t 2g (as in V3+) and t 2g (as in Cr3+). According to Hund’s rule, these electrons can have parallel spins (Fig. 10.41). A decision now has to be made: the fourth d electron (as in Mn3+) can occupy either the half-filled t2g set of orbitals or the empty eg orbitals. The advantage of the former arrangement is that the t2g orbitals lie lower in energy than the eg orbitals; the disadvantage is the significant electron–electron repulsions in a doubly filled orbital. The disadvantage of the second arrangement, which gives the configura3 1 eg, is the necessity of occupying a high-energy orbital, but the advantage is tion t 2g less electron–electron repulsion. This advantage is more important than might be 3 1 expected because all four electrons may have parallel spins in t 2g eg and Hund’s rule indicates that parallel spins are energetically favorable. 3 1 4 Which configuration, t2g eg or t2g , actually occurs depends on a variety of factors, an important one being the magnitude of the crystal-field splitting. If DO is large, 4 configuration, with its spin-paired arrangement, is favored. Such a molecule the t2g is called a low-spin complex (Fig. 10.42a). If DO is small, the advantage of minimizing electron–electron repulsion outweighs the disadvantage of occupying a 3 1 high-energy orbital and the t 2g eg configuration is expected, giving rise to a highspin complex with the maximum number of unpaired electrons (Fig. 10.42b). 2

2

2

393

Fig. 10.40 The classification of d orbitals in an octahedral environment.

Fig. 10.41 The occupation of energy levels in a d3 octahedral complex.

394

10 THE CHEMICAL BOND

Example 10.3

Low- and high-spin complexes of Fe(II) in hemoglobin

We saw in Case study 4.1 that O2 binds to and is transported through the body by the protein hemoglobin (Atlas P7), which contains the heme group (Atlas P6), a complex of the Fe2+ ion. Deoxygenated heme is a high-spin complex that makes a transition to a low-spin complex on binding O2 as a ligand of the Fe2+ ion. Predict the number of unpaired electrons in deoxygenated and oxygenated heme. Strategy Determine the electronic configuration of the Fe2+ ion according to

Fig. 10.42 The energy separation DO controls the electronic configuration of an octahedral d-metal complex, as shown here for a metal with four d electrons. (a) If DO is large, a low-spin complex results with a t 42g configuration. (b) If DO is small, a high-spin complex is favored with a t 32geg1 configuration.

the rules described in Section 9.11. Then apply the building-up principle to the two sets of d orbitals, allowing the maximum number of unpaired electrons to be the dominant factor in high-spin complexes, but not in low-spin complexes. Solution The ground-state electron configuration of an Fe atom is [Ar]3d64s2,

so the configuration of an Fe2+ ion is [Ar]3d6. In deoxygenated heme, a highspin complex, DO is small, so the first five electrons enter the t2g and eg orbitals with parallel spins. The sixth electron occupies the t2g orbital and must pair. 4 2 The configuration is, therefore, t 2g eg and there are four unpaired electrons. In oxygenated heme, DO is large and all six electrons occupy the t2g orbitals. To do so, they must have paired spins. The configuration is t 62g and there are no unpaired electrons. Self-test 10.12 Cobalt is present in vitamin B12. Predict the number of

unpaired electron spins in high-spin and low-spin complexes of a Co2+ ion. Answer: 3 and 1, respectively

(b) Ligand-field theory: s bonding

Crystal-field theory has a major deficiency: it attempts to ascribe the bonding of the complex to Coulombic interactions between d electrons localized on a central metal ion and electron pairs localized in orbitals confined to the ligands. However, we know from our discussion of MO theory that molecular orbitals spread over both metal atoms and ligands. Ligand-field theory develops this point of view in terms of molecular orbitals. It proceeds in three steps: • Identify combinations of the ligand orbitals that have symmetries that match the symmetries of the d orbitals of the central metal ion. • Form molecular orbitals by allowing overlap between these combinations and d orbitals of the same symmetry. • Use the building-up principle in the same way as in crystal-field theory. We shall represent (only for purposes of visualization) the ligand orbitals by six spheres, each occupied by two electrons (for concreteness, think of the spheres as representing the lone pairs of NH3 molecules). From these six atomic orbitals we construct the six combinations spreading over the six ligands shown in Fig. 10.43. We see that two of the six combinations have a shape that matches the two eg orbitals of the central ion, and four have the wrong shape for any net overlap with either the eg or t2g metal orbitals. As a result, only eg molecular orbitals can be formed between the d orbitals and the ligands, and as a result there are two eg bonding molecular orbitals and two eg* antibonding molecular orbitals. The three

10.8 D-METAL COMPLEXES

395

Fig. 10.43 The combinations of ligand orbitals (represented here by spheres) in an octahedral complex, shown alongside the atomic orbitals of the metal. Only the ligand orbitals labeled eg have the right shape to give nonzero overlap with eg orbitals of the metal. The metal’s t2g orbitals do not combine with the ligand orbitals.

396

10 THE CHEMICAL BOND

metal t2g orbitals are classified as non-bonding, in the sense that they do not interact to form bonding and antibonding combinations. The four remaining ligand orbitals (labeled, once again using notation that comes from group theory, as a1g and t1u in Fig. 10.43) have the appropriate symmetry to overlap with metal s and p orbitals, respectively, and form bonding and antibonding combinations with them. The energies of the full array of molecular orbitals are shown in Fig. 10.44. According to the building-up principle, we need to accommodate the appropriate number of electrons into the molecular orbitals of the complex. Each ligand provides two electrons, and the dn central ion provides n electrons, so we must accommodate 12 + n electrons. Of these electrons, four will occupy the two eg bonding molecular orbitals, eight will occupy the t1u and a1g bonding orbitals, and the remaining n electrons need to be distributed among the metal-centered t2g nonbonding orbitals and the eg* antibonding molecular orbitals. We see that there are similarities between the ligand-field and crystal-field formalisms because the n electrons contributed to the complex by the metal atom enter five orbitals split into a set of three and a set of two orbitals. The difference between the theories lies both in the source of the energy separation DO and in the spread of the eg* orbitals onto the ligands; the occurrence of low-spin and high-spin complexes is accounted for in terms of the energy splittings that result from the formation of bonding and antibonding molecular orbitals and not just in terms of metal–ligand Coulombic interactions. Fig. 10.44 The molecular orbital energy level diagram for an octahedral complex. The 12 electrons provided by the six ligands fill the lowest six orbitals, which are all bonding orbitals. The n d electrons provided by the central metal atom or ion are accommodated in the orbitals inside the box.

Fig. 10.45 The effect of p bonding on the magnitude of DO. (a) In this case, the antibonding p* orbital of the ligand is too high in energy to take part in bonding or it is absent; the interaction with the (full) p orbital of the ligand decreases DO. (b) In this case, the antibonding p* orbital of a ligand matches the metal orbital in energy; the interaction with the (empty) p* orbital of the ligand increases DO.

(c) Ligand-field theory: p bonding

So far, we have considered only ligand orbitals that point directly at the metal ion orbitals, forming s molecular orbitals. Ligand-field theory also takes into account the effects of ligand orbitals that participate in the formation of p molecular orbitals with metal ion orbitals. Figure 10.45 shows that a p orbital on the ligand perpendicular to the axis of the metal–ligand bond can overlap with one of the t2g orbitals. The resulting bonding combination lies below the energy of the original nonbonding t2g orbitals and the antibonding combination lies above them. Interactions between metal ion orbitals and ligand p orbitals can either decrease or increase DO. To see how this is so, consider the bonding schemes in Fig. 10.45. If a ligand p orbital and a nonbonding t2g orbital of the metal ion have similar energies, when they interact a likely outcome is shown in Fig. 10.45a. If the ligand p orbital supplies two electrons and the t2g orbital supplies one, then the result is a decrease in DO. On the other hand, if the metal t2g and ligand p* orbitals have

10.8 D-METAL COMPLEXES

397

similar energies, then they may interact in the manner shown in Fig. 10.45b, If the p* orbital is empty and t2g orbital supplies one, then the result is an increase in D O. Case study 10.4

Ligand-field theory and the binding of O2 to hemoglobin

Ligand-field theory provides excellent descriptions of the interactions between metal ions and ligands in metalloproteins. In this Case study, we apply the theory to an important biological process: the binding of O2 to hemoglobin. Nature makes unconscious use of ligand-field effects to pump and store oxygen throughout our bodies. Here we concentrate on hemoglobin (Hb, Atlas P7), the protein used to transport oxygen through our bodies, and myoglobin (Mb), the protein used to store oxygen in muscle tissue and to release it on demand (see also Case study 4.1). Hemoglobin is a tetramer of four myoglobinlike subunits, and each subunit, as in myoglobin, binds a single heme group, an almost flat ring-like structure with an iron atom at its center (Atlas P6). The oxygenated form of hemoglobin is called the relaxed state (R state) and the deoxygenated form is called the tense state (T state). The heme group binds oxygen when the Fe atom is present as iron(II) (Fig. 10.46). The Fe–O2 complex is held together by a s bond between an empty Fe(II) eg orbital and the full s orbital of O2 and a p bond between filled t2g orbitals on Fe(II) and the half-full p* orbitals of O2. The bound O2 molecule adopts a bent orientation with respect to the Fe atom, partly because that orientation maximizes interactions between orbitals, but also because it is consistent with the spatial constraints imposed by the arrangement of peptide residues in the pocket of the protein containing the heme group. Another important change that occurs when the Fe atom is oxygenated is the transition from an Fe(II) high-spin d6 configuration to an Fe(II) low-spin d6 configuration. This change accompanies the increase in the number of ligands of the Fe ion from five to six. In the deoxygenated form, the fifth location is taken up by the N atom of a histidine residue (His); in the oxygenated form that link remains, but the O2 molecule binds on the other side of the ring (as shown in Fig. 10.46b). The change from high spin to low spin results in a slightly smaller atom. As a result, instead of lying 60 pm above the plane of the heme ring, the Fe atom can fall back almost into the plane of the ring, and in the oxygenated form it lies only 20 pm above the plane. As it falls back, it pulls the histidine residue with it. In Case study 4.1 we discussed the thermodynamic view of the binding of O2 to hemoglobin. Now we can merge the thermodynamic and molecular views into a single model.2 When one of the subunits binds the first O2 molecule, the heme group and its ligands reorganize as described above with the further consequence that one pair of subunits rotates through 15° relative to the other pair and becomes offset by 80 pm. This realignment of two of the subunits relative to the other two disrupts an ionic His+...Asp− interaction that helps to stabilize the deoxygenated form, and as a result the partially oxygenated hemoglobin molecule is more capable of taking up the next O2 than the fully 2 J. Monod, J. Wyman, and J.-P. Changeux and later D. Koshland proposed the essential features of the model, which has been refined by structural studies with diffraction and spectroscopic techniques (discussed in Chapters 11 and 12, respectively).

Fig. 10.46 The change in molecular geometry that takes place when an O2 molecule attaches to an Fe atom in a hemoglobin molecule. (a) The deoxygenated heme group, with the Fe(II) ion in its low-spin configuration. (b) The oxygenated heme group, with the Fe(II) ion in its high-spin configuration. Note how the histidine residue is pulled into a different location by the motion of the iron atom.

398

10 THE CHEMICAL BOND

deoxygenated form was. In thermodynamic terms, the equilibrium constant for binding of the second O2 molecule is greater than the equilibrium constant for binding of the first O2 molecule. As each of the four subunits become oxygenated, the binding of O2 to the remaining deoxygenated subunits becomes successively more favorable thermodynamically. In other words, in hemoglobin there is a cooperative uptake of O2 molecules. The cooperative binding of O2 by hemoglobin is an example of an allosteric effect, in which an adjustment of the conformation of a molecule when one substrate binds affects the ease with which a subsequent substrate molecule binds. Oxygenated hemoglobin also unloads O2 cooperatively when conditions demand it. The result of cooperativity is that hemoglobin can release its O2 under conditions when myoglobin cannot, which is an ideal arrangement for a transport protein rather than a storage protein (see Case study 4.1).

Computational biochemistry Computational chemistry is now a standard part of chemical research. One major application is in pharmaceutical chemistry, where the likely pharmacological activity of a molecule can be assessed computationally from its shape and electron density distribution before expensive clinical trials are started. Commercial software is now widely available for calculating the electronic structures of molecules and displaying the results graphically. All such calculations work within the Born–Oppenheimer approximation and express the molecular orbitals as linear combinations of atomic orbitals. 10.9 Computational techniques Elaborate computational methods make reasonably accurate predictions of molecular properties, including their conformation, spectroscopic properties, and reactivity. Although these techniques tax computational resources heavily, they can be used in studies of moderately sized biological molecules.

There are two principal approaches to solving the Schrödinger equation for many-electron polyatomic molecules. In the semi-empirical methods, certain expressions that occur in the Schrödinger equation are set equal to parameters that have been chosen to lead to the best fit to experimental quantities, such as enthalpies of formation. Semi-empirical methods are applicable to a wide range of molecules with a virtually limitless number of atoms and are widely popular. In the more fundamental ab initio method, an attempt is made to calculate structures from first principles, using only the atomic numbers of the atoms present. Such an approach is intrinsically more reliable than a semi-empirical procedure. Both types of procedure typically adopt a self-consistent field (SCF) procedure, in which an initial guess about the composition of the LCAO is successively refined until the solution remains unchanged in a cycle of calculation. For example, the potential energy of an electron at a point in the molecule depends on the locations of the nuclei and all the other electrons. Initially, we do not know the locations of those electrons (more specifically, we do not know the detailed form of the wavefunctions that describe their locations, the molecular orbitals they occupy). First, then, we guess the form of those wavefunctions—we guess

10.9 COMPUTATIONAL TECHNIQUES

the values of the coefficients in the LCAO used to build the molecular orbitals— and solve the Schrödinger equation for the electron of interest on the basis of that guess. Now we have a first approximation to the molecular orbital of our electron (a reasonable estimate of the coefficients for its LCAO) and we repeat the procedure for all the other molecular orbitals in the molecule. At this stage, we have a new set of molecular orbitals, which in general will have coefficients that differ from our first guess, and we also have an estimate of the energy of the molecule. We use that refined set of molecular orbitals to repeat the calculation and calculate a new energy. In general, the coefficients in the LCAOs and the energy will differ from the new starting point. However, there comes a stage when repetition of the calculation leaves the coefficients and energy unchanged. The orbitals are now said to be self-consistent, and we accept them as a description of the molecule. (a) Semi-empirical methods

The Hückel method is a very primitive example of a semi-empirical method in which various integrals are set equal to either a or b and treated as empirical parameters; overlap integrals are ignored. The removal of the restriction of the Hückel method to planar hydrocarbon systems was achieved with the introduction of the extended Hückel theory (EHT) in about 1963. In heteroatomic non-planar systems (such as d-metal complexes) the separation of orbitals into p and s is no longer appropriate and each type of atom has a different value of HJJ (which in Hückel theory is set equal to a for all atoms). In this approximation, the overlap integrals are not set equal to zero but are calculated explicitly. Furthermore, the HJK, which in Hückel theory are set equal to b, in EHT are made proportional to the overlap integral between the orbitals J and K. Further approximations of the Hückel method were removed with the introduction of the complete neglect of differential overlap (CNDO) method, which is a slightly more sophisticated method for dealing with the terms HJK that appear in the secular equations for the coefficients. The introduction of CNDO opened the door to an avalanche of similar but improved methods and their accompanying acronyms, such as intermediate neglect of differential overlap (INDO), modified neglect of differential overlap (MNDO), and the Austin Model 1 (AM1, version 2 of MINDO). Software for all these procedures is now readily available, and reasonably sophisticated calculations can be run even on handheld computers. (b) Density functional theory

A semi-empirical technique that has gained considerable ground in recent years to become one of the most widely used techniques for the calculation of molecular structure is density functional theory (DFT). Its advantages include less demanding computational effort, less computer time, and—in some cases (particularly d-metal complexes)—better agreement with experimental values than is obtained from other procedures. The central focus of DFT is the electron density, r (rho), rather than the wavefunction y. When the Schrödinger equation is expressed in terms of r, it becomes a set of equations called the Kohn–Sham equations. As for the Schrödinger equation itself, this equation is solved iteratively and self-consistently. First, we guess the electron density. For this step it is common to use a superposition of atomic electron densities. Next, the Kohn–Sham equations are solved to obtain an initial set of orbitals. This set of orbitals is used to obtain a better

399

400

10 THE CHEMICAL BOND

A brief comment

In mathematics, when an entire function f(x) is associated with a single number, F, as when an entire wavefunction is associated with the energy of the state, the number is said to be a functional of the function and written F[ f ]. Thus, the energy E is a functional of the wavefunction and we could denote it E[y] to denote the functional dependence of the energy on the entire wavefunction. In DFT, the energy is regarded as a functional of the electron density, and written E[r].

Fig. 10.47

benzene.

The isodensity surface of

approximation to the electron density, and the process is repeated until the density and the energy are constant to within some tolerance. (c) Ab initio methods

The ab initio methods also simplify the calculations, but they do so by setting up the problem in a different manner, avoiding the need to estimate parameters by appeal to experimental data. In these methods, sophisticated techniques are used to solve the Schrödinger equation numerically. The difficulty with this procedure, however, is the enormous time it takes to carry out the detailed calculation. That time can be reduced by replacing the hydrogenic atomic orbitals used to form the LCAO by a Gaussian-type orbital (GTO) in which the exponential function e−r characteristic of actual orbitals is replaced by a sum of Gaussian functions of the form e−r (recall the relative shapes of exponential and Gaussian functions shown in Mathematical Toolkit F.2). 2

10.10 Graphical output One of the most significant developments in computational chemistry and its application to biology has been the introduction of graphical representations of molecular geometries, molecular orbitals, and electron densities.

The raw output of a molecular structure calculation is a list of the coefficients of the atomic orbitals in each molecular orbital and the energies of these orbitals. The graphical representation of a molecular orbital uses stylized shapes to represent the basis set and then scales their size to indicate the value of the coeffi cient in the LCAO. Different signs of the wavefunctions are represented by different colors (as we saw in Figs 10.36, 10.37, and 10.39). Once the coefficients are known, we can build up a representation of the electron density in the molecule by noting which orbitals are occupied and then forming the squares of those orbitals. The total electron density at any point is then the sum of the squares of the wavefunctions evaluated at that point. The outcome is commonly represented by an isodensity surface, a surface of constant total electron density (Fig. 10.47). There are several styles of representing an isodensity surface: as a solid form, as a transparent form with a ball-and-stick representation of the molecule within, or as a mesh. A related representation is a solvent-accessible surface, which is generated by plotting the location of the center of a sphere (representing a solvent molecule) that is imagined to roll across the exposed surfaces of the atoms. One of the most important aspects of a molecule other than its geometrical shape is the distribution of electric potential over its surface. A common procedure begins with calculation of the potential energy of a ‘probe’ charge at each point on an isodensity surface and interpreting its energy as an interaction with an electric potential at that point. The result is an electrostatic potential surface (an ‘elpot surface’) in which net positive potential is shown in one color and net negative potential is shown in another, with intermediate gradations of color (Fig. 10.48). 10.11 The prediction of molecular properties The results of quantum mechanical calculations are only approximate, with deviations from experimental values increasing with the size of the molecule. Therefore, one goal of computational biochemistry is to gain an insight into the trends in properties of biological molecules, without necessarily striving for ultimate accuracy.

10.11 THE PREDICTION OF MOLECULAR PROPERTIES

401

Computation is now used to explore far more than the electronic structures of molecules. We already saw in Section 1.12 that computational techniques can be used to estimate the enthalpies of formation of conformational isomers and the effect of solvent on the enthalpy of formation. (a) Electrochemical properties

Molecular orbital calculations may also be used to predict trends in electrochemical properties, such as standard potentials (Chapter 5). Several experimental and computational studies of aromatic hydrocarbons indicate that decreasing the energy of the LUMO enhances the ability of a molecule to accept an electron into the LUMO, with an attendant increase in the value of the molecule’s standard potential. The effect is also observed in quinones and flavins, co-factors involved in biological electron transfer reactions. For example, stepwise substitution of the hydrogen atoms in p-benzoquinone by methyl groups (–CH3) results in a systematic increase in the energy of the LUMO and a decrease in the standard potential for formation of the semiquinone radical:

Fig. 10.48 The electrostatic potential surfaces of (a) benzene and (b) pyridine. Note the accumulation of electron density on the N atom of pyridine at the expense of the other atoms.

The standard potentials of naturally occurring quinones are also modified by the presence of different substituents, a strategy that imparts specific functions to specific quinones. For example, the substituents in coenzyme Q are largely responsible for poising its standard potential so that the molecule can function as an electron shuttle between specific electroactive proteins in the respiratory chain (Section 5.10). (b) Spectroscopic properties

We remarked in Chapter 9 that a molecule can absorb or emit a photon of energy hc/l, resulting in a transition between two quantized molecular energy levels. The transition of lowest energy (and longest wavelength) occurs between the hom*o and LUMO. We can use calculations based on semi-empirical, ab initio, and DFT methods to correlate the hom*o–LUMO energy gap with the wavelength of absorption. For example, consider the linear polyenes shown in Table 10.3: ethene (C2H4), butadiene (C4H6), hexatriene (C6H8), and octatetraene (C8H10), all of which absorb in the ultraviolet region of the spectrum. The table also shows that, as expected, the wavelength of the lowest-energy electronic transition decreases as the energy separation between the hom*o and LUMO increases. We also see that the smallest hom*o–LUMO gap and longest transition wavelength correspond to octatetraene, the longest polyene in the group. It follows that the wavelength of the transition increases with increasing number of conjugated double bonds in linear polyenes. Extrapolation of the trend suggests that a sufficiently long linear polyene should absorb light in the visible region of the electromagnetic spectrum. This is indeed the case for b-carotene (Atlas E1), which absorbs light with l ≈ 450 nm. The ability of b-carotene to absorb visible light is part of the strategy employed by plants to harvest solar energy for use in photosynthesis (Chapter 12).

402

10 THE CHEMICAL BOND

Table 10.3

Summary of ab initio calculations and spectroscopic data for four linear

polyenes DEhom*o-LUMO/eV*

l transition /nm

18.1

163

14.5

217

12.7

252

11.6

304

1 eV = 1.602 × 10−19 J.

(c) Chemical reactivity

There are several ways in which molecular orbital calculations lend insight into reactivity. For example, electrostatic potential surfaces may be used to identify an electron-poor region of a molecule that is susceptible to association with or chemical attack by an electron-rich region of another molecule. Such considerations are important for assessing the pharmacological activity of potential drugs (Section 11.17(c)). An attractive feature of computational chemistry is its ability to model species that may be too unstable or short-lived to be studied experimentally. For this reason, quantum mechanical methods are often used to study the transition state, with an eye toward describing factors that stabilize it and increase the reaction rate. Systems as complex as enzymes are amenable to study by computational methods.

Checklist of key concepts 1. An ionic bond is formed by transfer of electrons from one atom to another and the attraction between the ions. A covalent bond is formed when two atoms share a pair of electrons. 2. In the Born–Oppenheimer approximation, nuclei are treated as stationary while electrons move around them. 3. In valence bond theory (VB theory), a bond is regarded as forming when an electron in an atomic orbital on one atom pairs its spin with that of an electron in an atomic orbital on another atom. 4. A valence-bond wavefunction with cylindrical symmetry around the internuclear axis is a s bond. A p bond arises from the merging of two p orbitals that approach side by side and the pairing of electrons that they contain.

5. Hybrid orbitals are mixtures of atomic orbitals on the same atom. In VB theory, hybridization is invoked to be consistent with molecular geometries. 6. Resonance is the superposition of the wavefunctions representing different electron distributions in the same nuclear framework. 7. In molecular orbital theory (MO theory), electrons are treated as spreading throughout the entire molecule. 8. A bonding orbital is a molecular orbital that, if occupied, contributes to the strength of a bond between two atoms. An antibonding orbital is a molecular orbital that, if occupied, decreases the strength of a bond between two atoms. 9. The building-up principle suggests procedures for constructing the electron configuration of molecules on the basis of their molecular orbital energy level diagram.

403

DISCUSSION QUESTIONS

10. When constructing molecular orbitals, we need to consider only combinations of atomic orbitals of similar energies and of the same symmetry around the internuclear axis. 11. The electronegativity of an element is the power of its atoms to draw electrons to itself when it is part of a compound. 12. In a bond between dissimilar atoms, the atomic orbital belonging to the more electronegative atom makes the larger contribution to the molecular orbital with the lowest energy. For the molecular orbital with the highest energy, the principal contribution comes from the atomic orbital belonging to the less electronegative atom. 13. Hückel theory is a simple treatment of the molecular orbitals of p-electron systems. In hydrocarbons the technique consists of forming linear combinations of unhybridized C2p orbitals. 14. In crystal-field theory, bonding in d-metal complexes arises from Coulomb interactions between electrons from the central metal ion and electrons from the

ligands. In an octahedral complex, the degenerate d atomic orbitals of the metal are split into two sets of orbitals separated by an energy DO: a triply degenerate set comprising the dxy, dyz, and dzx orbitals and labeled t2g and a doubly degenerate set comprising the dx −y and dz orbitals and labeled eg. 2

2

2

15. In a high-spin complex, the t2g and eg orbitals are filled in such a way as to maximize the number of unpaired d electrons. In a low-spin complex, the number of unpaired electrons is minimized. 16. Ligand-field theory is an adaptation of MO theory for complexes of the d metals. 17. In the self-consistent field procedure, an initial guess about the composition of the molecular orbitals is successively refined until the solution remains unchanged in a cycle of calculations. 18. In semi-empirical methods for the determination of electronic structure, the Schrödinger equation is written in terms of parameters chosen to agree with selected experimental quantities.

Checklist of key equations Property

Equation

Comment

A VB wavefunction

y(1,2) = yA(1)yB(2) + yA(2)yB(1)

yA and yB are atomic orbitals on different atoms

Resonance hybrid

y = ay1 + by2

y1 and y2 are wavefunctions for molecules with different electron distributions and the same nuclear locations

Molecular orbital

y = cAyA + cByB

An LCAO

Overlap integral

S=

Bond order

b = 12 (n − n*)

yAyB dt

Discussion questions 10.1 Compare the approximations built into valence bond theory and

10.5 Identify and justify the approximations used in the Hückel

molecular orbital theory.

theory of conjugated hydrocarbons.

10.2 Discuss the steps involved in the construction of sp3, sp2, and sp

10.6 Using information found in this and the previous chapter,

hybrid orbitals. 10.3 Distinguish between the Pauling and Mulliken electronegativity

discuss the unique role that carbon plays in biochemistry. 10.7 In the laboratory, the Fe2+ ion in the heme group of hemoglobin

scales.

can be removed and replaced by a Zn2+ ion. Discuss whether this modified protein is likely to bind O2 efficiently.

10.4 Use molecular orbital theory to discuss the biochemical

10.8 Distinguish between semi-empirical, ab initio, and density

reactivity of O2, N2, and NO.

functional theory methods of electronic structure determination.

404

10 THE CHEMICAL BOND

Exercises 10.9 Write down the valence bond wavefunction for a nitrogen

molecule. 10.10 Calculate the molar energy of repulsion between two hydrogen nuclei at the separation in H2 (74.1 pm). The result is the energy that must be overcome by the attraction from the electrons that form the bond. 10.11 Give the valence bond description of SO2 and SO3 molecules. 10.12 Write the Lewis structure for the peroxynitrite ion, ONOO−.

Label each atom with its state of hybridization and specify the composition of each of the different types of bond. 10.13 The structure of the visual pigment retinal is shown in (8). Label each atom with its state of hybridization and specify the composition of each of the different types of bond.

normalized to 1, then y is also normalized to 1. (c) To what values of q do the bonding and antibonding orbitals in a hom*onuclear diatomic molecule correspond? 10.21 Draw diagrams to show the various orientations in which a p orbital and a d orbital on adjacent atoms may form bonding and antibonding molecular orbitals. 10.22 How many molecular orbitals can be constructed from a diatomic molecule in which s, p, d, and f orbitals are all important in bonding? 10.23 Give the ground state electron configurations of (a) H2−, (b) N2,

and (c) O2. 10.24 Three biologically important diatomic species, either because they promote or inhibit life, are (a) CO, (b) NO, and (c) CN−. The first binds to hemoglobin, the second is a chemical messenger, and the third interrupts the respiratory electron transfer chain. Their biochemical action is a reflection of their orbital structure. Deduce their ground-state electron configurations. 10.25 Some chemical reactions proceed by the initial loss or transfer of an electron to a diatomic species. Which of the molecules N2, NO, O2, C2, F2, and CN would you expect to be stabilized by (a) the addition of an electron to form AB− and (b) the removal of an electron to form AB+? 10.26 Give the (g,u) parities of the wavefunctions for the first four

levels of a particle in a box. 10.14 Show that S = 2h1h2 dt = 0, where h1 = s + px + py + pz and

h2 = s − px − py + pz are hybrid orbitals. Hint: Each atomic orbital is individually normalized to 1. Also, note that S = 2spdt = 0, and that p orbitals with perpendicular orientations have zero overlap.

10.15 Show that the sp2 hybrid orbital (s + 21/2p)/31/2 is normalized to

1 if the s and p orbitals are each normalized to 1. 10.16 Find another sp2 hybrid orbital that has zero overlap with the hybrid orbital in the preceding problem. 10.17 Benzene is commonly regarded as a resonance hybrid of the two Kekulé structures shown in (4), but other structures can also contribute. Draw three other structures in which there are only covalent p bonds (allowing for bonding between some non-adjacent C atoms), and two structures in which there is one ionic bond. Why may these structures be ignored in simple descriptions of the molecule? 10.18 Before doing the calculation below, sketch how the overlap between a 1s orbital and a 2p orbital can be expected to depend on their separation. The overlap integral between a 1s orbital and a 2p orbital on nuclei separated by a distance R is S = (R/a0){1 + (R/a0) + 1 2 −R/a . Plot this function, and find the separation for which 3 (R/a0) }e the overlap is a maximum. 0

10.19 Suppose that a molecular orbital has the form N(0.145A + 0.844B). Find a linear combination of the orbitals A and B that has zero overlap with this combination. 10.20 Show, if overlap is ignored, (a) that any molecular orbital

expressed as a linear combination of two atomic orbitals may be written in the form y = yA cos q + yB sin q, where q is a parameter that varies between 0 and 12 p, and (b) that if yA and yB are orthogonal and

10.27 (a) Give the parities of the wavefunctions for the first four levels

of a harmonic oscillator. (b) How may the parity be expressed in terms of the quantum number v? 10.28 State the parities of the six p orbitals of benzene (see Fig. 10.39). 10.29 Two important diatomic molecules for the welfare of humanity are NO and N2: the former is both a pollutant and a chemical messenger, and the latter is the ultimate source of the nitrogen of proteins and other biomolecules. Use the electron configurations of NO and N2 to predict which is likely to have the greater bond dissociation energy and the shorter bond length. 10.30 Arrange the species O2+, O2, O2−, and O22− in order of increasing

bond length. 10.31 Construct the molecular orbital energy level diagrams of (a) ethene (ethylene) and (b) ethyne (acetylene) on the basis that the molecules are formed from the appropriately hybridized CH2 or CH fragments. 10.32 Many of the colors of vegetation are due to electronic transitions in conjugated p-electron systems. In the free-electron molecular orbital (FEMO) theory, the electrons in a conjugated molecule are treated as independent particles in a box of length L. Sketch the form of the two occupied orbitals in butadiene predicted by this model and predict the minimum excitation energy of the molecule. The tetraene CH2=CHCH=CHCH=CHCH=CH2 can be treated as a box of length 8R, where R = 140 pm (as in this case, an extra half bond length is often added at each end of the box). Calculate the minimum excitation energy of the molecule and sketch the hom*o and LUMO.

405

EXERCISES

10.33 Write down the Hückel secular determinant for cyclo-butadiene. 10.34 Solve the secular determinant for the allyl radical, CH2=CHCH2·. Hints: (a) regard the unpaired electron on the –CH2· fragment to be in a C2pz orbital, so that the electron can delocalize within the p system of the molecule. (b) See Mathematical toolkit 10.1. 10.35 It is important to understand the origins of stabilization of

linear conjugated molecules because they play important biological roles in plants and animals (see Case study 9.1). According to Hückel theory, the energies of the bonding p molecular orbitals of butadiene, CH2=CH2–CH2=CH2, are E = a + 1.62b and a + 0.62b. The energies of the antibonding p* molecular orbitals are E = a − 1.62b and a − 0.62b. The total p-electron binding energy, Ep, is the sum of the energies of each p electron. Recalling that there are four electrons to accommodate in the p molecular orbitals, calculate the p-electron binding energies of ethene (see Section 10.7) and butadiene. Is the energy of the butadiene molecule lower or higher than the sum of two individual p bonds? 10.36 Cyclic conjugated systems occur widely in biological macromolecules. Examples include the phenyl group of phenyalanine and a host of heterocyclic molecules, such as the purine and pyrimidine bases found in nucleic acids. In general, the delocalization energy of a conjugated system is

Edeloc = Ep − Ndb(2a + 2b) where Ndb is the number of double bonds, each contributing an energy 2a + 2b in the absence of conjugation. The most notable example of delocalization conferring extra stability is benzene and the aromatic molecules based on its structure. Predict the electronic configuration and delocalization energy of (a) the benzene anion and (b) the benzene cation. 10.37 The FEMO theory (Problem 10.32) of conjugated molecules is rather crude and better results are obtained with simple Hückel theory. (a) For a linear conjugated polyene with each of NC carbon atoms contributing an electron in a 2p orbital, the energies Ek of the resulting p molecular orbitals are given by

Ek = a + 2b cos

kp NC + 1

A 2 D C NC + 1 F

1/2

sin

jkp NC + 1

10.38 For monocyclic conjugated polyenes (such as cyclobutadiene and benzene) with each of NC carbon atoms contributing an electron in a 2p orbital, simple Hückel theory gives the following expression for the energies Ek of the resulting p molecular orbitals:

Ek = a + 2b cos

j = 1, 2, 3, . . . , NC

Determine the values of the coefficients of each of the six 2p orbitals in each of the six p molecular orbitals of hexatriene. Match each set of coefficients (that is, each molecular orbital) with a value of the energy

2kp NC

k = 0, ±1, ±2, . . . , ±NC/2

(even N)

k = 0, ±1, ±2, . . . , ±(NC − 1)/2

(odd N)

(a) Calculate the energies of the p molecular orbitals of benzene and cyclooctatetraene. Comment on the presence or absence of degenerate energy levels. (b) Calculate and compare the delocalization energies of benzene (using the expression above) and hexatriene (see Exercise 10.36). What do you conclude from your results? (c) Calculate and compare the delocalization energies of cyclooctaene and octatetraene. Are your conclusions for this pair of molecules the same as for the pair of molecules investigated in part (b)? 10.39 Experimentally, it is found that the value of DO varies with the

chemical nature of the ligand according to the spectrochemical series: S2− < Cl− < OH− ≈ RCO2− < H2O ≈ RS− < NH3 ≈ imidazole (the side chain of histidine) < CN− < CO. (a) Draw an energy level diagram like those in Fig. 10.41 showing the configuration of the d electrons on the metal ion in [Fe(OH2)6]3+ and [Fe(CN)6]3−. (b) Predict the number of unpaired electrons in each complex. 10.40 The terms low spin and high spin apply only to complexes of d-metal ions having certain numbers of d electrons. Put differently, certain d-metal ions can have only one electron configuration and a distinction between low- and high-spin complexes is not possible. For what number of d electrons are both high- and low-spin octahedral complexes possible? 10.41 Figures 10.41 and 10.42 show the result of an octahedral arrangement of ligands around a d-metal ion. In a tetrahedral complex, the dx −y and dz orbitals form a degenerate pair that is separated in energy from the degenerate dxy, dyz, and dzx orbitals by DT. In a square-planar complex with the ligand orbitals in the xy plane, the metal d orbitals increase in energy as follows: dxz = dyz < dz < dxy < dx −y . In a nickel-containing enzyme, the metal was shown to be in the +2 oxidation state and to have no unpaired electrons. What is the most probable geometry of the Ni2+ site? 2

k = 1, 2, 3, . . . , NC

Use this expression to determine a reasonable empirical estimate of the parameter b for the series consisting of ethene, butadiene, hexatriene, and octatetraene given that light-induced absorptions from the hom*o to the LUMO occur at 61 500, 46 080, 39 750, and 32 900 cm−1, respectively. (b) Calculate the delocalization energy of octatetraene (see Exercise 10.36). (c) In the context of this Hückel model, the p molecular orbitals are written as linear combinations of the carbon 2p orbitals. The coefficient of the jth atomic orbital in the kth molecular orbital is given by ckj =

calculated with the expression given in part (a) of the molecular orbital. Comment on trends that relate the energy of a molecular orbital with its ‘shape’, which can be inferred from the magnitudes and signs of the coefficients in the linear combination that describes the molecular orbital.

2

2

2

2

2

10.42 Ligands that interact with d metals as shown in Fig. 10.43 are called s-donor ligands. When p bonding is important, p-acceptor and p-donor ligands behave as shown in Fig. 10.45. If a ligand generates a weak ligand field around a d-metal ion, the result will be a small value of DO and a high-spin complex. Conversely, a strong ligand field leads to a large value of DO and a low-spin complex. (a) Justify the following statement: Cl− is a weak-field ligand because it is a p acceptor and CO is a strong-field ligand because it is a p donor. (b) Show that O2 is a p-acceptor ligand. (c) Using the information from parts (a) and (b) and from Case studies 4.1 and 10.4, propose a detailed mechanism for CO poisoning.

406

10 THE CHEMICAL BOND

Projects 10.43 In Section 10.2c we used VB theory to account for the planarity of the peptide link (1). Now we develop a molecular orbital theory treatment that provides a richer description of the factors that stabilize the planar conformation of the peptide link.

(a) Taking a hint from VB theory, we can suspect that delocalization of the p bond between the oxygen, carbon, and nitrogen atoms can be modeled by making LCAO-MOs from 2p orbitals perpendicular to the plane define by the atoms. The three combinations have the form y1 = acO + bcC + ccN

y2 = dcO − ecN

y3 = fcO − gcC + hcN

where the coefficients a to h are all positive. Sketch the orbitals y1, y2, and y3 and characterize them as bonding, non-bonding, or antibonding molecular orbitals. (b) Show that this treatment is consistent only with a planar conformation of the peptide link. (c) Draw a diagram showing the relative energies of these molecular orbitals and determine the occupancy of the orbitals. Hint: Convince yourself that there are four electrons to be distributed among the molecular orbitals. (d) Now consider a nonplanar conformation of the peptide link, in which the O2p and C2p orbitals are perpendicular to the plane defined by the O, C, and N atoms, but the N2p orbital lies on that plane. The LCAO-MOs are given by y4 = acO + bcC

y5 = ecN

y6 = fcO − gcC

Just as before, sketch these molecular orbitals and characterize them as bonding, nonbonding, or antibonding. Also, draw an energy level diagram and determine the occupancy of the orbitals.

between the hom*o and the LUMO. For example, a carbon–carbon bond in a linear polyene may have bonding character in the hom*o and antibonding character in the LUMO. Therefore, promotion of an electron from the hom*o to the LUMO weakens this carbon–carbon bond in the excited electronic state relative to the ground electronic state. (i) Use molecular modeling software to display the hom*o and LUMO of each molecule discussed in this project. (ii) Discuss in detail any changes in bond order that accompany the p-to-p* ultraviolet absorptions in these molecules. 10.45 Molecular orbital calculations may be used to predict trends in the standard potentials of conjugated molecules, such as the quinones and flavins, that are involved in biological electron transfer reactions (Chapter 5). It is commonly assumed that decreasing the energy of the LUMO enhances the ability of a molecule to accept an electron into the LUMO, with an attendant increase in the value of the molecule’s standard potential. Furthermore, a number of studies indicate that there is a linear correlation between the LUMO energy and the reduction potential of aromatic hydrocarbons.

(a) The biological standard potentials for the one-electron reduction of methyl-substituted p-benzoquinones (9) to their respective semiquinone radical anions are R2 H CH3 CH3 CH3 CH3

R3 H H H CH3 CH3

R5 H CH3 CH3 CH3 CH3

R6 H H H H CH3

E 9cell /V 0.078 0.023 −0.067 −0.065 −0.260

(e) Why is this arrangement of atomic orbitals consistent with a nonplanar conformation for the peptide link? (f) Does the bonding MO associated with the planar conformation have the same energy as the bonding MO associated with the nonplanar conformation? If not, which bonding MO is lower in energy? Repeat the analysis for the nonbonding and antibonding molecular orbitals. (g) Use your results from parts (a)–(f ) to construct arguments that support the planar model for the peptide link. The following projects require the use of molecular modeling software. 10.44 Here we explore further the application of molecular orbital

calculations to the prediction of spectroscopic properties of conjugated molecules. (a) Using data from Table 10.3, plot the hom*o-LUMO energy separations against the experimental frequencies for p-to-p* ultraviolet absorptions for ethene, butadiene, hexatriene, and octatetraene. Then use mathematical software to find the polynomial equation that best fits the data. (b) Using molecular modeling software and the computational method recommended by your instructor (extended Hückel, semiempirical, ab initio, or DFT methods), calculate the energy separation between the hom*o and LUMO of decapentaene. (c) Use your polynomial fit from part (a) to estimate the frequency of the p-to-p* absorption of decapentaene from the calculated hom*o-LUMO energy separation. (d) Discuss why the calibration procedure of part (a) is necessary. (e) Electronic excitation of a molecule may weaken or strengthen some bonds because bonding and antibonding characteristics differ

Using molecular modeling software and the computational method recommended by your instructor (extended Hückel, semi-empirical, ab initio, or DFT methods), calculate ELUMO, the energy of the LUMO 3 of each substituted p-benzoquinone, and plot ELUMO against E cell . Do 3 your calculations support a linear relation between ELUMO and E cell ? (b) The 1,4-benzoquinone for which R2 = R3 = CH3 and R5 = R6 = OCH3 is a suitable model of coenzyme Q, a component of the respiratory electron transport chain (Section 5.10). Determine ELUMO of this quinone and then use your results from part (a) to estimate its biological standard potential. (c) The p-benzoquinone for which R2 = R3 = R5 = CH3 and R6 = H is a suitable model of plastoquinone, a component of the photosynthetic electron transport chain (Section 5.11). Determine ELUMO of this quinone and then use your results from part (a) to estimate its standard potential. Is plastoquinone expected to be a better or worse oxidizing agent than coenzyme Q? (d) Based on your predictions and on basic concepts of biological electron transport (Sections 5.10 and 5.11), suggest a reason why coenzyme Q is used in respiration and plastoquinone is used in photosynthesis.

11

Macromolecules and self-assembly Biological cells are complex devices with outer shells built largely from lipids, sterols, and, in some organisms, complex carbohydrates. Inside the cells are information storage and retrieval systems—the chromosomes—and molecular machines—enzymes, ion channels and pumps, and so on—made from small molecules and macromolecules, such as proteins, nucleic acids, and polysaccharides. The construction of functional structures in the cell proceeds largely through self-assembly, the spontaneous formation of complex aggregates of molecules or macromolecules held together by a variety of molecular interactions of the kind described later in the chapter. We have already encountered a few examples of self-assembly, such as the formation of biological membranes from lipids and of a DNA double helix from two polynucleotide chains (Fundamentals F.1). In this chapter, we describe several techniques for the determination of the size and shape of biological macromolecules and aggregates, and then explore the interactions responsible for the shapes so found. These interactions contribute to a whole hierarchy of structure, from ‘no structure’ in fluids all the way up to the elaborate and functionally important structures of proteins and nucleic acids. We also describe computer-aided methods for building three-dimensional models of macromolecules in which the molecular interactions that promote self-assembly are optimized.

Determination of size and shape In this section we explore important methods used in modern biochemical research to determine the molar mass and structure of very large molecules. The most powerful of these techniques are based on the diffraction of X-rays from crystalline samples and reveal the position of almost every heavy atom (that is, every atom other than hydrogen) even in very large molecules.

Determination of size and shape 407 11.1 11.2 11.3 11.4

Ultracentrifugation Mass spectrometry Laser light scattering X-ray crystallography

In the laboratory 11.1

The crystallization of biopolymers

Because molar mass is so important for the identification of a molecule and the determination of its structure, we need to discuss sophisticated and accurate methods for its determination.

In a gravitational field, heavy particles settle toward the foot of a column of solution by the process called sedimentation. The rate of sedimentation depends on the strength of the field and on the masses and shapes of the particles. Spherical molecules (and compact molecules in general) sediment faster than rodlike or extended molecules. For example, DNA helices sediment much faster when they are denatured to a random coil, so sedimentation rates can be used to study

421

In the laboratory 11.2

Data acquisition in X-ray crystallography 422 Case Study 11.1 The structure of DNA from X-ray diffraction studies 423 The control of shape

Interactions between partial charges 11.6 Electric dipole moments 11.7 Interactions between dipoles 11.8 Induced dipole moments 11.9 Hydrogen bonding 11.10 The total interaction Case Study 11.2 Molecular recognition in biology and pharmacology

424

11.5

Levels of structure

11.1 Ultracentrifugation

407 410 412 414

425 426 429 431 433 435

437 438

11.11 Minimal order:

gases and liquids 11.12 Random coils 11.13 Proteins 11.14 Nucleic acids 11.15 Polysaccharides

438 440 442 446 448

11.16 Micelles and biological

membranes

449

11.17 Computer-aided

simulation Checklist of key concepts Checklist of key equations Discussion questions Exercises Projects

451 455 456 457 457 460

408

11 MACROMOLECULES AND SELF-ASSEMBLY

denaturation. When the sample is at equilibrium, the particles are dispersed over a range of heights because the gravitational field competes with the stirring effect of thermal motion. The spread of heights depends on the masses of the molecules, so the equilibrium distribution is another way to determine molar mass. Sedimentation is normally very slow, but it can be accelerated by ultracentrifugation, a technique that replaces the gravitational field with a centrifugal field. The effect can be achieved in an ultracentrifuge, which is essentially a cylinder that can be rotated at high speed about its axis with a sample in a cell near its periphery (Fig. 11.1). Modern ultracentrifuges can produce accelerations equivalent to about 105 that of gravity (‘105 g’). Initially the sample is uniform, but the ‘top’ (innermost) boundary of the solute moves outward as sedimentation proceeds. (a) The sedimentation rate

(a) An ultracentrifuge head. The sample on one side is balanced by a blank diametrically opposite. (b) Detail of the sample cavity: the ‘top’ surface is the inner surface, and the centrifugal force causes sedimentation toward the outer surface; a particle at a radius r experiences a force of magnitude meffrw 2.

Fig. 11.1

Solute particles in a spinning rotor adopt a constant speed away from the rotational axis because the outward, centrifugal force is balanced by a retarding, frictional force. The sedimentation constant, S, a measure of the rate at which a particle migrates in the centrifugal field, is defined as S=

s rw 2

Sedimentation constant

(11.1)

where s is the speed of sedimentation, r is the distance of the sample from the rotational axis, and w is the angular velocity of the rotor (in radians per second). For biological macromolecules, typical values of S are of the order of 10−13 s and depend on the shape and size of the particle, the temperature, and the viscosity of the solution. A common unit for S is the ‘svedberg’, denoted Sv and defined as 1 Sv = 10−13 s. For example, the sedimentation constant of the protein bovine serum albumin is 5.02 Sv in water at 25°C. We show in the following Justification that the molar mass of a macromolecule is related to its sedimentation constant, S, and diffusion constant, D, by the relation M=

SRT bD

Relation between the molar mass and the sedimentation constant

(11.2)

where b = 1 − rvs is a correction factor that takes into account the buoyancy of the solution, with r the mass density of the solvent (typically in grams per cubic centimeter) and vs the specific volume of the solute (typically in cubic centimeters per gram). Justification 11.1 The sedimentation constant

A solute particle of mass m has an effective mass meff = bm in the solution. The solute particles at a distance r from the axis of a rotor spinning at an angular velocity w experience a centrifugal force of magnitude meff rw 2. The acceleration outward is countered by a frictional force proportional to the speed, s, of the particles through the medium. This force is written fs, where f is the frictional coefficient. The particles therefore adopt a drift speed, a constant speed through the medium, which is found by equating the two forces meff rw 2 and fs. The forces are equal when s=

meff rw 2 bmrw 2 = f f

11.1 ULTRACENTRIFUGATION Next, we draw on the Stokes–Einstein relation, f = kT/D, between the frictional coefficient, f, and the diffusion coefficient, D, to write s=

bmrw 2D bMrw 2D = kT RT

where we have used the relation m = M/NA between the molecular mass and the molar mass M and the relation R = kNA between Boltzmann’s constant and the gas constant. Use of eqn 11.1 and rearrangement of this expression gives eqn 11.2.

The diffusion coefficient is related to the rate at which molecules migrate down a concentration gradient (it is treated in detail in Section 8.5) and can be measured by observing the rate at which a concentration boundary moves or the rate at which a more concentrated solution diffuses into a less concentrated one. The diffusion coefficient can also be measured by using laser light-scattering methods (Section 11.3). It follows that we can find the molar mass by combining measurements of sedimentation and diffusion rates (to obtain S and D, respectively). Self-test 11.1 Determine the molar mass of human hemoglobin, given that it has a sedimentation constant of 4.48 Sv and a diffusion coefficient of 6.9 × 10−11 m2 s−1 in a solution with rvs = 0.748 at 293 K.

Answer: 63 kg mol−1

(b) Sedimentation equilibrium

It is sometimes more convenient to measure the equilibrium distribution of molecules than the rate at which they sediment. At equilibrium, when the tendency of the solute to settle is balanced by the spreading effect of thermal motion, the molar mass can be obtained from the ratio of concentrations c2 /c1 of the macromolecules at two different radii r2 and r1, respectively, in a centrifuge operating at angular frequency w: M=

2RT c ln 2 2 2 共r − r 1)bw c1 2 2

Molar mass from sedimentation data

(11.3a)

where R is the gas constant. The centrifuge is run more slowly in this technique than in the sedimentation rate method to avoid having all the solute pressed in a thin film against the bottom of the cell. At these slower speeds, several days may be needed for equilibrium to be reached. Example 11.1

The molar mass of a protein from ultracentrifugation experiments

The data from an equilibrium ultracentrifugation experiment performed at 300 K on an aqueous solution of a protein show that a graph of ln c against (r/cm)2 is a straight line with a slope of 0.729. The rotational rate of the centrifuge was 50 000 rotations per minute and b = 0.70. Calculate the molar mass of the protein. Strategy We need to reinterpret eqn 11.3 in terms of the slope of a plot of ln c against r 2. To do so, we apply the relation ln(x/y) = ln x − ln y to eqn 11.3 and obtain, after minor rearrangement,

409

410

11 MACROMOLECULES AND SELF-ASSEMBLY

M=

2RT ln c2 − ln c1 × r 22 − r 12 bw 2

If a plot of ln c against r 2 is linear, then the ratio (ln c2 − ln c1)/(r 22 − r 12) has the form of the slope of the line. In practice, ln(c/g cm−3) is plotted against (r/cm)2 to give a dimensionless slope. It follows that M=

2RT × (slope × cm−2) bw 2

(11.3b)

and we can use the data provided to calculate the molar mass M. Each full revolution of the rotor corresponds to an angular change of 2p radians, so to obtain the angular frequency w, we multiply the rotation rate in cycles per second by 2p. Solution The angular frequency is A note on good practice

A molar mass (the mass per mole of molecules) of 4.4 kg mol−1 corresponds to a molecular mass (the mass of one molecule) of 4.4 kDa. Be careful to use the unit dalton (Da) to denote molecular, not molar, mass.

w = 2p × (50 000 min−1) ×

1 min 2p × 50 000 −1 = s 60 s 60

It follows from eqn 11.3 with 1 cm−2 = 104 m−2 and the slope 0.729 that the molar mass is M=

2 × (8.3145 J K−1 mol−1) × (300 K) × (0.729 × 104 m−2) 2 A 2p × 50 000 −1 D (1 − 0.70) × s C F 60

where we have used 1 J = 1 kg m2 s−2. The molar mass is therefore 4.4 kg mol−1. Self-test 11.2 The data from a sedimentation equilibrium experiment per-

formed at 293 K on a macromolecular solute in aqueous solution show that a graph of ln(c/g cm−3) against (r/cm)2 is a straight line with a slope of 0.821. The rotation rate of the centrifuge was 4500 Hz (1 Hz = 1 s−1) and rvs = 0.40. Calculate the molar mass of the solute. Answer: 8.3 kg mol−1, corresponding to a molecular mass of 8.3 kDa

11.2 Mass spectrometry

Diagram of a matrix-assisted laser desorption/ ionization time-of-flight (MALDI-TOF) mass spectrometer. A laser beam ejects macromolecules and ions from the solid matrix. The ionized macromolecules are accelerated by an electrical potential difference over a distance d and then travel through a drift region of length l. Ions with the smallest mass-to-charge ratio (m/z) reach the detector first. Fig. 11.2

The most precise technique for the determination of molar mass is mass spectrometry, and we need to know how to adapt traditional techniques developed for small molecules to the study of biological macromolecules.

In mass spectrometry, the sample is first ionized in the gas phase and then the mass-to-charge number ratios, m/z, of all the resulting ions are measured. Macromolecules present a challenge because it is difficult to produce gaseous ions of large species without fragmentation. However, two new techniques have emerged that circumvent this problem: matrix-assisted laser desorption/ionization (MALDI) and electrospray ionization. We shall discuss MALDI–TOF mass spectrometry, so-called because the MALDI technique is coupled to a time-offlight (TOF) ion detector. Figure 11.2 shows a schematic view of a MALDI–TOF mass spectrometer. The macromolecule is first embedded in a solid matrix that often consists of an organic acid such as 2,5-dihydroxybenzoic acid, nicotinic acid, or an a-cyanocarboxylic acid.

11.2 MASS SPECTROMETRY

This sample is then irradiated with a laser pulse. The pulse of electromagnetic energy ejects matrix ions, cations, and neutral macromolecules, thus creating a dense gas plume above the sample surface. The macromolecule is ionized by collisions and complexation with H+ cations. In the TOF spectrometer, the ions are accelerated over a short distance d by an electrical field of strength E and then travel through a drift region of length l. We show in the following Justification that the time of flight, t, required for an ion of mass m and charge number z to reach the detector at the end of the drift region is t=l

A m D2 C 2zeEd F

Time of flight in a TOF spectrometer

(11.4)

where e is the fundamental charge. Because d, l, and E are fixed for a given experiment, the time of flight of the ion is a direct measure of its m/z ratio, which is given by m/z = 2zeEd

A tD 2 C lF

The mass-to-charge number ratio in a TOF spectrometer

(11.5)

411

A note on good (in this case, common) practice

Strictly, the units of m/z are kilograms; however, it is conventional to interpret m as the ratio of the molecular mass to the atomic mass constant mu, in which case ‘m/z’ (strictly m/zmu) is dimensionless.

Justification 11.2 The time of flight of an ion in a mass spectrometer

Consider an ion of charge ze and mass m that is accelerated from rest by an electric field of strength E applied over a distance d. The kinetic energy, Ek, of the ion is Ek = 12 mv 2 = zeEd where v is the speed of the ion. The drift region, l, and the time of flight, t, in the mass spectrometer are both sufficiently short that we can ignore acceleration and write v = l/t. Then substitution into the expression for Ek gives 1 2

m

A lD2 = zeEd C tF

Rearrangement of this equation gives eqn 11.5.

Figure 11.3 shows the MALDI-TOF mass spectrum of bovine albumin. The MALDI technique produces unfragmented molecular ions of varying charges, with the singly charged ion often giving rise to the most prominent feature in the spectrum. The spectrum of a mixture of biopolymers consists of multiple peaks arising from molecules with different molar masses. The intensity of each peak is proportional to the abundance of each biopolymer in the sample.

Self-test 11.3 A MALDI-TOF mass spectrum consists of two intense features at m/z = 9912 and 4554. Does the sample contain one or two distinct biopolymers? Explain your answer.

Answer: Two distinct biopolymers because the feature at lower m/z probably does not arise from the unfragmented +2 cation of the species that gives rise to the feature at higher m/z.

Fig. 11.3 The MALDI-TOF mass spectrum of bovine albumin, a protein with molar mass 66.43 kg mol−1. During the MALDI process, the protein takes up one or two H+ ions, making molecular ions of charge +1 and +2, respectively. Because the protein does not fragment, the +2 ion gives rise to a peak in the spectrum at a m/z value that is one-half the value for the peak associated with the +1 ion. (Adapted from B.S. Larsen and C.N. McEwen in Mass spectrometry of biological materials, Marcel Dekker, New York (1998).)

412

11 MACROMOLECULES AND SELF-ASSEMBLY

Rayleigh scattering from a sample of point-like particles. The intensity of scattered light depends on the angle q between the incident and scattered beams. The treatment developed in the text corresponds to an experimental arrangement in which the plane of polarization of the laser beam (the dark blue plane in the inset) is perpendicular to the plane defined by the incident ray and the line from the sample to the detector (the light blue plane in the inset).

Fig. 11.4

11.3 Laser light scattering The analysis of the intensity of laser light scattered by a solution of a biological macromolecule yields information about its size and shape.

Scattering of light by particles with diameters much smaller than the wavelength of the incident radiation is called Rayleigh scattering. In the Rayleigh regime, the intensity of scattered light is proportional to the molar mass of the particle and to l−4, so shorter-wavelength radiation is scattered more intensely than longer wavelengths. For example, the blue of the sky arises from the more intense scattering of the blue component of white sunlight by the molecules of the atmosphere. (a) Rayleigh scattering A brief comment

The factor r 2 occurs in the definition of the Rayleigh ratio because the light wave spreads out over a sphere of radius r and surface area 4pr 2, so any sample of the radiation has its intensity I(q) decreased by a factor proportional to r 2. Therefore, the quantity I(q) × r 2, and not simply I(q), should be compared to I0 in forming the Rayleigh ratio. We also note that the definition of the Rayleigh ratio given here applies only to the experimental conditions in Fig. 11.4.

Consider the experimental arrangement shown in Fig. 11.4 for the measurement of light scattering from solutions of macromolecules. Typically, the sample is irradiated with monochromatic light from a laser. The intensity of scattered light is then measured as a function of the angle q that the direction of the laser beam makes with the direction of the detector from the sample at a distance r. Under these conditions, the intensity, I(q), of light scattered is written as the Rayleigh ratio: R(q) =

I(q) 2 ×r I0

Definition of the Rayleigh ratio

(11.6)

where I0 is the intensity of the incident laser radiation. A detailed examination of the scattering shows that the Rayleigh ratio depends on the mass concentration, cM (units: kg m−3), of the macromolecule and its molar mass M as: R(q) = KP(q)cMM

The relation of Rayleigh ratio to molar mass

(11.7)

where the constant K depends on the refractive index of the solution, the incident wavelength, and the distance between the detector and the sample, which is held constant during the experiment. The quantity P(q) is the structure factor, which is related to the size of the molecule. When the molecule is much smaller than the wavelength of the light, P(q) ≈ 1. However, when the size of the molecule is about one-tenth the wavelength of the incident radiation, it is possible to show that P(q) ≈ 1 −

16p 2R g2 sin2 12 q 3l2

Structure factor for small molecules

(11.8)

11.3 LASER LIGHT SCATTERING

where Rg is the radius of gyration of the macromolecule, a measure of its size (Section 11.12). Equation 11.7 applies only to ideal solutions. In practice, even relatively dilute solutions of macromolecules can deviate considerably from ideality, as we saw in In the laboratory 3.1. Being so large, macromolecules displace a large quantity of solvent instead of replacing individual solvent molecules with negligible disturbance. To take deviations from ideality into account, it is common to rewrite eqn 11.7 as KcM/R(q) = 1/P(q)M and to extend it to KcM 1 + BcM = R(q) P(q)M

(11.9)

where B is an empirical constant analogous to the osmotic virial coefficient (In the laboratory 3.1) and indicative of the effect of excluded volume. The preceding discussion shows that structural properties, such as the size and molar mass of a macromolecule, can be obtained from measurements of light scattering by a sample at several angles q relative to the direction of propagation on an incident beam. In modern instruments, lasers are used as the radiation sources.

Example 11.2

Determining the molar mass and size of a protein by laser light scattering

The following data for an aqueous solution of a protein with cM = 2.00 kg m−3 were obtained at 20°C with laser light at l = 532 nm. q/° R(q)/m2

15.0 23.8

45.0 22.9

70.0 21.6

85.0 20.7

90.0 20.4

In a separate experiment, it was determined that K = 2.40 × 10−2 mol m5 kg −2. From this information, calculate R g and M for the protein. Assume that B is negligibly small and that the protein is small enough that eqn 11.7 holds. Strategy Substituting the result of eqn 11.8 into eqn 11.7 we obtain, after some rearrangement:

D 1 1 A 16p 2Rg2D A 1 = + sin2 12 q 2 F C F C R(q) KcMM 3l R(q) Hence, a plot of 1/R(q) against {1/R(q)}sin2 12 q should be a straight line with slope 16p 2Rg2 /3l2 and y-intercept 1/KcMM. As usual, the plot should be of dimensionless quantities, so we actually plot 1/(R(q)/m2) against {1/(R(q)/m2)} × sin2 12 q, in which case the dimensionless slope is equal to the dimensionless quantity 16p 2Rg2 /3l2 and the dimensionless intercept is equal to 1/KcMM. Solution We construct a table of values of 1/R(q) and {1/R(q)}sin2 12 q and plot

the data (Fig. 11.5). 102/(R(q)/m2) 103 × (sin2 12 q)/(R(q)/m2)

4.20 0.716

4.37 6.40

4.63 15.2

4.83 22.0

4.90 24.5

The best straight line through the data has a slope of 0.295 and a y-intercept of 1/(R(q)/m2) = 4.18 × 10−2. From these values, we calculate Rg =

A 3l2 × slope D 1/2 A 3 × (532 nm)2 × 0.295 D 1/2 = = 39.8 nm C 16p 2 F C F 16p 2

Plot of the data for Example 11.2.

Fig. 11.5

413

414

11 MACROMOLECULES AND SELF-ASSEMBLY

M=

1 m2 (2.40 × 10−2 mol m5 kg−2) × (2.00 kg m−3) × (4.18 × 10−2)

= 4.98 × 102 kg mol−1 We conclude that the radius of gyration is 39.8 nm and the molar mass is 498 kg mol−1. Self-test 11.4 The following data for an aqueous solution of a macromolecule

were obtained at 20°C with plane-polarized light at l = 546 nm. q/º R(q)/m2

26.0 19.7

36.9 18.8

66.4 17.1

90.0 16.0

113.6 14.4

In separate experiments, it was determined that K = 6.42 × 10−5 mol m5 kg −2. From this information, and using cM = 311 kg m−3, calculate the R g and M of the macromolecule. State any assumptions you must make to solve this problem. Answer: R g = 46.9 nm and M = 987 kg mol−1

(b) Dynamic light scattering

A special laser scattering technique, dynamic light scattering, can be used to investigate the diffusion of macromolecules in solution. Consider two molecules being irradiated by a laser beam. Suppose that at a time t the scattered waves from these particles interfere constructively at the detector, leading to a large signal. However, as the molecules move through the solution, the scattered waves may interfere destructively at another time t′ and result in no signal. When this behavior is extended to a very large number of molecules in solution, it results in fluctuations in light intensity that depend on the diffusion coefficient, D. Hence, analysis of the fluctuations gives the diffusion coefficient and molecular size in cases where the molecular shape is known. Light scattering is a convenient method for the characterization of biological systems from proteins to viruses. Unlike mass spectrometry, laser light-scattering measurements may be performed in nearly intact samples; often the only preparation required is filtration of the sample.

11.4 X-ray crystallography The success of modern biochemistry in explaining such processes as DNA replication, protein biosynthesis, and enzyme catalysis is a direct result of developments in preparatory, instrumental, and computational procedures that have led to the determination of large numbers of structures of biological macromolecules by techniques based on X-ray diffraction.

Because much of our knowledge of the three-dimensional structures of biological macromolecules comes from studies of crystals of proteins and nucleic acids, we need to study the arrangements adopted by molecules when they stack together to form a crystalline solid. One of the most important techniques for the determination of the structures of crystals is X-ray diffraction. In its most sophisticated version, known as X-ray crystallography, X-ray diffraction provides detailed information about the location of all the atoms in molecules as complicated as biological macromolecules.

11.4 X-RAY CRYSTALLOGRAPHY

415

(a) Diffraction

A characteristic property of waves is that they interfere with one another, which means that they give a greater amplitude where their displacements add and a smaller amplitude where their displacements subtract (Section 9.1). Because the intensity of electromagnetic radiation is proportional to the square of the amplitude of the waves, the regions of constructive and destructive interference show up as regions of enhanced and diminished intensities. The phenomenon of diffraction is the interference caused by an object in the path of waves, and the pattern of varying intensity that results is called the diffraction pattern (Fig. 11.6). Diffraction occurs when the dimensions of the diffracting object are comparable to the wavelength of the radiation. Sound waves, with wavelengths of the order of 1 m, are diffracted by macroscopic objects. Light waves, with wavelengths of the order of 500 nm, are diffracted by narrow slits. X-rays have wavelengths comparable to bond lengths in molecules and the spacing of atoms in crystals (about 100 pm), so they are diffracted by them. By analyzing the diffraction pattern, it is possible to draw up a detailed picture of the location of atoms. The short-wavelength electromagnetic radiation we call X-rays is produced by bombarding a metal with high-energy electrons. The electrons decelerate as they plunge into the metal and generate radiation with a continuous range of wavelengths. This radiation is called bremsstrahlung.1 Superimposed on the continuum are a few high-intensity, sharp peaks. These peaks arise from the interaction of the incoming electrons with the electrons in the inner shells of the atoms. A collision expels an electron (Fig. 11.7), and an electron of higher energy drops into the vacancy, emitting the excess energy as an X-ray photon. An example of the process is the expulsion of an electron from the K shell (the shell with n = 1) of a copper atom, followed by the transition of an outer electron into the vacancy. If an electron from the L shell undergoes the transition, then the energy so released gives rise to copper’s Ka radiation of wavelength 154 pm. In 1912, the German physicist Max von Laue suggested that X-rays might be diffracted when passed through a crystal, for the wavelengths of X-rays are comparable to the separation of atoms. Laue’s suggestion was confirmed almost immediately by Walter Friedrich and Paul Knipping, and then developed by William and Laurence Bragg (father and son), who later jointly received the Nobel Prize. It has grown since then into a technique of extraordinary power.

The X-ray diffraction pattern obtained from a fiber of B-DNA. The black dots are the reflections, the points of maximum constructive interference, that are used to determine the structure of the molecule (see Case study 11.1). (Adapted from an illustration that appears in J.P. Glusker and K.N. Trueblood, Crystal structure analysis: A primer. Oxford University Press (1972).)

Fig. 11.6

(b) Crystal systems

X-ray diffraction is applied to crystalline arrays of molecules, so we need to know how to describe the arrangement of molecules in a crystal. The pattern that atoms, ions, or molecules adopt in a crystal is expressed in terms of an array of points making up the lattice that identify the locations of the individual species (Fig. 11.8). A unit cell of a crystal is the small three-dimensional figure obtained by joining typically eight of these points, which may be used to construct the entire crystal lattice by purely translational displacements, much as a wall may be constructed from bricks (Fig. 11.9). An infinite number of different unit cells can describe the same structure, but it is conventional to choose the cell with sides that have the shortest lengths and are most nearly perpendicular to one another. Unit cells are classified into one of seven crystal systems according to the symmetry they possess under rotations about different axes. The cubic system, for example, has four threefold axes (Fig. 11.10). A threefold axis is an axis of 1

Bremse is German for ‘brake’, Strahlung for ‘radiation’.

The formation of X-rays. When a metal is subjected to a high-energy electron beam, an electron in an inner shell of an atom is ejected. When an electron falls into the vacated orbital from an orbital of much higher energy, the excess energy is released as an X-ray photon.

Fig. 11.7

416

11 MACROMOLECULES AND SELF-ASSEMBLY

A unit cell, here shown in three dimensions, is like a brick used to construct a wall. Once again, only pure translations are allowed in the construction of the crystal. (Some bonding patterns for actual walls use rotations of bricks, so for these patterns a single brick is not a unit cell.) Fig. 11.9

(a) A crystal consists of a uniform array of atoms, molecules, or ions, as represented by these spheres. In many cases, the components of the crystal are far from spherical, but this diagram illustrates the general idea. (b) The location of each atom, molecule, or ion can be represented by a single point; here (for convenience only), the locations are denoted by a point at the center of the sphere. The unit cell, which is shown boxed, is the smallest block from which the entire array of points can be constructed without rotating or otherwise modifying the block. Fig. 11.8

Fig. 11.10 A unit cell belonging to the cubic system has four threefold axes (denoted C3) arranged tetrahedrally.

a rotation that restores the unit cell to the same appearance three times during a complete revolution, after rotations through 120°, 240°, and 360°. The four axes make the tetrahedral angle to each other. The monoclinic system has one twofold axis (Fig. 11.11). A twofold axis is an axis of a rotation that leaves the cell apparently unchanged twice during a complete revolution, after rotations through 180° and 360°. The essential symmetries, the properties that must be present for the unit cell to belong to a particular system, are listed in Table 11.1. A unit cell may have lattice points other than at its corners, so each crystal system can occur in a number of different varieties. For example, in some cases points may occur on the faces and in the body of the cell without destroying the cell’s essential symmetry. These various possibilities give rise to 14 distinct types of unit cell, called Bravais lattices. Three examples are shown in Fig. 11.12. (c) Crystal planes

The essential symmetries of the seven crystal systems

Table 11.1

System

Essential symmetries

Triclinic

None

Monoclinic

One twofold axis

Orthorhombic

Three perpendicular twofold axes

Rhombohedral

One threefold axis

Tetragonal

One fourfold axis

Hexagonal

One sixfold axis

Cubic

Four threefold axes in a tetrahedral arrangement

To specify a unit cell fully, we need to know not only its symmetry but its size, such as the lengths of its sides. There is a useful relation between the spacing of the planes passing through the lattice points, which (as we shall see) we can measure, and the lengths we need to know. Because two-dimensional arrays of points are easier to visualize than three-dimensional arrays, we shall introduce the concepts we need by referring to two dimensions initially and then extend the conclusions to three dimensions. Consider the two-dimensional rectangular lattice formed from a rectangular unit cell of sides a and b (Fig. 11.13). We can distinguish the four sets of planes shown in the illustration by the distances at which they intersect the axes. One way of labeling the planes would therefore be to denote each set by the smallest intersection distances. For example, we could denote the four sets in the illustration as (1a,1b), (3a,2b), (−1a,1b), and (∞a,1b). If, however, we agreed always to quote distances along the axes as multiples of the lengths of the unit cell, then we could omit the a and b and label the planes more simply as (1,1), (3,2), (−1,1), and (∞,1). Now suppose that the array in Fig. 11.13 is the top view of a three-dimensional rectangular lattice in which the unit cell has a length c in the z direction. All four

11.4 X-RAY CRYSTALLOGRAPHY

Fig. 11.11 A unit cell belonging to the monoclinic system has one twofold (denoted C2) axis (along b).

Fig. 11.12 The cubic unit cells. The letter P denotes a primitive unit cell, I a body-centered unit cell, and F a face-centered unit cell.

sets of planes intersect the z-axis at infinity, so the full labels of the sets of planes of lattice points are (1,1,∞), (3,2,∞), (−1,1,∞), and (∞,1,∞). The presence of infinity in the labels is inconvenient. We can eliminate it by taking the reciprocals of the numbers in the labels; this step also turns out to have further advantages, as we shall see. The resulting Miller indices, (hkl), are the reciprocals of the numbers in the parentheses with fractions cleared. For example, the (1,1,∞) planes in Fig. 11.13 are the (110) planes in the Miller notation. Similarly, the (3,2,∞) planes become first (13, 12,0) when reciprocals are formed and then (2,3,0) when fractions are cleared by multiplication through by 6, so they are referred to as the (230) planes. We write negative indices with a bar over the number: Fig. 11.13c shows the (1¯10) planes. Figure 11.14 shows some planes in three dimensions, including an example of a lattice with axes that are not mutually perpendicular. Self-test 11.5 A representative member of a set of planes in a crystal inter-

sects the axes at 3a, 2b, and 2c. What are the Miller indices of the planes? Answer: (233)

It is helpful to keep in mind the fact, as illustrated in Fig. 11.13, that the smaller the value of h in the Miller index (hkl), the more nearly parallel the plane is to the a axis. The same is true of k and the b axis and l and the c axis. When h = 0, the planes intersect the a axis at infinity, so the (0kl) planes are parallel to the a axis. Similarly, the (h0l) planes are parallel to b and the (hk0) planes are parallel to c.

417

Fig. 11.13 Some of the planes that can be drawn through the points of the space lattice and their corresponding Miller indices (hkl).

418

11 MACROMOLECULES AND SELF-ASSEMBLY

The Miller indices are very useful for calculating the separation of planes. For instance, they can be used to derive the following very simple expression for the separation, d, of the (hkl) planes in a rectangular lattice: 1 h2 k 2 l 2 = + + d 2 a2 b 2 c 2

Separation of planes

(11.10)

Justification 11.3 The separation of lattice planes

Consider the (hk0) planes of a rectangular lattice with sides of lengths a and b (Fig. 11.15). We can write the following trigonometric expressions for the angle f shown in the illustration: sin f =

d hd = (a/h) a

cos f =

d kd = (b/k) b

Then, because sin2 f + cos2 f = 1, we obtain h2d 2 k 2d 2 + 2 =1 a2 b which we can rearrange into 1 h2 k 2 = + d 2 a2 b 2 Fig. 11.14 Some representative planes in three dimensions and their Miller indices. Note that a 0 indicates that a plane is parallel to the corresponding axis. The indexing may also be used for unit cells with nonorthogonal axes.

Now consider an orthorhombic unit cell, a unit cell with perpendicular faces but different lengths of their edges (Fig. 11.16). In three dimensions, the expression above generalizes to eqn 11.10.

Example 11.3

Using the Miller indices

Calculate the separation of (a) the (123) planes and (b) the (246) planes of an orthorhombic cell with a = 0.84 nm, b = 0.96 nm, and c = 0.77 nm. Strategy For the first part, we simply substitute the information into eqn 11.10.

For the second part, instead of repeating the calculation, we should examine how d in eqn 11.10 changes when all three Miller indices are multiplied by 2 (or by a more general factor, n). Solution Substituting the data into eqn 11.10 gives

1 12 22 32 21 = + + = 2 2 2 2 d (0.84 nm) (0.96 nm) (0.77 nm) nm2 Fig. 11.15 The geometrical construction used to relate the separation of planes to the dimensions of a rectangular unit cell.

It follows that d = 0.22 nm. When the indices are all increased by a factor of 2, the separation becomes 1 (2 × 1)2 (2 × 2)2 (2 × 3)2 21 = + + =4× d 2 (0.84 nm)2 (0.96 nm)2 (0.77 nm)2 nm2 So, for these planes d = 0.11 nm. In general, increasing the indices uniformly by a factor n decreases the separation of the planes by n.

11.4 X-RAY CRYSTALLOGRAPHY

419

Self-test 11.6 Calculate the separation of the (133) and (399) planes in the

same lattice. Answer: 0.19 nm, 0.065 nm

(d) Bragg’s law

The earliest approach to the analysis of X-ray diffraction patterns treated a plane of atoms as a semitransparent mirror and modeled the crystal as stacks of reflecting planes of separation d (Fig. 11.17). The model makes it easy to calculate the angle the crystal must make to the incoming beam of X-rays for constructive interference to occur. It has also given rise to the name reflection to denote an intense spot arising from constructive interference. The path-length difference of the two rays shown in the illustration is

Fig. 11.16 An orthorhombic unit cell with sides of lengths a, b, and c.

AB + BC = 2d sin q where the angle q is often expressed as the glancing angle 2q. When the pathlength difference is equal to one wavelength (AB + BC = l), the reflected waves interfere constructively. It follows that a reflection should be observed when the glancing angle satisfies Bragg’s law: l = 2d sin q

Bragg’s law

(11.11a)

The primary use of Bragg’s law is to determine the spacing between the layers of atoms, for once the angle q corresponding to a reflection has been determined, d may readily be calculated. Equation 11.11a is sometimes written nl = 2d sin q

Alternative version of Bragg’s law

(11.11b)

with n = 1, 2, . . . denoting the order of the reflection, but the modern tendency is to incorporate n into the definition of d, as illustrated in Example 11.4. Example 11.4

Using Bragg’s law

A reflection from the (111) planes of a cubic crystal was observed at a glancing angle of 11.2° when Cu K a X-rays of wavelength 154 pm were used. What is the length of the side of the unit cell? Strategy We can find the separation, d, of the lattice planes from eqn 11.11

and the data. Then we find the length of the side of the unit cell by using eqn 11.10. Because the unit cell is cubic, a = b = c, so eqn 11.10 simplifies to 1 h2 + k 2 + l 2 = d2 a2 which rearranges to a = d × (h2 + k 2 + l 2)1/2 Solution According to Bragg’s law, the separation of the (111) planes respon-

sible for the diffraction is d=

l 154 pm = 2 sin q 2 sin 11.2°

It then follows that with h = k = l = 1, a=

154 pm × 31/2 = 687 pm 2 sin 11.2°

Fig. 11.17 The derivation of Bragg’s law treats each lattice plane as reflecting the incident radiation. The path lengths differ by AB + BC, which depends on the angle q. Constructive interference (a ‘reflection’) occurs when AB + BC is equal to an integral number of wavelengths.

420

11 MACROMOLECULES AND SELF-ASSEMBLY

Self-test 11.7 Calculate the angle at which the same lattice will give a reflec-

tion from the (123) planes. Answer: 24.8°

(e) Fourier synthesis

Bragg’s law is a very primitive approach to the interpretation of X-ray diffraction data. The huge amount of data obtained from a modern diffractometer has a much richer content than the separation of lattice planes, for in principle it contains information about the locations of individual atoms and the distribution of electron density throughout a unit cell. To derive the structure of the crystal from the intensities, Ihkl, we need to convert them to the amplitude of the wave responsible for the signal. For simplicity we shall focus on a one-dimensional crystal (a line of atoms) and write the diffraction intensities as Ih. Because the intensity of electromagnetic radiation is given by the square of the amplitude, we need to form the structure factors Fh = I h1/2. (Note that this ‘structure factor’ is entirely distinct from the structure factor of Rayleigh scattering, Section 11.3.) Here is the first difficulty: we do not know the sign to take. For instance, if Ih = 4, then Fh can be either +2 or −2. This ambiguity is the phase problem of X-ray diffraction. However, once we have the structure factors, we can calculate the electron density r(x) by forming the following sum: A brief comment

Formally, a Fourier synthesis is a reconstruction of a repetitive function as a superposition of sine or cosine waves. Long-wavelength waves account for the general features of the structure, and the details are gradually filled in by incorporating shorter-wavelength waves.

r(x) =

∞ 1! F0 + 2 ∑ Fh cos(2hpx)# $ V@ h=1

Fourier synthesis

(11.12)

where V is the volume of the unit cell. This expression is called a Fourier synthesis of the electron density: we show how it is used in the following Example. The point to note is that low values of the index h give the major features of the structure (they correspond to long-wavelength cosine terms), whereas the high values give the fine detail (short-wavelength cosine terms). Clearly, if we do not know the sign of Fh, we do not know whether the corresponding term in the sum is positive or negative and we get different electron densities, and hence crystal structures, for different choices of sign. Example 11.5

Calculating an electron density by Fourier synthesis

The determination of the three-dimensional structure of molecules is a key step in the rational design of therapeutic agents that bind specifically to receptor sites on proteins and nucleic acids (Case study 11.2). Consider the (h00) planes of a crystal of an organic molecule regarded as a candidate for a drug. In an X-ray analysis the structure factors were found as follows: h Fh

0 16

1 −10

2 2

3 −1

4 7

5 −10

h Fh

10 6

11 −5

12 3

13 −2

14 2

15 −3

6 8

7 −3

8 2

9 −3

Construct a plot of the electron density projected on to the x-axis of the unit cell. Strategy Evaluate the sum in eqn 11.12 (stopping at h = 15) for points

0 ≤ x ≤ 1:

15

Vr(x) = 16 + 2 ∑ Fh cos(2hpx) h=1

11.4 X-RAY CRYSTALLOGRAPHY

421

The task is made easier by using an electronic spreadsheet, which also can generate a plot of the results. Solution After introducing the data, eqn 11.12 takes the form

Vr(x) = 16 − 20 cos(2px) + 4 cos(4px) − · · · − 6 cos(30px) This function is shown in Fig. 11.18(a), and the locations of several types of atom are easy to identify as peaks in the electron density. The more terms there are included, the more accurate the density plot. Terms corresponding to high values of h (short-wavelength cosine terms in the sum) account for the finer details of the electron density; low values of h account for the broad features. Use an electronic spreadsheet to experiment with different structure factors (including changes in signs as well as amplitudes). For example, use the same values of Fh as above, but with positive signs for all values of h.

Self-test 11.8

Answer: Fig. 11.18(b)

The phase problem can be overcome to some extent by the method of isomorphous replacement, in which heavy atoms are introduced into the crystal. The technique relies on the fact that the scattering of X-rays is caused by the oscillations an incoming electromagnetic wave generates in the electrons of atoms, and heavy atoms with their large numbers of electrons give rise to stronger scattering than light atoms. Therefore, heavy atoms dominate the diffraction pattern and greatly simplify its interpretation. The phase problem can also be resolved by judging whether the calculated structure is chemically plausible, whether the electron density is positive throughout, and by using more refined mathematical techniques, with the help of powerful computers. Because biopolymers contain a great many atoms, overcoming the phase problem requires repeated rounds of isomorphous replacement and computeraided refinement, a process that can take several years to complete. As suggested by eqn 11.12 and Example 11.5, the more values of Ihkl that are collected, the richer the detail of the structure: analyzing few intensities leads to a fuzzy, low-resolution structure, whereas collecting more reflections results in a sharper, high-resolution structure. In practice, it is not the abundance of data, but rather the quality of the crystal—as determined by how perfectly ordered the molecules are packed in the solid—that limits the resolution of a structure. With current crystallization techniques, the best resolution of protein structures is approximately 200 pm, implying that two atoms cannot be located unambiguously if they are separated by less than this distance, which is greater than the average length of a carbon–carbon single bond (154 pm). In spite of this limitation, the identity and location of every atom in a biopolymer can be obtained by combining X-ray diffraction and sequencing data. In the laboratory 11.1

The crystallization of biopolymers

The first and often very demanding step in the structural analysis of biological macromolecules by X-ray diffraction methods is to form crystals in which the large molecules lie in orderly ranks. A technique that works well for charged proteins consists of adding large amounts of a salt, such as (NH4)2SO4, to a buffer solution containing the biopolymer. The increase in the ionic strength

Fig. 11.18 The plot of the electron density calculated in (a) Example 11.5 and (b) Self-test 11.8.

422

11 MACROMOLECULES AND SELF-ASSEMBLY

of the solution decreases the solubility of the protein to such an extent that the protein precipitates, sometimes as crystals that are amenable to analysis by X-ray diffraction (see Exercise 5.14 for an explanation of this effect).

Fig. 11.19 In a common implementation of the vapor diffusion method of biopolymer crystallization, a single drop of biopolymer solution hangs above a reservoir solution that is very concentrated in a nonvolatile solute. Solvent evaporates from the more dilute drop until the vapor pressure of water in the closed container reaches a constant equilibrium value. In the course of evaporation (denoted by the downward arrows), the biopolymer solution becomes more concentrated and, at some point, crystals may form.

Other common strategies for inducing crystallization involve the gradual removal of solvent from a biopolymer solution, either by dialysis (Section 3.10) or vapor diffusion. In one implementation of the vapor diffusion method, a single drop of biopolymer solution hangs above an aqueous solution (the reservoir), as shown in Fig. 11.19. If the reservoir solution is more concentrated in a nonvolatile solute (for example, a salt) than is the biopolymer solution, then solvent will evaporate slowly from the drop until the vapor pressure of water in the closed container reaches a constant, equilibrium value. At the same time, the concentration of biopolymer in the drop increases gradually until crystals begin to form. Special techniques are used to crystallize hydrophobic proteins, such as those spanning the bilayer of a cell membrane. In such cases, surfactant molecules, which, like phospholipids, contain polar head groups and hydrophobic tails, are used to encase the protein molecules and make them soluble in aqueous buffer solutions. Dialysis or vapor diffusion may then be used to induce crystallization.

In the laboratory 11.2

Data acquisition in X-ray crystallography

After suitable crystals are obtained, X-ray diffraction data are collected and analyzed. Laue’s original method consisted of passing a beam of X-rays of a wide range of wavelengths into a single crystal and recording the diffraction pattern photographically. The idea behind the approach was that a crystal might not be suitably orientated to act as a diffraction grating for a single wavelength, but whatever its orientation Bragg’s law would be satisfied for at least one of the wavelengths when a range of wavelengths is present in the beam. An alternative technique was developed by Peter Debye and Paul Scherrer and independently by Albert Hull. They used monochromatic (single-frequency) X-rays and a powdered sample. When the sample is a powder, we can be sure that some of the randomly distributed crystallites will be orientated so as to satisfy Bragg’s law. For example, some of them will be orientated so that their (111) planes, of spacing d, give rise to a reflection at a particular angle, and others will be orientated so that their (230) planes give rise to a reflection at a different angle. Each set of (hkl) planes gives rise to reflections at a different angle. In the modern version of the technique, which uses a powder diffractometer, the sample is spread on a flat plate and the diffraction pattern is monitored electronically. The major application is for qualitative analysis because the diffraction pattern is a kind of fingerprint and may be recognizable (Fig. 11.20). The technique is also used for the characterization of substances that cannot be crystallized or the initial determination of the dimensions and symmetries of unit cells. Modern X-ray crystallography, which utilizes an X-ray diffractometer (Fig. 11.21), is now a highly sophisticated technique. By far the most detailed information comes from developments of the techniques pioneered by the Braggs, in which a single crystal is employed as the diffracting object and a

11.4 X-RAY CRYSTALLOGRAPHY

423

monochromatic beam of X-rays is used to generate the diffraction pattern. The single crystal (which may be only a fraction of a millimeter in length) is rotated relative to the beam, and the diffraction pattern is monitored and recorded electronically for each crystal orientation. The primary data are therefore a set of intensities arising from the Miller planes (hkl), with each set of planes giving a reflection of intensity Ihkl.

Case study 11.1

The structure of DNA from X-ray diffraction studies

Bragg’s law helps us understand the features of one of the most seminal X-ray images of all time, the characteristic X-shaped pattern obtained by Rosalind Franklin and Maurice Wilkins from strands of DNA, and used by James Watson and Francis Crick in their construction of the double-helix model of DNA (Fig. 11.22). To interpret this image by using Bragg’s law, we have to be aware that it was obtained by using a fiber consisting of many DNA molecules oriented with their axes parallel to the axis of the fiber, with X-rays incident from a perpendicular direction. All the molecules in the fiber are parallel (or nearly so) but are randomly distributed in the perpendicular directions; as a result, the diffraction pattern exhibits the periodic structure parallel to the fiber axis superimposed on a general background of scattering from the distribution of molecules in the perpendicular directions. There are two principal features in Fig. 11.22: the strong ‘meridional’ scattering upward and downward by the fiber and the X-shaped distribution at smaller scattering angles. Because scattering through large angles occurs for closely spaced features (from l = 2d sin q, if d is small then q has to be large to preserve the equality), we can infer that the meridional scattering arises from closely spaced components and that the inner X-shaped pattern arises from features with a longer periodicity. Because the meridional pattern occurs at a distance of about 10 times that of the innermost spots of the X pattern, the large-scale structure is about 10 times bigger than the small-scale structure. From the geometry of the instrument, the wavelength of the radiation, and Bragg’s law, we can infer that the periodicity of the small-scale feature is 340 pm, whereas that of the large-scale feature is 3400 pm (that is, 3.4 nm). To see that the cross is characteristic of a helix, look at Fig. 11.22. Each turn of the helix defines two planes, one orientated at an angle a to the horizontal and the other at −a. As a result, to a first approximation, a helix can be thought of as consisting of an array of planes at an angle a together with an array of planes at an angle −a with a separation within each set determined by the pitch of the helix. Thus, a DNA molecule is like two arrays of planes, each set corresponding to those treated in the derivation of the Bragg law, with a perpendicular separation d = p cos a, where p is the pitch of the helix, each canted at the angles ±a to the horizontal. The diffraction spots from one set of planes therefore occur at an angle a to the vertical, giving one leg of the X, and those of the other set occur at an angle −a, giving rise to the other leg of the X. The experimental arrangement has up–down symmetry, so the diffraction pattern repeats to produce the lower half of the X. The sequence of spots outward along a leg corresponds to first-, second-, . . . order diffraction (n = 1, 2, . . . in eqn 11.11b). Therefore from the X-ray pattern, we see at once that the molecule is helical and we can measure the angle directly and find a = 40°. Finally, with the angle

Fig. 11.20 A typical X-ray powder diffraction pattern that can be used to identify the material and determine the size of its unit cell: (a) NaCl, (b) KCl.

Fig. 11.21 The geometry of a four-circle diffractometer. The settings of the orientations of the components are controlled by computer; each reflection is monitored in turn, and their intensities are recorded.

424

11 MACROMOLECULES AND SELF-ASSEMBLY

Fig. 11.22 The origin of the X pattern characteristic of diffraction by a helix. (a) A helix can be thought of as consisting of an array of planes at an angle a together with an array of planes at an angle –a. (b) The diffraction spots from one set of planes appear at an angle a to the vertical, giving one leg of the X, and those of the other set appear at an angle −a, giving rise to the other leg of the X. The lower half of the X appears because the helix has up–down symmetry in this arrangement. (c) The sequence of spots outward along a leg of the X corresponds to first-, second-, . . . order diffraction (n = 1, 2, . . .). Fig. 11.23 The effect of the internal structure of the helix on the X-ray diffraction pattern. (a) The residues of the macromolecule are represented by points. (b) Parallel planes passing through the residues are perpendicular to the axis of the molecule. (c) The planes give rise to strong diffraction with an angle that allows us to determine the layer spacing h from l = 2h sin q.

a and the pitch p determined, we can determine the radius r of the helix from tan a = p/4r, from which it follows that r = (3.4 nm)/(4 tan 40°) = 1.0 nm. To derive the relation between the helix and the cross-like pattern, we have ignored the detailed structure of the helix, the fact that it is a periodic array of nucleotide bases, not a smooth wire. In Fig. 11.23 we represent the bases by points and see that there is an additional periodicity of separation h, forming planes that are perpendicular to the axis of the molecule (and the fiber). These planes give rise to the strong meridional diffraction with an angle that allows us to determine the layer spacing from Bragg’s law in the form l = 2d sin q as h = 340 pm.

The control of shape The conformation of a biological molecule that has been determined by one of the techniques described so far is of crucial importance for its function, and we need to understand the forces that bring about the shape we observe. Interactions between molecules include the attractive and repulsive interactions between the partial electric charges of polar molecules and of polar functional groups in

11.5 INTERACTIONS BETWEEN PARTIAL CHARGES

425

macromolecules and the repulsive interactions that prevent the complete collapse of matter to densities as high as those characteristic of atomic nuclei. The repulsive interactions arise from the exclusion of electrons from regions of space where the orbitals of closed-shell species overlap. One class of interaction, those proportional to the inverse sixth power of the separation, are termed van der Waals interactions. However, these are not the only interactions, and in the following paragraphs we describe the principal nonbonding interactions that occur between molecules and between different parts of the same molecule. All these interactions are much weaker—in some cases by several orders of magnitude—than those responsible for the formation of chemical bonds. 11.5 Interactions between partial charges The Coulomb interaction between charges is our starting point for the discussion of the assembly of biological structures.

Atoms in molecules in general have partial charges. Table 11.2 gives the partial charges typically found on the atoms in peptides. If these charges were separated by a vacuum, they would attract or repel each other in accordance with Coulomb’s law (Fundamentals F.3), and we would write V=

Q1Q2 4pε0r

Coulomb’s law (vacuum)

(11.13a)

where Q1 and Q2 are the partial charges and r is their separation. However, we should take into account the possibility that other parts of the molecule, or other molecules, lie between the charges and decrease the strength of the interaction. We therefore write QQ V= 1 2 4pεr

Coulomb’s law (in any medium)

Partial charges in polypeptides Table 11.2

Atom

Partial charge/e

C(=O)

+0.45

C(–CO)

+0.06

H(–C)

+0.02

H(–N)

+0.18

H(–O)

+0.42

N

−0.36

O

−0.38

(11.13b)

where ε is the permittivity of the medium lying between the charges. The permittivity is usually expressed as a multiple of the vacuum permittivity by writing ε = εrε0, where εr is the relative permittivity (formerly known as the dielectric constant). The effect of the medium can be very large: for water εr = 78, so the potential energy of two charges separated by bulk water is reduced by nearly two orders of magnitude compared to the value it would have if the charges were separated by a vacuum (Fig. 11.24). The problem is made worse in calculations on polypeptides and nucleic acids by the fact that two partial charges may have water and a biopolymer chain lying between them. Various models have been proposed to take this awkward effect into account, the simplest being to set εr = 3.5 and to hope for the best. A brief illustration

The energy of interaction between a partial charge of −0.36 (that is, Q1 = −0.36e) on the N atom of a peptide link and the partial charge of +0.45 (Q2 = +0.45e) on the carbonyl C atom at a distance of 3.0 nm on the assumption that the medium between them is a vacuum is V=

(−0.36e) × (0.45e) −0.36 × 0.45 × (1.602 × 10−19 C)2 = 4pε0 × (3.0 nm) 4p × (8.854 × 10−12 J−1 C2 m−1) × (3.0 × 10−9 m) = −1.2 × 10−20 J

Fig. 11.24 The Coulomb potential energy of two charges Q1 and Q2 and its dependence on their separation. The two curves correspond to different relative permittivities (εr =1 for a vacuum, 3 for a fluid).

426

11 MACROMOLECULES AND SELF-ASSEMBLY This energy (after multiplication by Avogadro’s constant) corresponds to −7.5 kJ mol−1. However, if the medium has a ‘typical’ relative permittivity of 3.5, then the interaction energy is reduced to −2.1 kJ mol−1. For bulk water as the medium, with the H2O molecules able to rotate in response to a field, the energy of interaction would be reduced by a factor of 78, to only −0.96 kJ mol−1.

11.6 Electric dipole moments Many physical and chemical properties are related to the distribution of partial charges in a molecule or group (such as the peptide group), and here we start to identify them.

Dipole moments and mean polarizability volumes

Table 11.3

m/D

a′/(10−30 m3)

Ar

1.66

CCl4

10.3

C6H6

10.4

H2

0.819

H2O

1.85

1.48

NH3

1.47

2.22

HCl

1.08

2.63

HBr

0.80

3.61

HI

0.42

5.45

At its simplest, an electric dipole consists of two charges Q and −Q separated by a distance l. The product Ql is called the electric dipole moment, m. We represent dipole moments by an arrow with a length proportional to m and pointing from the negative charge to the positive charge (1).2 Because a dipole moment is the product of a charge (in coulombs, C) and a length (in meters, m), the SI unit of dipole moment is the coulomb meter (C m). However, it is often much more convenient to report a dipole moment in the non-SI unit debye, D, where 1 D = 3.335 × 10−30 C m, because experimental values for molecules are then close to 1 D (Table 11.3).3 The dipole moment of two charges e and −e separated by 100 pm is 1.6 × 10−29 C m, corresponding to 4.8 D. Dipole moments of small molecules are typically smaller than that, at about 1 D. A polar molecule is a molecule with a permanent electric dipole moment arising from the partial charges on its atoms (Section 10.5). A nonpolar molecule is a molecule that has no permanent electric dipole moment. All heteronuclear diatomic molecules are polar because the difference in electronegativities of their two atoms results in nonzero partial charges. Typical dipole moments are 1.08 D for HCl and 0.42 D for HI (Table 11.3). A very approximate relation between the dipole moment and the difference in Pauling electronegativities (Table 10.2) of the two atoms, Dc, is m/D ≈ Dc

Relation between dipole moment and electronegativity

(11.14)

A brief illustration

The electronegativities of hydrogen and bromine are 2.1 and 2.8, respectively. The difference is 0.7, so we predict an electric dipole moment of about 0.7 D for HBr. The experimental value is 0.80 D.

Because it attracts the electrons more strongly, the more electronegative atom is usually the negative end of the dipole. However, there are exceptions, particularly when antibonding orbitals are occupied. Thus, the dipole moment of NO is very small (0.07 D), but the negative end of the dipole is on the N atom even 2 Be careful with this convention: for historical reasons the opposite convention is still widely adopted. 3 The unit is named after Peter Debye, the Dutch pioneer of the study of dipole moments of molecules.

11.6 ELECTRIC DIPOLE MOMENTS

427

Mathematical toolbox 11.1 Addition and subtraction of vectors

Consider two vectors 1 and 2 making an angle q shown below

The first step in the addition of 2 to 1 consists of joining the tail of 2 to the head of 1:

The subtraction of vectors follows the same principles outlined above for addition by noting that subtraction of 2 from 1 amounts to addition of −2 to 1:

In the second step, we draw a vector res, the resultant vector, originating from the tail of 1 to the head of 2:

though the O atom is more electronegative. This apparent paradox is resolved as soon as we realize that antibonding orbitals are occupied in NO (see Fig. 10.35) and because electrons in antibonding orbitals tend to be found closer to the less electronegative atom, they contribute a negative partial charge to that atom. If this contribution is larger than the opposite contribution from the electrons in bonding orbitals, then the net effect will be a small negative partial charge on the less electronegative atom. Molecular symmetry is of the greatest importance in deciding whether a polyatomic molecule is polar or not. Indeed, molecular symmetry is more important than the question of whether or not the atoms in the molecule belong to the same element. hom*onuclear polyatomic molecules may be polar if they have low symmetry and the atoms are in inequivalent positions. For instance, the angular molecule ozone, O3 (2), is hom*onuclear; however, it is polar because the central O atom is different from the outer two (it is bonded to two atoms, they are bonded only to one). Moreover, the dipole moments associated with each bond make an angle to each other and do not cancel (see Mathematical toolkit 11.1). Heteronuclear polyatomic molecules may be nonpolar if they have high symmetry, because individual bond dipoles may then cancel. The heteronuclear linear triatomic molecule CO2, for example, is nonpolar because, although there are partial charges on all three atoms, the dipole moment associated with the OC bond points in the opposite direction to the dipole moment associated with the CO bond, and the two cancel (3).

Self-test 11.9 Ozone, carbon dioxide, water, and methane are all components

of the Earth’s atmosphere that absorb heat emanating from the surface of the planet, thus maintaining temperatures consistent with the proliferation of life. Predict whether methane and water molecules are polar or nonpolar. Answer: An H2O molecule is angular and polar; a CH4 molecule is tetrahedral and nonpolar

428

11 MACROMOLECULES AND SELF-ASSEMBLY

A useful approach to the calculation of dipole moments is to take into account the locations and magnitudes of the partial charges on all the atoms. These partial charges are included in the output of many molecular structure software packages. Indeed, the programs calculate the dipole moments of the molecules by noting that an electric dipole moment is actually a vector, m, with three components, mx, my, and mz (4). The direction of m shows the orientation of the dipole in the molecule, and the length of the vector is the magnitude, m, of the dipole moment. In common with all vectors (Mathematical toolkit 9.1), the magnitude is related to the three components by Magnitude of the dipole moment vector

m = (mx2 + m2y + m2z)1/2

(11.15a)

To calculate m, we need to calculate the three components and then substitute them into this expression. To calculate the x-component, for instance, we need to know the magnitude of the partial charge on each atom and the atom’s x-coordinate relative to a point in the molecule and form the sum mx =

∑J QJxJ

Calculation of a component of the dipole moment vector

(11.15b)

Here QJ is the partial charge of atom J, xJ is the x-coordinate of atom J, and the sum is over all the atoms in the molecule. Similar expressions are used for the y- and z-components. For an electrically neutral molecule, the origin of the coordinates is arbitrary, so it is best chosen to simplify the measurements.

Example 11.6

Calculating the dipole moment of the peptide group

Estimate the electric dipole moment of the peptide group (5) by using the partial charges and the locations of the atoms shown in pm. Strategy We use eqn 11.15b to calculate each of the components of the dipole moment. Then we use eqn 11.15a to assemble the three components into the magnitude of the dipole moment. Note that the partial charges are multiples of the fundamental charge e = 1.602 × 10−19 C. Solution The expression for mx is

mx = (−0.36e) × (132 pm) + (0.45e) × (0 pm) + (0.18e) × (182 pm) + (−0.38e) × (−62.0 pm) = 8.8e pm = 8.8 × (1.602 × 10−19 C) × (10−2 m) = 1.4 × 10−30 C m corresponding to mx = +0.42 D. The expression for my is my = (−0.36e) × (0 pm) + (0.45e) × (0 pm) + (0.18e) × (−87 pm) + (−0.38e) × (107 pm) = −56e pm = −9.0 × 10−30 C m It follows that my = −2.7 D. Therefore, because mz = 0, m = {(0.42 D)2 + (−2.7 D)2}1/2 = 2.7 D We can find the orientation of the dipole moment by arranging an arrow of length 2.7 units of length to have x-, y-, and z-components of 0.42, −2.7, and 0 units; the orientation is superimposed on (5).

11.7 INTERACTIONS BETWEEN DIPOLES

Self-test 11.10 Calculate the electric dipole moment of formaldehyde, using the information in (6).

Answer: 3.2 D

11.7 Interactions between dipoles When molecules or groups are widely separated, it is simpler to express their interaction in terms of the dipole moments rather than with each partial charge. We need to know how to handle these interactions because they are important for the assembly of biological macromolecules.

The potential energy of a dipole m1 in the presence of a charge Q2 is calculated by taking into account the interaction of the charge with the two partial charges of the dipole, one resulting in a repulsion and the other an attraction. The result for the arrangement shown in (7) is V=−

Q2m1 4pε0r 2

Charge–dipole interaction energy (as in 7)

(11.16a)

Justification 11.4 The interaction of a charge with a dipole

When the charge and dipole are collinear, as in (7), the potential energy is V= =

=

Q1Q2 Q1Q2 − 1 4pε0(r + 2 l) 4pε0(r − 12 l) Q1Q2 Q1Q2 − A A lD lD 4pε0r 1 + 4pε0r 1 − C C 2rF 2rF 1 D Q1Q2 A 1 − E B 4pε0r B l l 1+ 1− E C 2r 2r F

Next, we suppose that the separation of charges in the dipole is much smaller than the distance of the charge Q2 in the sense that l/2r 54.7° because then like charges are closer than unlike charges. The potential energy is zero on the lines at 54.7° and 180° − 54.7° = 125.3° because at those angles the two attractions and the two repulsions cancel (10). The average potential energy of interaction between polar molecules that are freely rotating in a fluid (a gas or liquid) is zero because the attractions and 4

For a derivation of eqn 11.17, see our Physical chemistry (2010).

11.8 INDUCED DIPOLE MOMENTS

431

repulsions cancel. However, because the potential energy of a dipole near another dipole depends on their relative orientations, the molecules exert forces on each other and therefore do not in fact rotate completely freely, even in a gas. As a result, the lower energy orientations are marginally favored, so there is a nonzero interaction between rotating polar molecules (Fig. 11.25). The detailed calculation of the average interaction energy is quite complicated, but the final answer is very simple: V=−

2m12 m22 3(4pε0)2kTr 6

Average dipole–dipole interaction energy (freely rotating dipoles)

(11.18)

The important features of this expression are the dependence of the average interaction energy on the inverse sixth power of the separation (which identifies it as a van der Waals interaction) and its inverse dependence on the temperature. The temperature dependence reflects the way that the greater thermal motion overcomes the mutual orientating effects of the dipoles at higher temperatures. Equation 11.18 is applicable when both molecules are free to rotate or when one is fixed and only the other is free to rotate, as for a small polar molecule near a macromolecule. A brief illustration

Suppose a water molecule (m = 1.85 D) can rotate freely at 1.0 nm from a peptide group (m = 2.7 D): the energy of their interaction at 25°C (298 K) is V=−

2 × (1.85 × 3.336 × 10−30 C m)2 × (2.7 × 3.336 × 10−30 C m)2 3(4p × (8.854 × 10−12 J −1 C2 m−1)2 × 1.381 × 10−23 J K −1 × 298 K × (1.0 × 10−9 m)6

= −4.0 × 10−23

C4 m4 J −2 C4 m−2 J K−1 K m6

= −4.0 × 10−23 J This interaction energy corresponds (after multiplication by Avogadro’s constant) to −24 J mol−1. When the temperature is raised to body temperature, 37°C (310 K), the H2O molecule rotates more vigorously and the average interaction is reduced to −23 J mol−1.

11.8 Induced dipole moments The structures and properties of biological assemblies also emerge from interactions that involve nonpolar species, such as nonpolar groups on the peptide residues of a protein.

A nonpolar molecule may acquire a temporary induced dipole moment, m*, as a result of the influence of an electric field generated by a nearby ion or polar molecule. The field distorts the electron distribution of the molecule and gives rise to an electric dipole in it. The molecule is said to be polarizable. The magnitude of the induced dipole moment is proportional to the strength of the electric field, E, and we write m* = aE

A dipole–dipole interaction. When a pair of molecules can adopt all relative orientations with equal probability, the favorable orientations and the unfavorable ones cancel, and the average interaction is zero. In an actual fluid, the favorable interactions slightly predominate.

Fig. 11.25

Polarizability

(11.19)

The proportionality constant a is the polarizability of the molecule. The larger the polarizability of the molecule, the greater is the distortion caused by a given

A note on good practice

Note how the units are included in the calculation and cancel to give the result in joules. It is far better to include the units at each stage of the calculation and treat them as algebraic quantities that can be multiplied and canceled than to guess the units at the end of the calculation.

432

11 MACROMOLECULES AND SELF-ASSEMBLY

strength of electric field. If the molecule has few electrons, they are tightly controlled by the nuclear charges and the polarizability of the molecule is low. If the molecule contains large atoms with electrons some distance from the nucleus, the nuclear control is less and the polarizability of the molecule is greater. The polarizability also depends on the orientation of the molecule with respect to the field unless the molecule is tetrahedral (such as CCl4), octahedral (such as SF6), or icosahedral (such as C60). Atoms, tetrahedral, octahedral, and icosahedral molecules have isotropic (orientation-independent) polarizabilities; all other molecules have anisotropic (orientation-dependent) polarizabilities. The polarizabilities reported in Table 11.3 are given as polarizability volumes, a′: a′ =

a 4pε0

Polarizability volume

(11.20)

The polarizability volume has the dimensions of volume (hence its name) and is comparable in magnitude to the volume of the molecule. Self-test 11.11 What strength of electric field is required to induce an electric dipole moment of 1.0 mD in a molecule of polarizability volume 2.6 × 10−30 m3 (like CO2)?

Answer: 11 kV m−1

(a) Dipole–induced-dipole interactions

A polar molecule with dipole moment m1 can induce a dipole moment in a polarizable molecule (which may itself be either polar or nonpolar) because the partial charges of the polar molecule give rise to an electric field that distorts the second molecule. That induced dipole interacts with the permanent dipole of the first molecule, and the two are attracted together (Fig. 11.26). The formula for the dipole–induced-dipole interaction energy is V=−

m12a2′ 4pε0r 6

Dipole–induced-dipole interaction energy

(11.21)

where a2 is the polarizability of molecule 2. The negative sign shows that the interaction is attractive. For a molecule with m = 1 D (such as HCl) near a molecule of polarizability volume a′ = 1.0 × 10−29 m3 (such as benzene, Table 11.3), the average interaction energy is about −0.8 kJ mol−1 when the separation is 0.3 nm. (b) Dispersion interactions

Fig. 11.26 A dipole–induceddipole interaction. The induced dipole follows the changing orientation of the permanent dipole.

Despite the absence of partial charges, we know that uncharged, nonpolar species can interact because they form condensed phases, such as benzene, liquid hydrogen, and liquid xenon. The dispersion interaction, or London interaction, between nonpolar species arises from the transient dipoles that they possess as a result of fluctuations in the instantaneous positions of their electrons (Fig. 11.27). Suppose, for instance, that the electrons in one molecule flicker into an arrangement that results in partial positive and negative charges and thus gives it an instantaneous dipole moment m1. While it exists, this dipole can polarize the other molecule and induce in it an instantaneous dipole moment m2. The two dipoles attract each other and the potential energy of the pair is lowered. Although the first molecule will go on to change the size and direction of its dipole (perhaps

11.9 HYDROGEN BONDING

433

within 10−16 s), the second will follow it; that is, the two dipoles are correlated in direction like two meshing gears, with a positive partial charge on one molecule appearing close to a negative partial charge on the other molecule and vice versa. Because of this correlation of the relative positions of the partial charges, and their resulting attractive interaction, the attraction between the two instantaneous dipoles does not average to zero. Instead, it gives rise to a net attractive interaction. Polar molecules interact by a dispersion interaction as well as by dipole–dipole interactions. The strength of the dispersion interaction depends on the polarizability of the first molecule because the magnitude of the instantaneous dipole moment m1 depends on the looseness of the control that the nuclear charge has over the outer electrons. If that control is loose, the electron distribution can undergo relatively large fluctuations. Moreover, if the control is loose, then the electron distribution can also respond strongly to applied electric fields and hence have a high polarizability. It follows that a high polarizability is a sign of large fluctuations in local charge density. The strength also depends on the polarizability of the second molecule, for that polarizability determines how readily a dipole can be induced in molecule 2 by molecule 1. We therefore expect that V ∝ a1a2. The actual calculation of the dispersion interaction is quite involved, but a reasonable approximation to the interaction energy is the London formula: V = − 32 ×

a1′a 2′ I1I2 × r 6 I1 + I2

London formula

(11.22)

where I1 and I2 are the ionization energies of the two molecules. A brief illustration

If two phenylalanine residues are separated by 3.0 nm in a polypeptide, the dispersion interaction between their phenyl groups is calculated from eqn 11.22 by setting a1′ = a2′ and I1 = I2 = I: V = − 34 ×

a1′2 ×I r6

We treat the phenyl groups as benzene rings of polarizability volume 1.0 × 10−29 m3: V = − 34 ×

(1.0 × 10−29 m3)2 × I = −1.0 × 10−7 × I (3.0 × 10−9 m)6

If we suppose that the ionization energy of the phenyl group is about 5 eV (about 500 kJ mol−1), this energy is approximately −23 mJ mol−1.

11.9 Hydrogen bonding Strong interactions of the type X–H···Y (with X, Y = N or O) are responsible for the formation of well-defined three-dimensional structures in proteins and nucleic acids. We need to understand the origin of the strength of these very important interactions.

The strongest intermolecular interaction arises from the formation of a hydrogen bond, in which a hydrogen atom lies between two strongly electronegative atoms and binds them together. The bond is normally denoted X–H···Y, with X and Y

In the dispersion interaction, an instantaneous dipole on one molecule induces a dipole on another molecule, and the two dipoles then interact to lower the energy. The directions of the two instantaneous dipoles are correlated and, although they occur in different orientations at different instants, the interaction does not average to zero.

Fig. 11.27

434

11 MACROMOLECULES AND SELF-ASSEMBLY

being N, O, or F. Unlike the other interactions we have considered, hydrogen bonding is not universal but is restricted to molecules that contain these atoms. The most elementary description of the formation of a hydrogen bond is that it is the result of a Coulombic interaction between the partly exposed positive charge of a proton bound to an electron-withdrawing X atom (in the fragment X–H) and the negative charge of a lone pair on the second atom Y, as in d−X–Hd+ :Y d−. A slightly more sophisticated version of the electrostatic description is to regard hydrogen bond formation as the formation of a Lewis acid–base complex in which the partly exposed proton of the X–H group is the Lewis acid and :Y, with its lone pair, is the Lewis base, as in X–H + :Y → X–H:Y. A brief illustration

A common hydrogen bond is that formed between O–H groups and O atoms, as in liquid water and ice. In Exercise 11.42, you are invited to use the electrostatic model to calculate the dependence of the potential energy of interaction on the OOH angle, denoted q in (11), and the results are plotted in Fig. 11.28. We see that at q = 0 when the OHO atoms lie in a straight line; the molar potential energy is −19 kJ mol−1. Note how sharply the energy depends on angle: it is negative only with ±12° of linearity.

Molecular orbital theory provides an alternative description that is more in line with the concept of delocalized bonding and the ability of an electron pair to bind more than one pair of atoms (Section 10.6). Thus, if the X–H bond is regarded as formed from the overlap of an orbital on X, yX, and a hydrogen 1s orbital, yH, and the lone pair on Y occupies an orbital on Y, yY, then when the two molecules are close together, we can build three molecular orbitals from the three basis orbitals: y = c1yX + c2yH + c3yY

Fig. 11.28 The variation of the energy of interaction (on the electrostatic model) of a hydrogen bond as the angle between the O–H and :O groups is changed.

One of the molecular orbitals is bonding, one almost nonbonding, and the third antibonding (Fig. 11.29). These three orbitals need to accommodate four electrons (two from the original X–H bond and two from the lone pair of Y), so two enter the bonding orbital and two enter the nonbonding orbital. Because the antibonding orbital remains empty, the net effect—depending on the precise location of the almost nonbonding orbital—may be a lowering of energy. Experimental evidence and theoretical arguments have been presented in favor of both the electrostatic and molecular orbital view of hydrogen bonding. For example, recent experiments suggest that the hydrogen bonds in ice have significant covalent character, providing support for the molecular orbital treatment. However, the matter has not yet been resolved. Hydrogen bond formation dominates all other interactions between electrically neutral molecules when it can occur (Table 11.4). It has a typical strength of the order of 20 kJ mol−1, as can be inferred from the enthalpy of vaporization of water, 40.7 kJ mol−1, for vaporization involves the breaking of two hydrogen bonds to each water molecule. Hydrogen bonding accounts for the rigidity of molecular solids such as sucrose and ice; the low vapor pressure, high viscosity, and surface tension of liquids such as water; the secondary structure of proteins (the formation of helices and sheets of polypeptide chains); the structure of DNA and hence

11.10 THE TOTAL INTERACTION

Table 11.4

435

Interaction potential energies

Interaction type

Distance dependence of potential energy

Typical energy (kJ mol−1)

Ion–ion

1/r

250

Hydrogen bond

20

Comment

Only between ions Occurs in X–H· · ·Y, where X, Y = N, O, or F

Ion–dipole

1/r 2

15

Dipole–dipole

1/r 3

2

Between stationary polar molecules

1/r 6

0.6

Between rotating polar molecules

1/r 6

2

Between all types of molecules and ions

London (dispersion)

the transmission of genetic information; and the attachment of drugs to receptors sites in proteins (Case study 11.2). Hydrogen bonding also contributes to the solubility in water of species such as ammonia and compounds containing hydroxyl groups and to the hydration of anions. In this last case, even ions such as Cl− and HS− can participate in hydrogen bond formation with water, for their charge enables them to interact with the hydroxylic protons of H2O. 11.10 The total interaction To treat the myriad interactions in biological assemblies quantitatively, we need simple formulas that express the strengths of the attractions and repulsions.

Fig. 11.29 A schematic portrayal of the molecular orbitals that can be formed from an X, H, and Y orbital and that give rise to an X–H···Y hydrogen bond. The lowest-energy combination is fully bonding, the next nonbonding, and the uppermost is antibonding. The antibonding orbital is not occupied by the electrons provided by the X–H bond and the :Y lone pair, so the configuration shown may result in a net lowering of energy in certain cases (namely, when the X and Y atoms are N, O, or F).

Table 11.4 summarizes the strengths and distance dependence of the attractive interactions that we have considered so far. The total attractive interaction energy between rotating molecules that cannot participate in hydrogen bonding is the sum of the contributions from the dipole–dipole, dipole–induced-dipole, and dispersion interactions. Only the dispersion interaction contributes if both molecules are nonpolar. All three interactions vary as the inverse sixth power of the separation, so we may write V=−

C r6

(11.23)

where C is a coefficient that depends on the identity of the molecules and the type of interaction between them. As we have remarked, the energy of a hydrogen bond X–H···Y is typically 20 kJ mol−1 and occurs on contact for X, Y = N, O, or F. Repulsive terms become important and begin to dominate the attractive forces when molecules are squeezed together (Fig. 11.30), for instance, during the impact of a collision, under the force exerted by a weight pressing on a substance, or simply as a result of the attractive forces drawing the molecules together. These repulsive interactions arise in large measure from the Pauli exclusion principle, which forbids pairs of electrons being in the same region of space. The repulsions increase steeply with decreasing separation in a way that can be deduced only by very extensive, complicated molecular structure calculations. In many cases, however, progress can be made by using a greatly simplified representation of

Fig. 11.30 The general form of an intermolecular potential energy curve (the graph of the potential energy of two closed shell species as the distance between them is changed). The attractive (negative) contribution has a long range, but the repulsive (positive) interaction increases more sharply once the molecules come into contact.

436

11 MACROMOLECULES AND SELF-ASSEMBLY

the potential energy, where the details are ignored and the general features are expressed by a few adjustable parameters. One such approximation is to express the short-range repulsive potential energy as inversely proportional to a high power of r: V=+

C* rn

(11.24)

where C* is another constant (the asterisk signifies repulsion). Typically, n is set equal to 12, in which case the repulsion dominates the 1/r 6 attractions strongly at short separations because then C*/r 12 >> C/r 6. The sum of the repulsive interaction with n = 12 and the attractive interaction given by eqn 11.23 is called the Lennard-Jones (12,6) potential. It is normally written in the form V = 4ε ! The Lennard-Jones potential is another approximation to the true intermolecular potential energy curves. It models the attractive component by a contribution that is proportional to 1/r 6 and the repulsive component by a contribution that is proportional to 1/r12. Specifically, these choices result in the Lennard-Jones (12,6) potential. Although there are good theoretical reasons for the former, there is plenty of evidence to show that 1/r12 is only a very poor approximation to the repulsive part of the curve.

A s D 12 A s D 6 # − @C r F C r F $

Fig. 11.31

Lennard-Jones (12,6) potential

(11.25)

and is drawn in Fig. 11.31. The two parameters are ε (epsilon), the depth of the well, and s, the separation at which V = 0; some typical values are listed in Table 11.5. The well minimum occurs at r = 21/6s. Although the (12,6) potential has been used in many calculations, there is plenty of evidence to show that 1/r 12 is a very poor representation of the repulsive potential and that the exponential form e−r/s is superior. An exponential function is more faithful to the exponential decay of atomic wavefunctions at large distances and hence to the distance dependence of the overlap that is responsible for repulsion. However, a disadvantage of the exponential form is that it is slower to compute, which is important when considering the interactions between the large numbers of atoms in liquids and macromolecules. A further computational advantage of the (12,6) potential is that once r 6 has been calculated, r 12 is obtained simply by taking the square.

Self-test 11.12 At what separation does the minimum of the potential energy

Lennard-Jones parameters for the (12,6) potential

curve occur for a Lennard-Jones potential? Hint: Solve for r after setting the first derivative of the potential energy function to zero.

Table 11.5

ε/(kJ mol−1)

Answer: r = 21/6s s/pm

Ar

128

342

Br2

536

427

C6H6

454

527

Cl2

368

412

H2

34

297

He

11

258

Xe

236

406

With the advent of atomic force microscopy (AFM), in which the force between a molecular sized probe and a surface is monitored (see In the laboratory 9.2), it has become possible to measure directly the forces acting between molecules. The force, F, is the negative slope of the potential energy (F = −dV/dr), so for a Lennard-Jones potential between individual molecules we write F=−

dV 24ε ! A s D 13 A s D 7 # = 2 − C rF $ dr s @ C rF

(11.26)

The net attractive force is greatest (from dF/dr = 0) at r = (267)1/6s, or 1.244s, and at that distance is equal to −144(267 )7/6ε/13s, or −2.397ε/s. For typical parameters, the magnitude of this force is about 10 pN.

11.10 THE TOTAL INTERACTION

Case study 11.2

Molecular recognition in biology and pharmacology

Molecular interactions are responsible for the assembly of many biological structures. Hydrogen bonding and hydrophobic interactions are primarily responsible for the three-dimensional structures of biopolymers, such as proteins, nucleic acids, and cell membranes. The binding of a ligand, or guest, to a biopolymer, or host, is also governed by molecular interactions. Examples of biological host–guest complexes include enzyme–substrate complexes, antigen–antibody complexes, and drug–receptor complexes. In all these cases, a site on the guest contains functional groups that can interact with complementary functional groups of the host. For example, a hydrogen bond donor group of the guest must be positioned near a hydrogen bond acceptor group of the host for tight binding to occur. It is generally true that many specific intermolecular contacts must be made in a biological host–guest complex and, as a result, a guest binds only hosts that are chemically similar. The strict rules governing molecular recognition of a guest by a host control every biological process, from metabolism to immunological response, and provide important clues for the design of effective drugs for the treatment of disease. Interactions between nonpolar groups can be important in the binding of a guest to a host. For example, many enzyme active sites have hydrophobic pockets that bind nonpolar groups of a substrate. Coulombic interactions can be important in the interior of a biopolymer host, where the relative permittivity can be much lower than that of the aqueous exterior. For example, at physiological pH, amino acid side chains containing carboxylic acid or amine groups are negatively and positively charged, respectively, and can attract each other. Dipole–dipole interactions are also possible because many of the building blocks of biopolymers are polar, including the peptide link, –CONH– (see Example 11.6). However, hydrogen bonding interactions are by far the most prevalent in biological host–guest complexes. Many effective drugs bind tightly and inhibit the action of enzymes that are associated with the progress of a disease. In many cases, a successful inhibitor will be able to form the same hydrogen bonds with the binding site that the normal substrate of the enzyme can form, except that the drug is chemically inert toward the enzyme. This strategy has been used in the design of drugs for the treatment of HIV-AIDS. Here we describe the properties of a drug that fights HIV infection, highlighting the importance of molecular interactions. For mature HIV particles to form in cells of the host organism, several large proteins encoded by the viral genetic material must be cleaved by a protease enzyme. The drug Crixivan (12) is a competitive inhibitor of HIV protease and has several molecular features that optimize binding to the enzyme’s active site. First, the hydroxyl group highlighted in (12) displaces an H2O molecule that acts as the nucleophile in the hydrolysis of the substrate. Second, the carbon atom to which the key –OH group is bound has a tetrahedral geometry that mimics the structure of the transition state of the peptide hydrolysis reaction. However, the tetrahedral moiety in the drug is not cleaved by the enzyme. Third, the inhibitor is anchored firmly to the active site by a network of hydrogen bonds involving the carbonyl groups of the drug, a water molecule, and peptide NH groups from the enzyme, as shown in (12).

437

438

11 MACROMOLECULES AND SELF-ASSEMBLY

Levels of structure The concept of the ‘structure’ of a macromolecule takes on different meanings at the different levels at which we think about the arrangement of the chain or network of monomers. The term configuration refers to the structural features that can be changed only by breaking chemical bonds and forming new ones. Thus, the chains –A–B–C– and –A–C–B– have different configurations. The term conformation refers to the spatial arrangement of the different parts of a chain, and one conformation can be changed into another by rotating one part of a chain around a bond. In the following sections we explore the molecular interactions responsible for the different levels of structure of biological macromolecules (primary, secondary, etc., as explained in Fundamentals F.1) and assemblies (such as biological membranes; see Fundamentals F.1). We draw from the concepts developed in Sections 11.5–11.10 and describe computational techniques that can help with the prediction of the three-dimensional structure of polypeptides and polynucleotides. 11.11 Minimal order: gases and liquids Many biochemical processes take place in the aqueous intracellular space, so we need to understand the structure of liquids in general and of water in particular.

The form of matter with the least order is a gas. In a perfect gas there are no intermolecular interactions and the distribution of molecules is completely random. In a real gas there are weak attractions and repulsions that have minimal effect on the relative locations of the molecules but that cause deviations from the perfect gas law for the dependence of pressure on the volume, temperature, and amount. Normally there is no need to consider such deviations in biological applications. The attractions between molecules are responsible for the condensation of gases into liquids at low temperatures. First, at low enough temperatures the molecules of a gas have insufficient kinetic energy to escape from each other’s attraction and they stick together. Second, although molecules attract each other when they are a few diameters apart, as soon as they come into contact, they repel

11.11 MINIMAL ORDER: GASES AND LIQUIDS

439

Fig. 11.32 (a) In a perfect crystal at T = 0, the distribution of molecules (or ions) is highly regular, and the radial distribution function has a series of sharp peaks that show the regular organization of rings of neighbors around any selected central molecule or ion. (b) In a liquid, there remain some elements of structure close to each molecule, but the greater the distance, the less the correlation. The radial distribution function now shows a pronounced (but broadened) peak corresponding to the nearest neighbors of the molecule of interest (which are only slightly more disordered than in the solid) and a suggestion of a peak for the next ring of molecules, but little structure at greater distances.

each other. This repulsion is responsible for the fact that liquids and solids have a definite bulk and do not collapse to an infinitesimal point. The molecules are held together by molecular interactions, but their kinetic energies are comparable to their potential energies. As a result, although the molecules of a liquid are not free to escape completely from the bulk, the whole structure is very mobile and we can speak only of the average relative locations of molecules. The average locations of the molecules in a liquid are described in terms of the radial distribution function, g(r). This function is defined so that g(r)dr is the probability that a molecule will be found at a distance between r and r + dr from another molecule.5 It follows that if g(r) passes through a maximum at a radius of, for instance, 0.5 nm, then the most probable distance (regardless of direction) at which a second molecule will be found will be at 0.5 nm from the first molecule. In a crystal, g(r) is an array of sharp spikes, representing the certainty (in the absence of defects and thermal motion) that particles lie at definite locations. This regularity continues out to large distances (to the edge of the crystal, billions of molecules away), so we say that crystals have long-range order. When the crystal melts, the long-range order is lost and wherever we look at long distances from a given particle there is equal probability of finding a second particle. Close to the first particle, however, there may be a remnant of order (Fig. 11.32). Its nearest neighbors might still adopt approximately their original positions, and even if they are displaced by newcomers, the new particles might adopt their vacated positions. It may still be possible to detect, on average, a sphere of nearest neighbors at a distance r1 and perhaps beyond them a sphere of next-nearest neighbors at r2. The existence of this short-range order means that g(r) can be expected to have a broad but pronounced peak at r1, a smaller and broader peak at r2, and perhaps some more structure beyond that. As an illustration, Fig. 11.33 shows the radial distribution function for water at a series of temperatures. The shells of local structure shown are unmistakable. Closer analysis shows that any given H2O molecule is surrounded by other molecules at the corners of a tetrahedron, similar to the arrangement in ice (Fig. 11.34). The form of g(r) at 100°C shows that the intermolecular forces (in this case, largely hydrogen bonds) are strong enough to affect the local structure right up to the boiling point. 5 Recall the analogous quantity used to describe the distance of an electron from an atom, Section 9.8.

Fig. 11.33 The experimentally determined radial distribution function of the oxygen atoms in liquid water at three temperatures. Note the expansion as the temperature is raised.

440

11 MACROMOLECULES AND SELF-ASSEMBLY

Fig. 11.34 A fragment of the crystal structure of ice. Each O atom is at the center of a tetrahedron of four O atoms at a distance of 276 pm. The central O atom is attached by two short O–H bonds to two H atoms and by two relatively long O···H bonds to two neighboring H2O molecules. Overall, the structure consists of planes of hexagonal puckered rings of H2O molecules (like the chair form of cyclohexane).

Fig. 11.35 A freely jointed chain is like a three-dimensional random walk, each step being in an arbitrary direction but of the same length.

11.12 Random coils The next stage for understanding the link between the structure and properties of a biological macromolecule is to consider the least organized structure of a chain of atoms, a dynamically active random coil.

Unlike the molecules of a liquid, the atoms and subunits of a macromolecule are tied together by chemical bonds. However, the atoms may still have considerable freedom of location on account of the ability of the units to rotate relative to their neighbors. A random coil is a disorganized conformation of a flexible macromolecule. The simplest model of a random coil is a freely jointed chain, in which any bond is free to make any angle with respect to the preceding one (Fig. 11.35). We assume that the residues occupy zero volume, so different parts of the chain can occupy the same region of space. The model is obviously an oversimplification because a bond is actually constrained to a cone of angles around a direction defined by its neighbor. In a hypothetical one-dimensional freely jointed chain all the residues lie in a straight line, and the angle between neighbors is either 0° or 180°. The residues in a three-dimensional freely jointed chain are not restricted to lie in a line or a plane. The probability, f(r)dr, that the distance between the ends of a threedimensional freely jointed chain of N units of length l lies in the range r to r + dr is6 f(r)dr = 4p

A a D 3 2 −a r A 3 D 1/2 r e dr, a = C p1/2F C 2Nl 2 F 2 2

Distribution of the separation of the ends of a three-dimensional chain

(11.27)

In some coils, the ends may be far apart, whereas in others their separation is small. Note that it is very unlikely that the two ends will be found either very close together (r = 0), because the factor r 2 vanishes, or stretched out in an almost straight line, because the exponential factor then vanishes. An alternative interpretation of f(r) is to regard each coil in a sample as ceaselessly writhing from one conformation to another; then f(r)dr is the probability that at any instant the chain will be found with the separation of its ends between r and r + dr. (a) Measures of size

There are several measures of the geometrical size of a three-dimensional random coil. The root mean square separation, Rrms, is a measure of the average separation of the ends of the coil: Rrms = N 1/2l

Root mean square separation

(11.28)

We see that as the number of residues N (each of length l) increases, the root mean square separation of its ends increases as N 1/2, and consequently the volume 6

See our Physical chemistry (2010) for full derivations of eqns 11.27, 11.28, and 11.31.

11.12 RANDOM COILS

441

of the coil increases as N 3/2. The contour length, Rc, is the length of the macromolecule measured along its backbone from atom to atom: Rc = Nl

Contour length

(11.29)

Another convenient measure of size is the radius of gyration of the macromolecule, the radius of a thin hollow spherical shell of the same mass and moment of inertia as the molecule (Fig. 11.36). For example, a solid sphere of radius R has R g = (35)1/2R and a long thin rod of length l has R g = l/121/2 for rotation about an axis perpendicular to the long axis. For a random coil, Rg =

A N D 1/2 l C 6F

Radius of gyration

(11.30)

and we see that, for specified values of N and l, R rms > R g (Fig. 11.37). Table 11.6 lists some experimental values of R g.

A spherical molecule of radius R and the smaller hollow spherical shell that has the same rotational characteristics. The radius of the hollow shell is the radius of gyration Rg of the molecule.

Fig. 11.36

A brief illustration

With a powerful microscope it is possible to see that a long piece of doublestranded DNA is flexible and writhes as if it were a random coil. However, small segments of the macromolecule resist bending, so it is more appropriate to visualize DNA as a freely jointed chain with N and l as the number and length, respectively, of these rigid units. The length l, the persistence length, is approximately 45 nm, corresponding to approximately 130 base pairs. It follows that for a piece of DNA with N = 200, we estimate (by using 103 nm = 1 mm) from eqn 11.29: R c = 200 × 45 nm = 9.0 mm from eqn 11.28: R rms = (200)1/2 × 45 nm = 0.64 mm from eqn 11.30: R g =

A 200D 1/2 × 45 nm = 0.26 mm C 6 F

A random coil in three dimensions. This one contains about 200 units. The root mean square distance between the ends (Rrms) and the radius of gyration (Rg) are indicated. Fig. 11.37

The random coil model ignores the role of the solvent: a ‘poor’ solvent will tend to cause the coil to tighten so that solute–solvent contacts are minimized; a ‘good’ solvent does the opposite. Therefore, calculations based on this model are better regarded as lower bounds to the dimensions for a coil in a good solvent and as an upper bound for a coil in a poor solvent. (b) Conformational entropy

Because a random coil is the least structured conformation of an idealized polymer chain, it corresponds to the state of greatest entropy. Any stretching of the coil introduces order and reduces the entropy. Conversely, the formation of a random coil from a more extended form is a spontaneous process (provided enthalpy contributions do not interfere). The change in conformational entropy, the entropy arising from the arrangement of bonds, when a coil containing N bonds of length l is stretched or compressed by nl is DS = − 12 kN ln {(1 + n)1+n(1 − n)1−n}

n = n/N

Conformational entropy

(11.31)

where k is Boltzmann’s constant and the maximum value of n is N, corresponding to maximum extension. This function is plotted in Fig. 11.38, and we see that minimum extension—fully coiled—corresponds to maximum entropy.

Radii of gyration of biological macromolecules and assemblies Table 11.6

M/(kg mol−1) Rg /nm DNA

4 × 103

Myosin

493

Serum albumin

66

Tobacco 3.9 × 104 mosaic virus

117.0 46.8 2.98 92.4

442

11 MACROMOLECULES AND SELF-ASSEMBLY

A brief illustration

Suppose that N = 1000 and l = 150 pm. The change in entropy when the (onedimensional) random coil is stretched through 1500 pm (corresponding to n = 1500 pm/150 pm = 10 and n = 1/100) is DS = −0.050k. The change in molar entropy is therefore DSm = −0.050R or −0.42 J K−1 mol−1 (we have used R = NAk).

11.13 Proteins We now need to understand how proteins attain complex structures.

Fig. 11.38 The change in molar entropy of a freely jointed chain as its extension changes; n = 1 corresponds to complete extension; n = 0, the conformation of highest entropy, corresponds to the random coil.

For a protein to function correctly, it needs to have a well-defined conformation. For example, an enzyme has its greatest catalytic efficiency only when it is in a specific conformation. In this section we explore the covalent and noncovalent interactions that cause polypeptides to fold into complex assemblies. (a) The secondary structure of a protein

The origin of the secondary structure of a protein is found in the rules formulated by Linus Pauling and Robert Corey in 1951. The essential feature is the stabilization of structures by hydrogen bonds involving the peptide link. The latter can act both as a donor of the H atom (the NH part of the link) and as an acceptor (the CO part). The Corey–Pauling rules are as follows (Fig. 11.39): 1. The four atoms of the peptide link lie in a relatively rigid plane. The planarity of the link is due to delocalization of p electrons over the O, C, and N atoms and the maintenance of maximum overlap of their p orbitals (see Exercise 10.41). 2. The N, H, and O atoms of a hydrogen bond lie in a straight line (with displacements of H tolerated up to not more than 30° from the N–O vector). 3. All NH and CO groups are engaged in hydrogen bonding.

The (a) angles and (b) bond lengths (pm) that characterize the peptide link. The C–NH–CO–C atoms define a plane (the C–N bond has partial double-bond character), but there is rotational freedom around the C–CO and N–C bonds.

Fig. 11.39

The rules are satisfied by two structures. One, in which hydrogen bonding between peptide links leads to a helical structure, is the a helix. The other, in which hydrogen bonding between peptide links leads to a planar structure, is the b sheet;7 this form is the secondary structure of the protein fibroin, the constituent of silk. The a-helix is illustrated in Fig. 11.40. Each turn of the helix contains 3.6 amino acid residues, so the period of the helix corresponds to five turns (18 residues). The pitch of a single turn (the distance between points separated by 360°) is 544 pm. The N–H···O bonds lie parallel to the axis and link every fourth group (so residue i is linked to residues i − 4 and i + 4). All the R groups point away from the major axis of the helix. There is freedom for the helix to be arranged as either a right- or a left-handed screw, but the overwhelming majority of natural polypeptides are right-handed on account of the preponderance of the l-configuration of the naturally occurring amino acids, as we explain below. The reason for their preponderance is not known. A polypeptide chain adopts a conformation corresponding to a minimum Gibbs energy, which depends on the conformational energy, the energy of 7

The sheet is often called the pleated sheet.

11.13 PROTEINS

interaction between different parts of the chain, and the energy of interaction between the chain and surrounding solvent molecules. In the aqueous environment of biological cells, the outer surface of a protein molecule is covered by a mobile sheath of water molecules, and its interior contains pockets of water molecules. These water molecules play an important role in determining the conformation that the chain adopts through hydrophobic interactions and hydrogen bonding to amino acids in the chain. The simplest calculations of the conformational energy of a polypeptide chain ignore entropy and solvent effects and concentrate on the total potential energy of all the interactions between nonbonded atoms. For example, these calculations predict that a right-handed a-helix of l-amino acids is marginally more stable than a left-handed helix of the same amino acids. To calculate the energy of a conformation, we need to make use of many of the molecular interactions described earlier in the chapter and also of some additional interactions: 1. Bond stretching. Bonds are not rigid, and it may be advantageous for some bonds to stretch and others to be compressed slightly as parts of the chain press against one another. If we liken the bond to a spring, then the potential energy takes the form corresponding to a Hooke’s law of force (restoring force proportional to the displacement; eqn 9.28) and is Contribution of bond stretching to the conformational energy

Vstretch = 12 k f,stretch(R − R e)2

(11.32)

where R e is the equilibrium bond length and k f,stretch is the stretching force constant, a measure of the stiffness of the bond in question. Self-test 11.13 The equilibrium bond length of a carbon–carbon single bond is 152 pm. Given a C–C force constant of 400 N m−1, how much energy, in kilojoules per mole, would it take to stretch the bond to 165 pm?

Answer: 3.38 × 10−20 J, equivalent to 20.3 kJ mol−1

2. Bond bending. An O–C–H bond angle (or some other angle) may open out or close in slightly to enable the molecule as a whole to fit together better. If the equilibrium bond angle is qe, we write Vbend = 12 k f,bend(q − qe)2

Contribution of bond bending to the conformational energy

(11.33)

where kf,bend is the bending force constant, a measure of how difficult it is to change the bond angle. Self-test 11.14 Theoretical studies have estimated that the lumiflavin isoalloazine ring system (13) has an energy minimum at the bending angle of 15°, but that it requires only 8.5 kJ mol−1 to increase the angle to 30°. If there are no other compensating interactions, what is the force constant for lumiflavin bending?

Answer: 1.26 × 10−22 J deg−2, equivalent to 75.6 J mol−1 deg−2

443

Fig. 11.40 The polypeptide a-helix, with poly-l-alanine as an example. There are 3.6 residues per turn and a translation along the helix of 150 pm per residue, giving a pitch of 544 pm. The diameter (ignoring side chains) is about 600 pm.

444

11 MACROMOLECULES AND SELF-ASSEMBLY

3. Bond torsion. There is a barrier to internal rotation of one bond relative to another (just like the barrier to internal rotation in ethane).

Fig. 11.41 The definition of the torsional angles y and f between two peptide units.

Because the planar peptide link is relatively rigid, the geometry of a polypeptide chain can be specified by the two angles that two neighboring planar peptide links make to each other. Figure 11.41 shows the two angles f and y commonly used to specify this relative orientation. The sign convention is that a positive angle means that the front atom must be rotated clockwise to bring it into an eclipsed position relative to the rear atom. For an all-trans form of the chain, all f and y are 180°. A helix is obtained when all the f are equal and when all the y are equal. For a right-handed a-helix, all f = 57° and all y = 47°. For a left-handed a-helix, both angles are positive. The torsional contribution to the total potential energy is Vtorsion = A(1 + cos 3f) + B(1 + cos 3y)

Contribution of bond torsion to the conformational energy

(11.34)

in which A and B are constants of the order of 1 kJ mol−1. Because only two angles are needed to specify the conformation of a helix, and they range from −180° to +180°, the torsional potential energy of the entire molecule can be represented on a Ramachandran plot, a contour diagram in which one axis represents f and the other represents y. 4. Interaction between partial charges. If the partial charges Qi and Qj on the atoms i and j are known, a Coulombic contribution of the form given in eqn 11.13 can be included, using the partial charges quoted in Table 11.2. The interaction between partial charges does away with the need to take dipole– dipole interactions into account, for they are taken care of by dealing with each partial charge explicitly. 5. Dispersive and repulsive interactions. The interaction energy of two atoms separated by a distance r (which we know once f and y are specified) can be given by the Lennard-Jones (12,6) form, eqn 11.25. 6. Hydrogen bonding. In some models of structure, the interaction between partial charges is judged to take into account the effect of hydrogen bonding. In other models, hydrogen bonding is added as another interaction of the form VH−bonding =

Contour plots of potential energy against the torsional angles y and f, also known as Ramachandran plots, for (a) a glycyl residue of a polypeptide chain and (b) an alanyl residue. The glycyl diagram is symmetrical, but that for alanyl is unsymmetrical, with the area shaded in red corresponding to an a-helix. (After T. Hovmoller et al., Acta Cryst. D58, 768 (2002).) Fig. 11.42

E F − r 12 r 10

Contribution of hydrogen bonding to the conformational energy

(11.35)

The total potential energy of a given conformation (f,y) can be calculated by summing the contributions given by eqns 11.32 through 11.35 and the contributions from Coulombic and dispersion interactions for all bond angles (including torsional angles) and pairs of atoms in the molecule. Figure 11.42 shows the potential energy contours for the helical form of polypeptide chains formed from the nonchiral amino acid glycine (R = H) and the chiral amino acid l-alanine (R = CH3). The contours were computed by summing all the contributions described above for each choice of angles and then plotting contours of equal potential energy. The glycine map is symmetrical, with minima corresponding to the formation of right- and left-handed helices. In contrast, the map for l-alanine is unsymmetrical, with the lowest being consistent with the formation of an a-helix. A b sheet is formed by hydrogen bonding between two extended polypeptide chains (large absolute values of the torsion angles f and y). Some of the R groups

11.13 PROTEINS

445

point above and some point below the sheet. Two types of structures can be distinguished from the pattern of hydrogen bonding between the constituent chains. In an antiparallel b sheet (Fig. 11.43a), f = −139°, y = +113°, and the N–H–O atoms of the hydrogen bonds form a straight line. This arrangement is a consequence of the antiparallel arrangement of the chains: every N–H bond on one chain is aligned with a C–O bond from another chain. Antiparallel b sheets are very common in proteins. In a parallel b sheet (Fig. 11.43b), f = −119° and y = +113°, and the N–H–O atoms of the hydrogen bonds are not perfectly aligned. This arrangement is a result of the parallel arrangement of the chains: each N–H bond on one chain is aligned with an N–H bond of another chain and, as a result, each C–O bond of one chain is aligned with a C–O bond of another chain. These structures are not common in proteins. Although we do not know all the rules that govern protein folding, X-ray diffraction studies of water-soluble natural proteins and synthetic polypeptides show that some amino acid residues appear in helical segments more frequently than in sheets, whereas others exhibit the opposite behavior. Table 11.7 summarizes the available data. (a) An antiparallel b sheet (f = −139°, y = +113°), in which the N–H–O atoms of the hydrogen bonds form a straight line. (b) A parallel b sheet (f = −119° and y = +113°), in which the N–H–O atoms of the hydrogen bonds are not perfectly aligned.

Fig. 11.43

(b) Higher-order structures of proteins

In an aqueous environment, chains fold in such a way as to place nonpolar R groups in the interior (which is often not very accessible to solvent) and charged R groups on the surface (in direct contact with the polar solvent). A wide variety of structures can result from these broad rules. Among them, a four-helix bundle (Fig. 11.44), which is found in proteins such as cytochrome b562 (an electrontransport protein, Atlas P5), forms when each helix has a nonpolar region along its length. The four nonpolar regions pack together to form a nonpolar interior. Similarly, interconnected b sheets may interact to form a b barrel (Fig. 11.45), the interior of which is populated by nonpolar R groups and which has an exterior rich in charged residues. The retinol-binding protein of blood plasma (Atlas P11), which is responsible for transporting vitamin A, is an example of a b barrel structure. Factors that promote the folding of proteins include covalent –S–S– disulfide links between cysteine residues (14), Coulombic interactions between ions (which depend on the degree of protonation of groups and therefore on the pH), hydrogen bonding (such as O–H···O), van der Waals interactions, and hydrophobic interactions. The clustering of nonpolar, hydrophobic amino acids into the interior of a protein is driven primarily by hydrophobic interactions (Section 2.7). Proteins with M > 50 kg mol−1 are often found to be aggregates of two or more polypeptide chains. Hemoglobin, which consists of four myoglobin-like chains (Fig. 11.46 and Atlas P7), is an example of a quaternary structure. Myoglobin (Atlas P10) is an oxygen-storage protein. The subtle differences that arise when four such molecules coalesce to form hemoglobin result in the latter being an oxygen transport protein, able to load O2 cooperatively and to unload it cooperatively too (see Case studies 4.1 and 10.4). Proteins can also self-assemble into rather large aggregates. Collagen (Atlas P4), the most abundant protein in mammals and responsible for imparting mechanical strength to tissues and organs, consists of three long helices wound around each other. The protein actin forms thin, rodlike filaments that, when associated with several copies of the protein myosin, play an important role in the mechanism of muscle contraction. The microtubules that participate in the

446

11 MACROMOLECULES AND SELF-ASSEMBLY

Relative frequencies of amino acid residues in helices and sheets

Table 11.7

Amino acid

a helix

b sheet

Alanine

1.29

0.90

Arginine

0.96

0.99

Asparagine

0.90

0.76

Aspartic acid

1.04

0.72

Cysteine

1.11

0.74

Glutamic acid

1.44

0.75

Glutamine

1.27

0.80

Glycine

0.56

0.92

Histidine

1.22

1.08

Isoleucine

0.97

1.45

Leucine

1.30

1.02

Lysine

1.23

0.77

Methionine

1.47

0.97

Phenylalanine

1.07

1.32

Proline

0.52

0.64

Serine

0.82

0.95

Threonine

0.82

1.21

Tryptophan

0.99

1.14

Tyrosine

0.72

1.25

Valine

0.91

1.49

Data from T.E. Creighton, Proteins: structures and molecular properties, W. H. Freeman and Co., New York (1992).

A four-helix bundle forms from the interactions between nonpolar amino acids on the surfaces of each helix, with the polar amino acids exposed to the aqueous environment of the solvent.

Fig. 11.44

Eight antiparallel b sheets, each represented by an arrow and linked by short random coils fold together as a barrel. Nonpolar amino acids are in the interior of the barrel.

Fig. 11.45

separation of chromosomes during cell division, provide structural rigidity in cells, and participate in the motile function of flagella are hollow cylinders formed by aggregation of the protein tubulin. Not all protein aggregates are beneficial. In patients afflicted with sickle-cell anemia, hemoglobin molecules aggregate into rods, rendering the red blood cell unable to transport O2 efficiently. Also, the presence of aggregates of proteins in the brain appears to be associated with several serious conditions. For example, the amyloid plaques found in postmortem analysis of the brains of patients with Alzheimer’s disease are a mixture of damaged neurons and aggregates of the b amyloid protein, which is an extended antiparallel b sheet. 11.14 Nucleic acids Of crucial biological importance are the conformations adopted by nucleic acids, the key components of the mechanism of storage and transfer of genetic information in biological cells.

We saw in Fundamentals F.1 that DNA and RNA are polynucleotides, polymers of base–sugar–phosphate units linked by phosphodiester bonds, that self-assemble

Fig. 11.46 A hemoglobin molecule consists of four myoglobin-like units. An O2 molecule attaches to the ion atom in the heme group indicated by the arrow.

11.14 NUCLEIC ACIDS

447

Fig. 11.47 The structural features of the most abundant forms of DNA in the cell: (a) B-DNA, (b) A-DNA.

into complex three-dimensional structures. An example of secondary structure in nucleic acids is the winding of two polynucleotide chains around each other to form a DNA double helix, as shown in Fig. 11.47. Figure 11.47 also shows that different forms of the double helix are possible. In B-DNA, the most abundant form of DNA in the cell (Fig. 11.47a), the rodlike double helix is right-handed with a diameter of 2.37 nm and a pitch of 3.54 nm. The base pairs are approximately parallel to each other and perpendicular to the long axis of the rod. In A-DNA (Fig. 11.47b), the double helix is right-handed but slightly wider, with a diameter of approximately 2.55 nm and a pitch of 2.53 nm. The base pairs are parallel to each other but not perpendicular to the long axis of the helix. Double-stranded RNA and hybrid RNA-DNA, the assembly of one strand of ribonucleic acid strand with a DNA strand, assume the A form. A third form of DNA, called Z-DNA, is a left-handed helix with a diameter of 1.84 nm, a pitch of 4.56 nm, and a slightly tilted arrangement of the base pairs relative to the long axis of the helix. The physiological role of Z-DNA is not certain. We saw in Section 3.5 that base pairing by hydrogen bonding is largely responsible for the thermal stability of DNA. A more subtle interaction that confers stability to DNA is base stacking, in which dispersion interactions bring together the planar systems of bases. Experiments show that stacking interactions are stronger between C–G base pairs than between A–T base pairs. It follows that two factors render DNA sequences rich in C–G base pairs more stable than sequences rich in A–T base pairs: more hydrogen bonds between the bases (Section 3.5) and stronger stacking interactions between base pairs. Some drugs with planar p systems, shown as a gray rectangle in the illustration, are effective because they intercalate between base pairs through stacking interactions, causing the helix to unwind slightly and altering the function of DNA (Fig. 11.48). Because a long stretch of DNA is flexible, it can undergo further folding into a variety of tertiary structures. Two examples are shown in Fig. 11.49. Supercoiled DNA is found in the chromosome and can be visualized as the twisting of closed circular DNA (ccDNA), much like the twisting of a rubber band. Before it can participate in the transmission of genetic information, supercoiled DNA must be uncoiled. Both coiling and uncoiling are catalyzed by enzymes belonging to the topoisomerase family. There are important differences in the chemical compositions of RNA and DNA that translate into different secondary and tertiary structures. In RNA the sugar is b-d-ribose (Atlas S1), whereas in DNA it is b-d-2-deoxyribose (Atlas S2). Although adenine, cytosine, and guanine are found in both DNA and RNA, in RNA uracil (Atlas B5) replaces thymine. As in DNA, the secondary and tertiary structures of RNA arise primarily from the pattern of hydrogen bonding between bases of one or more chains. The extra –OH group in b-d-ribose imparts enough steric strain to a polynucleotide chain that stable double helices cannot form in RNA. Therefore, RNA exists primarily as single chains that can fold into complex

Some drugs with planar p systems, shown as a rectangle, intercalate between base pairs of DNA.

Fig. 11.48

A long section of DNA may form closed circular DNA (ccDNA) by covalent linkage of the two ends of the chain. Twisting of ccDNA leads to the formation of supercoiled DNA.

Fig. 11.49

448

11 MACROMOLECULES AND SELF-ASSEMBLY

structures by formation of A–U and G–C base pairs. One example of this effect is the structure of transfer RNA (tRNA), shown schematically in Fig. 11.50, in which base-paired regions are connected by loops and coils. Transfer RNAs help assemble polypeptide chains during protein synthesis in the cell. 11.15 Polysaccharides To understand the connection between structure and biological function of carbohydrates, we need to examine the conformations adopted by their polymers.

The structure of a transfer RNA (tRNA).

Fig. 11.50

We saw in Fundamentals F.1 that polysaccharides are polymers of simple carbohydrates. Carbohydrate units are linked together in polysaccharides by glycosidic bonds that form between hydroxyl groups and result in C–O–C ether moieties. The orientation of one linked ring relative to another depends on which hydroxyl groups are linked and on their stereochemistry. Consider the a and b isomers of glucose (15a and 15b, respectively), which differ in the configuration of the C1 carbon. Linking the C1 and C4 carbons by glycosidic bonds, so-called 1,4-glycosidic bonds, results in either a bent (16) or a linear (17) chain, depending on whether the monomer is a- or b-glucose, respectively. Branched structures are also possible when a monomer makes three glycosidic bonds, as shown in (18).

11.16 MICELLES AND BIOLOGICAL MEMBRANES

449

Like polypeptides and polynucleotides, polysaccharides also possess different levels of structure. In cellulose, linear chains of glucose, such as those shown in (17), interact through hydrogen bonds involving hydroxyl groups and ring oxygen atoms. The resulting structure is a thin but strong fiber that is used to construct the wall of a plant cell. In amylase, which stores glucose molecules for future use by the plant cells, a bent chain, such as that in (16), coils into a helical structure held together by hydrogen bonds. Glycogen, which stores glucose in animals and microbes, and amylopectin, which—like amylase—performs the same function in plants, also feature a-1,4-linkages, but because of branching points (as in 18), these polymers do not adopt regular secondary structures. 11.16 Micelles and biological membranes We need to understand the factors that optimize the self-assembly of cell membranes.

We saw in Fundamentals F.1 that phospholipids are amphipathic molecules that can group together through hydrophobic interactions to form bilayer structures and cell membranes (Fig. F.1). Here we explore details of the self-assembly of amphipathic molecules into a variety of structures with significance to biology and medicine. (a) Micelles

In aqueous environments amphipathic molecules can group together as micelles, in which hydrophobic tails congregate, leaving hydrophilic heads exposed to the solvent (Fig. 11.51). Micelles are important in industry and biology on account of their solubilizing function: matter can be transported by water after it has been dissolved in their hydrocarbon interiors. Micelles form only above a certain concentration of amphiphiles called the critical micelle concentration (CMC) and above the Krafft temperature. Nonionic amphipathic molecules may cluster together in clumps of 1000 or more, but ionic species tend to be disrupted by the electrostatic repulsions between head groups and are normally limited to groups of fewer than about 100. The interior of a micelle is like a droplet of oil, and experiments show that the hydrophobic tails are mobile, but slightly more restricted than in the bulk. Different molecules tend to form micelles of different shapes. For example, ionic species such as sodium dodecyl sulfate (SDS) and cetyl trimethylammonium bromide (CTAB) form rods at moderate concentrations, whereas sugar molecules form small, approximately spherical micelles. Broadly speaking, the shapes of micelles vary with the shape of the constituent molecules, their concentration, and the temperature. A useful predictor of the shape of the micelle, the surfactant parameter, Ns, is defined as Ns =

V Al

Surfactant parameter

A representation of a spherical micelle. The hydrophilic groups are represented by the red spheres and the hydrophobic hydrocarbon chains are represented by the stalks. The latter are mobile.

Fig. 11.51

Table 11.8 Variation of micelle shape with the surfactant parameter

Value or range of the surfactant parameter, Ns

Micelle shape

< 0.33

Spherical

0.33–0.50

Cylindrical rods

0.50–1.00

Vesicles

1.00

Planar bilayers

> 1.00

Reverse micelles and other shapes

(11.36)

where V is the volume of the hydrophobic tail, A is the area of the hydrophilic head group, and l is the maximum length of the tail. Table 11.8 summarizes the dependence of micelle shape on the surfactant parameter. Under certain experimental conditions, a liposome may form, with an inward pointing inner surface of molecules surrounded by an outward pointing outer layer (Fig. 11.52). Liposomes may be used to carry nonpolar drug molecules in blood. Reverse micelles form in nonpolar solvents, with small polar head groups

Fig. 11.52 The cross-sectional structure of a spherical liposome.

450

11 MACROMOLECULES AND SELF-ASSEMBLY

in a micellar core and more voluminous hydrophobic tails extending into the organic bulk phase. These spherical aggregates can solubilize water in organic solvents by creating a pool of trapped water molecules in the micellar core. (b) Bilayers, vesicles, and membranes

Gibbs energies of transfer of amino acid residues in a helix from the interior of a membrane to water

Table 11.9

Amino acid

D transferG/ (kJ mol−1)

Some micelles at concentrations well above the CMC form extended parallel sheets two molecules thick, called planar bilayers. The individual molecules lie perpendicular to the sheets, with hydrophilic groups on the outside in aqueous solution and on the inside in nonpolar media. When segments of planar bilayers fold back on themselves, unilamellar vesicles may form where the spherical hydrophobic bilayer shell separates an inner aqueous compartment from the external aqueous environment. Bilayers show a close resemblance to biological membranes and are often a useful model on which to base investigations of biological structure. However, actual membranes are highly sophisticated structures, in which phospholipid molecules form layers instead of micelles because the hydrocarbon chains are too bulky to allow packing into nearly spherical clusters. The bilayer is a highly mobile structure. Not only are the hydrocarbon chains ceaselessly twisting and turning in the region between the polar groups, but the phospholipid and cholesterol molecules migrate over the surface. It is better to think of the membrane as a viscous fluid rather than a permanent structure, with a viscosity about 100 times that of water. In common with diffusional behavior in general (Section 8.5), the average distance a phospholipid molecule diffuses is proportional to the square root of the time; more precisely, for a molecule confined to a two-dimensional plane, the average distance traveled in a time t is equal to (4Dt)1/2, where D is the diffusion constant. Typically, a phospholipid molecule migrates through about 1 mm in about 1 min.

Phenylalanine

15.5

Methionine

14.3

(c) Interactions between proteins and biological membranes

Isoleucine

13.0

Leucine

11.8

Valine

10.9

Peripheral proteins are proteins attached to the bilayer. Integral proteins are proteins embedded in the mobile but viscous bilayer. Examples include complexes I–IV of oxidative phosphorylation (Section 5.10), ion channels, and ion pumps (Section 5.3). Integral proteins may span the depth of the bilayer and consist of tightly packed a helices or, in some cases, b sheets containing hydrophobic residues that sit comfortably within the hydrocarbon region of the bilayer. The hydrophobicity of a residue can be assessed by measuring the Gibbs energy of transfer of the corresponding amino acid from an aqueous solution to the interior of a membrane (Table 11.9). Amino acids with negative values of the Gibbs energy of transfer are likely to be found in the membrane-spanning regions of integral proteins. There are two views of the motion of integral proteins in the bilayer. In the fluid mosaic model shown in Fig. 11.53, the proteins are mobile, but their diffusion coefficients are much smaller than those of the lipids. In the lipid raft model, a number of lipid and cholesterol molecules form ordered structures, or ‘rafts,’ that envelop proteins and help carry them to specific parts of the cell. The mobility of the bilayer enables it to flow around a molecule close to the outer surface, to engulf it, and to incorporate it into the cell by the process of endocytosis. Alternatively, material from the cell interior wrapped in cell membrane may coalesce with the cell membrane itself, which then withdraws and ejects the material in the process of exocytosis. An important function of the proteins embedded in the bilayer, however, is to act as devices for transporting matter into and out of the cell in a more subtle manner, as discussed in Section 8.6.

Cysteine

8.4

Tryptophan

8.0

Alanine

6.7

Threonine

5.0

Glycine

4.2

Serine

2.5

Proline

−0.8

Tyrosine

−2.9

Histidine

−12.6

Glutamine

−17.2

Asparagine

−20.2

Glutamic acid

−34.4

Lysine

−37.0

Aspartic acid

−38.6

Arginine

−51.7

Data from D.M. Engelman, T.A. Steitz, and A. Goldman, Ann. Rev. Biophys. Biophys. Chem. 15, 330 (1986).

11.17 COMPUTER-AIDED SIMULATION

451

11.17 Computer-aided simulation To understand the various approaches to the prediction of structure, we need to see how to take into account a balance of interactions that give a biological macromolecule its native conformation or hold a drug and receptor together.

We saw in Chapter 10 that ideas derived from quantum mechanics can be used to predict the structures and the physical and chemical properties of molecules. Semi-empirical, ab initio, and density functional methods work very well for molecules of modest size but require too much computational power and time to be suitable for predicting the structures of macromolecules. The problem is particularly acute when the surrounding water plays an important role in governing structure. For this reason, biochemists often rely on other techniques to generate three-dimensional models of proteins, nucleic acids, lipid bilayers, and drug–receptor complexes. Computational methods based on the principles of classical physics lead to the visual representation of atomic motions in biopolymers, thereby opening a window onto the molecular factors that are responsible for such dynamic processes as protein folding and enzyme catalysis. Yet other strategies can give insight into the structural features of a drug that optimize its docking to a receptor site. (a) Molecular mechanics calculations

We saw in Section 11.13 that the conformational energy, VC, of a biopolymer can be calculated by adding the contributions from steric interactions (bond stretching, bending, and torsion and dispersive interactions), electrostatic interactions, and hydrogen bonding: VC = Vstretch + Vbend + Vtorsion + VCoulomb + VLJ + VH-bonding

Conformational energy

(11.37)

Fig. 11.53 In the fluid mosaic model of a biological cell membrane, integral proteins diffuse through the lipid bilayer. In the alternative lipid raft model, a number of lipid and cholesterol molecules envelop and transport the protein around the membrane.

In a molecular mechanics simulation, the locations of the atoms are changed until the conformation with the lowest value of VC is found. For a macromolecule, a plot of the conformational energy against bond distance or bond angle often shows several local minima and a global minimum, which is associated with the preferred conformation (Fig. 11.54). Commercially available molecular modeling software packages include schemes for modifying and searching for these minima systematically. Molecular mechanics calculations are fast and do not require a great deal of computing power. However, they are of limited utility because the structure corresponding to the global minimum is a snapshot of the molecule at T = 0. That is, only the potential energy is included in the calculation; contributions to the total energy from kinetic energy are excluded. Also, the method does not handle interactions with a solvent. (b) Molecular dynamics and Monte Carlo simulations

Biological macromolecules (like all except the smallest molecules) are flexible and move ceaselessly. Atomic fluctuations and side-chain motions have amplitudes of 1–500 pm and characteristic times ranging from 1 fs to 0.1 s. Rigid body motions, such as the motions of helices and subunits, have amplitudes of 0.1–1.0 pm and characteristic times of 1 ns to 1 s. Folding transitions and the formation of quaternary structure from large structures have amplitudes greater than 0.5 nm and occur over a time span of from 100 ns to several hours. In a molecular dynamics simulation, the molecule is set in motion by treating it as though it has been heated to a specified temperature and the possible

For large molecules, a plot of potential energy against the molecular geometry often shows several local minima and a global minimum.

Fig. 11.54

452

11 MACROMOLECULES AND SELF-ASSEMBLY

trajectories of all atoms under the influence of the intermolecular potentials are calculated. To appreciate what is involved, we consider the motion of an atom in one dimension. We show in the following Justification that after a time interval Dt, the position of an atom changes from xi−1 to a new value xi given by xi = xi−1 + vi−1Dt

(11. 38)

where vi−1 is the velocity of the atom when it was at xi−1, its location at the start of the interval. The velocity at xi is related to vi−1, the velocity at the start of the interval, by vi = vi−1 − m−1

dVC(x) dx

(11. 39)

Dt xi−1

where the derivative of the conformational energy VC(x) is evaluated at xi−1. The time interval Dt is approximately 1 fs (10−15 s), which is shorter than the average time for the fastest atomic motions in a macromolecule. The calculation of xi and vi is then repeated for tens of thousands of such steps.

Justification 11.5 The atomic trajectories according to molecular dynamics

Consider an atom of mass m moving along the x direction with an initial velocity v1 given by vi = Dx/Dt. If the initial and new positions of the atom are x1 and x2, then Dx = x2 − x1 and x2 = x1 + v1Dt. This expression generalizes to eqn 11.38 for the calculation of a position xi from a previous position xi−1 and velocity vi−1. The atom moves under the influence of a force arising from interactions with other atoms in the molecule. From Newton’s second law of motion, we write the force F1 at x1 as F1 = ma1, where the acceleration a1 at x1 is given by a1 = Dv/Dt. If the initial and new velocities are v1 and v2, then Dv = v2 − v1 and v2 = v1 + a1Dt = v1 +

F1 Dt m

Because F = −dV/dx, the force acting on the atom is related to the potential energy of interaction with other nearby atoms, the conformational energy VC(x), by F1 = −

dVC(x) dx

x1

where the derivative is evaluated at x1. It follows that v2 = v1 − m−1

dVC(x) Dt dx x 1

This expression generalizes to eqn 11.39 for the calculation of a velocity vi from a previous velocity vi−1.

Self-test 11.15 Consider a particle of mass m connected to a stationary wall with a spring of force constant k f . Write an expression for the velocity of this particle once it is set into motion in the x direction from an equilibrium position x0.

Answer: vi = vi−1 + (k f /m)(xi−1 − x0)Dt

11.17 COMPUTER-AIDED SIMULATION

Commercially available software packages use versions of eqns 11.38 and 11.39 to calculate the trajectories of a large number of atoms in three dimensions. The trajectories correspond to the conformations that the molecule can sample at the temperature selected for the simulation. At very low temperatures, the molecule cannot overcome some of the potential energy barrier given by eqn 11.37, atomic motion is restricted, and only a few conformations are possible. At high temperatures, more potential energy barriers can be overcome and more conformations are accessible. Computational methods also allow for the simulation of a solvent cage around the macromolecule. In the Monte Carlo method, the atoms of a macromolecule are moved through small but otherwise random distances, and the change in conformational energy, DVC, is calculated. If the conformational energy is not greater than before the change, then the conformation is accepted. However, if the conformational energy is greater than before the change, it is necessary to check if the new conformation is reasonable and can exist in equilibrium with structures of lower conformational energy at the temperature of the simulation. To make progress, we use the Boltzmann distribution (Fundamentals F.3) to write that at equilibrium, the ratio of populations of two states with energy separation DVC is e−DV /kT, where k is Boltzmann’s constant. Because we are testing the viability of a structure with a higher conformational energy than the previous structure in the calculation, DVC > 0 and the exponential factor varies between 0 and 1. In the Monte Carlo method, the exponential factor is compared with a random number between 0 and 1; if the factor is larger than the random number, the conformation is accepted; if the factor is not larger, the conformation is rejected. Molecular dynamics and Monte Carlo simulations are much faster than quantum chemical calculations and can handle with relative ease the effect of solvent on the structure of a biopolymer. However, neither method is likely to yield the native structure of a large biopolymer from its sequence because of the very large number of states that must be sampled during the calculation. Nevertheless, the methods can be used to predict the effect of a minor change in the sequence of a nucleic acid or protein of known structure. Because in such a case the chemical substitution is not expected to result in a large deviation from the native structure, the calculation needs to sample only a manageable (but still large) number of conformations. This approach allows for the systematic investigation of a very large number of biopolymers, potentially leading to the determination of the chemical rules for stabilization of biomolecular structure. In much the same vein, the combination of molecular dynamics and Monte Carlo simulations can be used to investigate the thermodynamics of interaction between a drug and a biopolymer. C

(c) Quantitative structure–activity relationships

Computational approaches are having a considerable impact on the processes of drug discovery. To devise efficient therapies, it is necessary to know how to characterize and optimize both the three-dimensional structure of the drug and the molecular interactions between the drug and its target. Computational studies of the types described in Chapter 10 and this chapter can identify regions of a molecule that have high or low electron densities and result in specific interactions between the host protein and the guest agent. The graphical representation of numerical results brings these interactions vividly to life and allows modifications to be analyzed in the hope of improving specificity. In structure-based design, new drugs are developed on the basis of the known structure of the receptor site of a known target. However, in many cases a number of so-called lead compounds are known to have some biological activity but little

453

454

11 MACROMOLECULES AND SELF-ASSEMBLY

information is available about the target. To design a molecule with improved pharmacological efficacy, quantitative structure–activity relationships (QSAR) are often established by correlating data on activity of lead compounds with molecular properties, also called molecular descriptors, which can be determined either experimentally or computationally. The first stage of the QSAR method consists of compiling molecular descriptors for a very large number of lead compounds. Descriptors such as molar mass, molecular dimensions and volume, and relative solubility in water and nonpolar solvents are available from routine experimental procedures. Quantum mechanical descriptors determined by calculations of the type described in Chapter 10 include bond orders and hom*o and LUMO energies. In the second stage of the process, biological activity is expressed as a function of the molecular descriptors. An example of a QSAR equation is: activity = c0 + c1d1 + c2d 12 + c3d2 + c4d 22 + · · ·

A QSAR correlation equation

(11.40)

where di is the value of the descriptor and ci is a coefficient calculated by fitting the data by regression analysis. The quadratic terms account for the fact that biological activity can have a maximum or minimum value at a specific descriptor value. For example, a molecule might not cross a biological membrane and become available for binding to targets in the interior of the cell if it is too hydrophilic, in which case it will not partition into the hydrophobic layer of the cell membrane, or too hydrophobic, for then it may bind too tightly to the membrane. It follows that the activity will peak at some intermediate value of a parameter that measures the relative solubility of the drug in water and organic solvents. In the final stage of the QSAR process, the activity of a drug candidate can be estimated from its molecular descriptors and the QSAR equation either by interpolation or extrapolation of the data. The predictions are more reliable when a large number of lead compounds and molecular descriptors are used to generate the QSAR equation. The traditional QSAR technique has been refined into 3D QSAR, in which sophisticated computational methods are used to gain further insight into the three-dimensional features of drug candidates that lead to tight binding to the receptor site of a target. The process begins by using a computer to superimpose three-dimensional structural models of lead compounds and looking for common features, such as similarities in shape, location of functional groups, and electrostatic potential plots. The key assumption of the method is that common structural features are indicative of molecular properties that enhance binding of the drug to the receptor. The collection of superimposed molecules is then placed inside a three-dimensional grid of points. An atomic probe, typically an sp3hybridized carbon atom, visits each grid point and two energies of interaction are calculated: Esteric, the steric energy reflecting interactions between the probe and electrons in uncharged regions of the drug, and Eelec, the electrostatic energy arising from interactions between the probe and a region of the molecule carrying a partial charge. The measured equilibrium constant for binding of the drug to the target, Kbind, is then assumed to be related to the interaction energies at each point r by the 3D QSAR equation log Kbind = c0 + ∑ {csteric(r)Esteric(r) + celec(r)Eelec(r)} r

A 3D-QSAR equation

(11.41)

where the c(r) are coefficients calculated by regression analysis, with the coefficients csteric and celec reflecting the relative importance of steric and electrostatic

CHECKLIST OF KEY CONCEPTS

455

A 3D QSAR analysis of the steroids to human corticosteroid-binding globulin (CBG). The tinted regions indicate areas in the protein’s binding site with positive and negative electrostatic potentials and with little or much steric crowding. [Adapted from P. Krogsgaard-Larsen, T. Liljefors, and U Marsden (cd.) A textbook of drug design and discovery, Taylor & Francis, London (2002).]

Fig. 11.55

interactions, respectively, at the grid point r. Visualization of the regression analysis is facilitated by coloring each grid point according to the magnitude of the coefficients. Figure 11.55 shows results of a 3D QSAR analysis of the binding of steroids, molecules with the carbon skeleton shown, to human corticosteroidbinding globulin (CBG). Indeed, we see that the technique lives up to the promise of opening a window into the chemical nature of the binding site even when its structure is not known.

Checklist of key concepts 1. In ultracentrifugation, a sample is exposed to a strong centrifugal field generated by rotation at high speeds and the molar mass of a biopolymer is calculated from the sedimentation constant. 2. MALDI-TOF mass spectrometry is a technique for the determination of molar masses in which a sample is ionized in the gas phase and the mass-to-charge number ratios of all ions are measured. 3. In laser light scattering the size and shape of a macromolecule are obtained from analysis of the intensity of light scattered by the sample. 4. X-ray crystallography is a collection of X-ray diffraction techniques based on applications of Bragg’s law to the determination of the three-dimensional structures of small and large molecules, including biopolymers. 5. Unit cells are classified into seven crystal systems according to their rotational symmetries. 6. Crystal planes are specified by a set of Miller indices (hkl). 7. In X-ray crystallography the electron density is calculated from the intensities Ih of scattered X-rays and the structure factors. 8. Crystals of proteins amenable to analysis by X-ray diffraction techniques can be made by adding a large amount of a salt, such as (NH4)2SO4, to a solution containing a charged protein. Detergents are often used to crystallize hydrophobic proteins. 9. A van der Waals interaction between closed-shell molecules is inversely proportional to the sixth power of their separation.

10. A polar molecule is a molecule with a permanent electric dipole moment; the magnitude of a dipole moment is the product of the partial charge and the separation. 11. The following interactions are important in biological self-assembly: charge–charge, charge–dipole, dipole–dipole, dipole–induced-dipole, dispersion (London), hydrogen bonding. 12. A hydrogen bond is an interaction of the form X–H···Y, where X and Y are N, O, or F. 13. The Lennard-Jones (12,6) potential is a model of the total intermolecular potential energy. 14. The relative locations of molecules in a liquid are reported in terms of the radial distribution function, g(r). 15. The least structured model of a macromolecule, such as a long stretch of DNA or a denatured protein, is as a random coil. 16. The conformational entropy of a random coil is the entropy arising from the arrangement of bonds. 17. The secondary structure of a polypeptide chain can be specified by the two angles, f and y, that two neighboring planar peptide links make to each other. 18. The different forms of double-helical DNA (B-, A-, and Z-) differ in diameter, pitch, and tilt of the base pairs relative to the long axis of the helix. 19. Supercoiled DNA is formed by twisting of closed circular DNA (ccDNA). 20. RNA exists primarily as single chains that can fold into complex structures by formation of base pairs.

456

11 MACROMOLECULES AND SELF-ASSEMBLY

21. Carbohydrate units are linked together in polysaccharides by glycosidic bonds between hydroxyl groups; bent, linear, or branched chains can result depending on which hydroxyl groups are linked.

25.

22. Micelles form in aqueous environments when the hydrophobic tails of amphipathic molecules congregate and their hydrophilic heads are exposed to the surrounding water molecules.

26.

23. In the fluid mosaic model of the cell membrane, integral proteins are mobile. In the lipid raft model, a number of lipid and cholesterol molecules form ordered structures, or ‘rafts’, that envelop proteins and help carry them to specific parts of the cell. 24. A biopolymer adopts a conformation corresponding to a minimum Gibbs energy, which depends on the conformational energy, the energy of interaction

27.

28.

between different parts of the polymer, and the energy of interaction between the polymer and surrounding solvent molecules. In a molecular mechanics simulation, the locations of the atoms are changed until the conformation with the lowest value of the total potential energy is found. In a molecular dynamics simulation, the molecule is set in motion by supposing that it has been heated to a specified temperature and the possible trajectories of all atoms under the influence of the intermolecular potentials are calculated. In a Monte Carlo simulation, the atoms of a macromolecule are moved through small but otherwise random distances, and the change in conformational energy is calculated. In the QSAR technique, pharmacological activity is correlated with a variety of molecular characteristics.

Checklist of key equations Property

Equation

Molar mass obtained from: sedimentation rate equilibrium sedimentation Mass-to-charge ratio in a TOF spectrometer Molar mass from the Rayleigh ratio Separation between crystal planes Bragg’s law

M = SRT/bD, S = s/rw 2 M = [2RT/{(r 22 − r 21)bw 2}]ln(c2/c1) m/z = 2eEd(t/l)2 R(q) = KP(q)cMM 1/d 2 = h2/a 2 + k 2/b 2 + l 2/c 2 l = 2d sin q

Fourier synthesis

r(x) = (1/V ) F0 + 2 ∑ Fh cos(2hpx)

Coulomb’s law in any medium Dipole moment in terms of electronegativity difference Charge–dipole interaction energy

V = Q1Q2/4pεr m/D ≈ Dc V = −Q2 m1/(4pε0r 2) V = −Q2 m1 cos q/(4pε0r 2) V = m1m2(1 − 3 cos2 q)/(4pε0r 3) V = −2m21 m22/{3(4pε0)2kTr 6} a′ = a/4pε0 V = −m21a2/(4pε0r 6) V = − 32 × (a1′a2′/r 6) × {I1I2/(I1 + I2)} V = 4ε{(s/r)12 − (s/r)6}

Dipole–dipole interaction energy Polarizability volume Dipole–induced-dipole interaction energy London formula Lennard-Jones (12,6) potential energy Measures of size: root mean square separation contour length radius of gyration Conformational entropy Conformational energy bond stretching bond bending bond torsion hydrogen bonding Surfactant parameter A QSAR equation

Comment

h=1

Orthogonal lattice

冧 See 7 See 8 See 9 Freely rotating dipoles

Random coil Rrms = N 1/2l Rc = Nl Rg = (N/6)1/2l DS = 12 kN ln{(1 + n)1+n(1 − n)1−n} VC = Vstretch + Vbend + Vtorsion + VCoulomb + VLJ + VH-bonding Vstretch = 12 k f,stretch(R − Re)2 Vbend = 12 k f,bend(q − qe )2 Vtorsion = A(1 + cos 3f) + B(1 + cos 3y) VH-bonding = E/r 12 − F/r 10 Ns = V/Al Activity = c0 + c1d1 + c2d 12 + c3d2 + c4d 22 + . . .

Random coil

Definition

457

EXERCISES

Discussion questions 11.1 What features in an X-ray diffraction pattern suggest a helical

conformation for a biological macromolecule?

or (c) hydrophobic interactions (Section 2.7)? Hint: Consult data from Tables 4.6 and 11.9.

11.2 Describe the phase problem in X-ray diffraction and explain

11.9 Why are DNA sequences rich in C–G base pairs more stable than

how it may be overcome.

sequences rich in A–T base pairs?

11.3 Explain how the permanent dipole moment and the

11.10 Discuss the factors that lead to bent, linear, and branched

polarizability of a molecule arise.

structures in polysaccharides.

11.4 Describe the formation of a hydrogen bond in terms of

11.11 (a) Distinguish between micelles, liposomes, bilayers, vesicles, and membranes. (b) Discuss the role of the surfactant parameter as a predictor of the shape of a micelle.

(a) an electrostatic interaction and (b) molecular orbital theory. 11.5 Distinguish between contour length, root-mean-square

separation, and radius of gyration of a random coil. 11.6 Identify the terms in and limit the generality of the following

expressions: (a) V = −Q2 m1/4pε0r 2, (b) V = Q2 m1 cos q/4pε0r 2, (c) V = m2 m1(1 − 3 cos2 q)/4pε0r 3, (d) Rrms = N 1/2l, and (e) R g = (N/6)1/2l. 11.7 Distinguish between an a helix, an anti-parallel b sheet, and a

parallel b sheet. 11.8 Which amino acids have side chains that can interact with

molecules (such as other amino acids or enzyme substrates) at pH 7 through (a) Coulombic interactions, (b) hydrogen bonding,

11.12 It is observed that the critical micelle concentration of sodium dodecyl sulfate in aqueous solution decreases as the concentration of added sodium chloride increases. Explain this effect. 11.13 Distinguish between the fluid mosaic and lipid raft models for motion of integral proteins in a biological membrane. 11.14 Distinguish between molecular mechanics, molecular dynamics, and Monte Carlo calculations. Why are these methods generally more popular in biochemical research than the quantum mechanical procedures discussed in Chapter 10?

Exercises 11.15 The data from a sedimentation equilibrium experiment performed at 300 K on a macromolecular solute in aqueous solution show that a graph of ln c against r 2 is a straight line with slope 729 cm−2. The rotational rate of the centrifuge was 50 000 r.p.m. The specific volume of the solute is vs = 0.61 cm3 g−1. Calculate the molar mass of the solute. Hint: Use eqn 11.3 and take r =1.00 g cm−3. 11.16 Find the drift speed of a particle of radius 20 m and density

t/ms nbp

39.03 9

66.43 34

96.28 76

121.25 123

t/ms nbp

189.67 307

217.23 404

247.81 527

269.05 622

154.01 201

(a) Plot nbp against t and then against t2. Which plot is linear? Explain the physical origin of the linear relationship. (b) What time of flight would be observed for a fragment with 238 base pairs?

1750 kg m−3 that is settling from suspension in water (density = 1000 kg m−3) under the influence of gravity alone. The viscosity of water is 8.9 × 10 4 kg m−1 s−1.

11.20 Draw a set of points as a rectangular array based on unit cells of side a and b, and mark the planes with Miller indices (10), (01), (11), (12), (23), (41), and (41).

11.17 At 20°C the diffusion coefficient of a macromolecule is found to be 8.3 × 10−11 m2 s−1. Its sedimentation constant is 3.2 Sv in a solution of density 1.06 g cm−3. The specific volume of the macromolecule is 0.656 cm3 g−1. Determine the molar mass of the macromolecule.

11.21 Repeat Exercise 11.20 for an array of points in which the a and b axes make 60° to each other.

11.18 Calculate the speed of operation (in r.p.m.) of an ultracentrifuge needed to obtain a readily measurable concentration gradient in a sedimentation equilibrium experiment. Take that gradient to be a concentration at the bottom of the cell about five times greater than that at the top. Use rtop = 5.0 cm, rbottom = 7.0 cm, M ≈ 105 g mol−1, rvs ≈ 0.75, T = 298 K. 11.19 Mass spectrometry can be used for sizing DNA molecules. To appreciate the power of the technique, consider the analysis by MALDI-TOF of a mixture of fragments of pBR 322 DNA. It was observed that the time of flight, t, varied with nbp, the number of base pairs, as follows:

11.22 In a certain unit cell, planes cut through the crystal axes at (2a,3b,c), (a,b,c), (6a,3b,3c), and (2a,−3b,−3c). Identify the Miller indices of the planes. 11.23 Draw an orthorhombic unit cell and mark on it the (100), (010), (001), (011), (101), and (101) planes. 11.24 (a) Calculate the separations of the planes (111), (211), and (100) in a crystal in which the cubic unit cell has sides of length 532 pm. (b) Calculate the separations of the planes (123) and (236) in an orthorhombic crystal in which the unit cell has sides of lengths 0.754, 0.623, and 0.433 nm. 11.25 The glancing angle of a Bragg reflection from a set of crystal planes separated by 97.3 pm is 19.85°. Calculate the wavelength of the X-rays.

458

11 MACROMOLECULES AND SELF-ASSEMBLY

11.26 Construct the electron density along the x-axis of a crystal given the following structure factors:

11.34 (a) What are the units of the polarizability a? (b) Show that the units of polarizability volume are cubic meters (m3).

h Fh

0 +30.0

1 +8.2

2 +6.5

3 +4.1

4 +5.5

h Fh

5 −2.4

6 +5.4

7 +3.2

8 −1.0

9 +1.1

11.35 The magnitude of the electric field at a distance r from a point charge Q is equal to Q/4pε0r 2. How close to a water molecule (of polarizability volume 1.48 × 10−30 m3) must a proton approach before the dipole moment it induces is equal to the permanent dipole moment of the molecule (1.85 D)?

h Fh

10 +6.5

11 +5.2

12 −4.3

13 −1.2

14 +0.1

15 +2.1

11.27 Consider the electrostatic model of the hydrogen bond. The N–C distance of the hydrogen bonded groups in proteins, such as occur in an a helix, is 0.29 nm. How much energy (in kJ mol−1) is required to break the hydrogen bond (a) in a vacuum (εr = 1), (b) in a membrane (essentially a liquid hydrocarbon with εr = 2.0), and (c) in water (εr = 80.0)?

11.36 Phenylanine (21 and Atlas A14) is a naturally occurring amino acid with a benzene ring. What is the energy of interaction between its benzene ring and the electric dipole moment of a neighboring peptide group? Take the distance between the groups as 4.0 nm and treat the benzene ring as benzene itself and the phenyl group as benzene molecules. The dipole moment of the peptide group is m = 2.7 D and the polarizability volume of benzene is a′ = 1.04 × 10−29 m3.

11.28 Estimate the dipole moment of an HCl molecule from the electronegativities of the elements and express the answer in debye and coulomb meters (C m). 11.29 The technique of vector addition can be used to predict the dipole moment of a molecule. The resultant mres of two dipole moments m1 and m2 that make an angle q to each other is approximately

mres ≈ (m12 + m22 + 2m1m2 cos q)1/2 (a) Calculate the resultant of two dipoles of magnitude 1.50 D and 0.80 D that make an angle 109.5° to each other. (b) Estimate the ratio of the electric dipole moments of ortho (1,2-) and meta (1,3-) disubstituted benzenes. 11.30 Calculate the electric dipole moment of a glycine molecule using the partial charges in Table 11.2 and the locations of the atoms shown in (19).

11.37 Now consider the London interaction between the benzene rings of two Phe residues (see Exercise 11.36). Estimate the potential energy of attraction between two such rings (treated as benzene molecules) separated by 4.0 nm. For the ionization energy, use I = 5.0 eV. 11.38 In a region of the oxygen-storage protein myoglobin, the OH group of a tyrosine residue is hydrogen bonded to the N atom of a histidine residue in the geometry shown in (22). Use the partial charges in Table 11.2 to estimate the potential energy of this interaction.

11.31 (a) Plot the magnitude of the electric dipole moment of hydrogen peroxide as the H–O–O–H (azimuthal) angle f changes. Use the dimensions shown in (20). (b) Devise a way for depicting how the angle as well as the magnitude changes.

11.32 Calculate the molar energy required to reverse the direction of a water molecule located (a) 100 pm and (b) 300 pm from a Li+ ion initially with the O atom closest to the ion. Take the dipole moment of water as 1.85 D. 11.33 Show, by following the procedure in Justification 11.4, that

eqn 11.17 describes the potential energy of two electric dipole moments in the orientation shown in structure (9) of the text.

11.39 Given that force is the negative slope of the potential energy, calculate the distance dependence of the force acting between two nonbonded groups of atoms in a polypeptide chain that have a London dispersion interaction with each other. What is the separation at which the force is zero? Hint: Calculate the slope by considering the potential energy at R and R + dR, with dR 650 nm. Drugs based on hematoporphyrin do not meet this criterion very well, so novel porphyrin and related macrocycles with more desirable electronic properties are being synthesized and tested. At the same time, the quantum yield of triplet formation and of 1O2 formation must be high so many drug molecules can be activated and many oxidation reactions can occur during a short period of laser irradiation. Photodynamic therapy has been used successfully in the treatment of macular degeneration, a disease of the retina that leads to blindness, and in a number of cancers, including those of the lung, bladder, skin, and esophagus.

CHECKLIST OF KEY CONCEPTS

507

Checklist of key concepts 1. Spectroscopy is the analysis of the electromagnetic radiation emitted, absorbed, or scattered by atoms and molecules. 2. A spectrometer consists of a source of radiation, a dispersing element (or an interferometer), and a detector. 3. In a Raman spectrum lines shifted to lower frequency than the incident radiation are called Stokes lines and lines shifted to higher frequency are called anti-Stokes lines. 4. The intensity of a transition is proportional to the square of the transition dipole moment. 5. A selection rule is a statement about when the transition dipole can be nonzero. 6. A gross selection rule specifies the general features a molecule must have if it is to have a spectrum of a given kind. 7. A specific selection rule is a statement about which changes in quantum number may occur in a transition. 8. The gross selection rule for infrared absorption spectra is that the electric dipole moment of the molecule must change during the vibration. 9. The specific selection rule for vibrational transitions is Dv = ±1. 10. The gross selection rule for the vibrational Raman spectrum of a polyatomic molecule is that the normal mode of vibration is accompanied by a changing polarizability. 11. The exclusion rule states that if the molecule has a center of inversion, then no modes can be both infrared and Raman active. 12. In resonance Raman spectroscopy, radiation that nearly coincides with the frequency of an electronic transition is used to excite the sample and the result is a much greater intensity in the scattered radiation. 13. In conventional microscopy, the diffraction limit prevents the study of specimens that are much smaller than the wavelength of light used as a probe.

14. In vibrational microscopy, an infrared or Raman spectrometer is combined with a microscope to yield the vibrational spectrum of molecules in small specimens, such as single cells. 15. The Franck–Condon principle states that because nuclei are so much more massive than electrons, an electronic transition takes place faster than the nuclei can respond. 16. A chromophore is a group with characteristic optical absorption: chromophores include d-metal complexes, the carbonyl group, and the carbon– carbon double bond. 17. Chiral molecules may show optical activity and circular dichroism, the differential absorption of left- and right-circularly polarized light. 18. In fluorescence, the spontaneously emitted radiation ceases quickly after the exciting radiation is extinguished. 19. In phosphorescence, the spontaneous emission may persist for long periods; the process involves intersystem crossing into a triplet state. 20. In fluorescence microscopy, images of biological cells at work are obtained by attaching a large number of fluorescent molecules to proteins, nucleic acids, and membranes, and then measuring the distribution of fluorescence intensity within the illuminated area. Special techniques permit the observation of fluorescence from single molecules in cells. 21. The primary quantum yield of a photochemical reaction is the number of events producing specified primary products for each photon absorbed; the overall quantum yield is the number of reactant molecules that react for each photon absorbed. 22. Collisional deactivation, electron transfer, and resonance energy transfer are common fluorescence quenching processes. The rate constants of electron and resonance energy transfer decrease with increasing separation between donor and acceptor molecules. 23. Fluorescence resonance energy transfer (FRET) forms the basis of a technique for measuring distances between molecules in biological systems.

508

12 OPTICAL SPECTROSCOPY AND PHOTOBIOLOGY

Checklist of key equations Property

Equation

Comment

Beer–Lambert law

I = I0e−ε[J]L

Uniform solution

Absorbance

A = ε[J]L

Definition

Transmittance

T = I/I0

Definition

Transition dipole moment

mfi =

Definition

Lifetime broadening

dE ≈ ħ/t

In practice: d6 ≈ (5.3 cm−1)/(t/ps)

Vibrational selection rule

Dv = ±1

Harmonic oscillator model

Number of vibrational modes

(a) 3N − 6, (b) 3N − 5

(a) Nonlinear molecules, (b) linear molecules

Primary quantum yield

f = rate/Iabs

Observed fluorescence lifetime

t0 = fF /kF

Stern–Volmer equation

fF,0 /fF = 1 + t0 kQ[Q]

y*m f myi dt

Energy transfer efficiency

hT = 1 − fF/fF,0

Förster theory

hT = R06/(R 06 + R6)

Definition

Discussion questions 12.1 Describe the physical origins of linewidths in absorption and

emission spectra. 12.2 (a) Discuss the physical origins of the gross selection rules for

infrared spectroscopy and Raman spectroscopy. (b) Suppose that you wish to characterize the normal modes of benzene in the gas phase. Why is it important to obtain both infrared absorption and Raman spectra of your sample?

12.5 Provide examples of common chromophores. 12.6 Describe the mechanisms of photon emission by fluorescence

and phosphorescence. 12.7 (a) Summarize the main features of the Förster theory of

resonance energy transfer. (b) Discuss FRET and photosynthetic light harvesting in terms of Förster theory.

12.3 Explain how color can arise from molecules. 12.4 Explain the origin of the Franck–Condon principle and how

it leads to the appearance of vibrational structure in an electronic transition.

Exercises 12.8 Express a wavelength of 670 nm as (a) a frequency and

(b) a wavenumber. 12.9 What is (a) the wavenumber and (b) the wavelength of

the radiation used by an FM radio transmitter broadcasting at 92.0 MHz? 12.10 When light of wavelength 410 nm passes through 2.5 mm of a solution of the dye responsible for the yellow of daffodils at a concentration 0.433 mmol dm−3, the transmission is 71.5 per cent.

Calculate the molar absorption coefficient of the coloring matter at this wavelength and express the answer in centimeters squared per mole (cm2 mol−1). 12.11 An aqueous solution of a triphosphate derivative of molar mass 602 g mol−1 was prepared by dissolving 30.2 mg in 500 cm3 of water and a sample was transferred to a cell of length 1.00 cm. The absorbance was measured as 1.011. (a) Calculate the molar absorption coefficient. (b) Calculate the transmittance, expressed as a percentage, for a solution of twice the concentration.

EXERCISES

12.12 A swimmer enters a gloomier world (in one sense) on diving to greater depths. Given that the mean molar absorption coefficient of seawater in the visible region is 6.2 × 10−5 dm3 mol−1 cm−1, calculate the depth at which a diver will experience (a) half the surface intensity of light and (b) one-tenth that intensity. 12.13 Consider a solution of two unrelated substances A and B. Let their molar absorption coefficients be equal at a certain wavelength and write their total absorbance A. Show that we can infer the concentration of A and B from the total absorbance at some other wavelength provided we know the molar absorption coefficients at that different wavelength. (See eqn 12.7.) 12.14 The molar absorption coefficients of tryptophan and tyrosine at

240 nm are 2.00 × 103 dm3 mol−1 cm−1 and 1.12 × 104 dm3 mol−1 cm−1, respectively, and at 280 nm they are 5.40 × 103 dm3 mol−1 cm−1 and 1.50 × 103 dm3 mol−1 cm−1. The absorbance of a sample obtained by hydrolysis of a protein was measured in a cell of thickness 1.00 cm and was found to be 0.660 at 240 nm and 0.221 at 280 nm. What are the concentrations of the two amino acids? 12.15 A solution was prepared by dissolving tryptophan and tyrosine in 0.15 m NaOH(aq) and a sample was transferred to a cell of length 1.00 cm. The two amino acids share the same molar absorption coefficient at 294 nm (2.38 × 103 dm3 mol−1 cm−1), and the absorbance of the solution at that wavelength is 0.468. At 280 nm the molar absorption coefficients are 5.23 × 103 and 1.58 × 103 dm3 mol−1 cm−1, respectively and the total absorbance of the solution is 0.676. What are the concentrations of the two amino acids? Hint: It would be sensible to use the result derived in Exercise 12.13, but this specific example could be worked through without using that general case. 12.16 In many cases it is possible to assume that an absorption band has a Gaussian line shape (one proportional to e−x ) centered on the band maximum. (a) Assume such a line shape and show that 2

A = ε(6)d6 ≈ 1.0645εmax D61/2 where D61/2 is the width at half-height. (b) The electronic absorption bands of many molecules in solution have half-widths at half-height of about 5000 cm−1. Estimate the integrated absorption coefficients of bands for which (i) εmax ≈ 1 × 104 dm3 mol−1 cm−1 and (ii) εmax ≈ 5 × 102 dm3 mol−1 cm−1. 12.17 *Ozone absorbs ultraviolet radiation in a part of the electromagnetic spectrum energetic enough to disrupt DNA in biological organisms and absorbed by no other abundant atmospheric constituent. This spectral range, denoted UVB, spans wavelengths from about 290 nm to 320 nm. (a) The abundance of ozone is typically inferred from measurements of UV absorption and is often expressed in terms of Dobson units (DU): 1 DU is equivalent to a layer of pure ozone 10 mm thick at 1 atm and 0°C. Compute the absorbance of UV radiation at 300 nm expected for an ozone abundance of 300 DU (a typical value) and 100 DU (a value reached during seasonal Antarctic ozone depletions) given a molar absorption coefficient of 476 dm3 mol−1 cm−1. (b) The molar extinction coefficient of ozone over the UVB range is given in the table below. Compute the integrated absorption coefficient of ozone over the wavelength range 290–320 nm. Hint: ε(6) can be fitted to an exponential function quite well.

* Adapted from a problem supplied by Charles Trapp and Carmen Giunta.

l/nm ε/(dm3 mol−1 cm−1) l/nm ε/(dm3 mol−1 cm−1)

292.0 1512 310.1 135.9

296.3 865 315.0 69.5

300.8 477 320.0 34.5

509

305.4 257

12.18 The Beer–Lambert law is derived on the basis that the concentration of absorbing species is uniform (see Justification 12.1). Suppose instead that the concentration falls exponentially as [J] = [J]0e−x/l. Derive an expression for the variation of I with sample length: suppose that l >> l. Hint: Work through Justification 12.1, but use this expression for the concentration. 12.19 Assume that the electronic states of the p electrons of a conjugated molecule can be approximated by the wavefunctions of a particle in a one-dimensional box and that the dipole moment can be related to the displacement along this length by m = −ex. Show that the transition probability for the transition n = 1 → n = 2 is nonzero, whereas that for n = 1 → n = 3 is zero. Hint: The following relations will be useful:

sin x sin y = 12 cos(x − y) − 12 cos(x + y)

x cos ax dx =

1 x cos ax + sin ax a2 a

12.20 Estimate the lifetime of a state that gives rise to a line of width (a) 0.1 cm−1, (b) 1 cm−1, and (c) 1.0 GHz. 12.21 A molecule in a liquid undergoes about 1 × 1013 collisions in each second. Suppose that (a) every collision is effective in deactivating the molecule vibrationally and (b) that one collision in 200 is effective. Calculate the width (in cm−1) of vibrational transitions in the molecule. 12.22 Suppose that the C=O group in a peptide bond can be regarded as isolated from the rest of the molecule. Given that the force constant of the bond in a carbonyl group is 908 N m−1, calculate the vibrational frequency of (a) 12C=16O and (b) 13C=16O. 12.23 The hydrogen halides have the following fundamental vibrational wavenumbers:

6/cm−1

HF 4141.3

HCl 2988.9

HBr 2649.7

HI 2309.5

(a) Calculate the force constants of the hydrogen–halogen bonds. (b) From the data in part (a), predict the fundamental vibrational wavenumbers of the deuterium halides. 12.24 Which of the following molecules may show infrared absorption spectra: (a) H2, (b) HCl, (c) CO2, (d) H2O, (e) CH3CH3, (f ) CH4, (g) CH3Cl, and (h) N2? 12.25 How many normal modes of vibration are there for (a) NO2, (b) N2O, (c) cyclohexane, and (d) hexane? 12.26 Consider the vibrational mode that corresponds to the uniform expansion of the benzene ring. Is it (a) Raman or (b) infrared active? 12.27 Suppose that three conformations are proposed for the nonlinear molecule H2O2 (8, 9, and 10). The infrared absorption spectrum of gaseous H2O2 has bands at 870, 1370, 2869, and

510

12 OPTICAL SPECTROSCOPY AND PHOTOBIOLOGY

3417 cm−1. The Raman spectrum of the same sample has bands at 877, 1408, 1435, and 3407 cm−1. All bands correspond to fundamental vibrational wavenumbers, and you may assume that (i) the 870 and 877 cm−1 bands arise from the same normal mode and (ii) the 3417 and 3407 cm−1 bands arise from the same normal mode. (a) If H2O2 were linear, how many normal modes of vibration would it have? (b) Determine which of the proposed conformations is inconsistent with the spectroscopic data. Explain your reasoning. 12.28 The compound CH3CH=CHCHO has a strong absorption in the ultraviolet at 46 950 cm−1 and a weak absorption at 30 000 cm−1. Justify these features in terms of the structure of the compound. 12.29 Figure 12.54 shows the UV–visible absorption spectra of a selection of amino acids. Suggest reasons for their different appearances in terms of the structures of the molecules.

12.32 Consider some of the precautions that must be taken when conducting single-molecule spectroscopy experiments. (a) What is the molar concentration of a solution in which there is, on average, one solute molecule in 1.0 mm3 (1.0 fL) of solution? (b) It is important to use pure solvents in single-molecule spectroscopy because optical signals from fluorescent impurities in the solvent may mask optical signals from the solute. Suppose that water containing a fluorescent impurity of molar mass 100 g mol−1 is used as solvent and that analysis indicates the presence of 0.10 mg of impurity per 1.0 kg of solvent. On average, how many impurity molecules will be present in 1.0 mm3 of solution? You may take the density of water as 1.0 g cm−3. Comment on the suitability of this solvent for single-molecule spectroscopy experiments. 12.33 Light-induced degradation of molecules, also called photobleaching, is a serious problem in single-molecule spectroscopy. A molecule of a fluorescent dye commonly used to label biopolymers can withstand about 106 excitations by photons before light-induced reactions destroy its p system and the molecule no longer fluoresces. For how long will a single dye molecule fluoresce while being excited by 1.0 mW of 488 nm radiation from a laser? You may assume that the dye has an absorption spectrum that peaks at 488 nm and that every photon delivered by the laser is absorbed by the molecule. 12.34 Consider a unimolecular photochemical reaction with rate constant kr = 1.7 × 104 s−1 that involves a reactant with an observed fluorescence lifetime of 1.0 ns and an observed phosphorescence lifetime of 1.0 ms. Is the excited singlet state or the excited triplet state the most likely precursor of the photochemical reaction? 12.35 In a photochemical reaction A → 2 B + C, the quantum yield

Fig. 12.54

12.30 Suppose that you are a color chemist and have been asked to intensify the color of a dye without changing the type of compound and that the dye in question is a polyene. (a) Would you choose to lengthen or to shorten the chain? (b) Would the modification to the length shift the apparent color of the dye toward the red or the blue? 12.31 Dansyl chloride (11), which absorbs maximally at 330 nm and fluoresces maximally at 510 nm, can be used to label amino acids in fluorescence microscopy and FRET studies. Tabulated below is the variation of the fluorescence intensity of an aqueous solution of dansyl chloride with time after excitation by a short laser pulse (with I0 the initial fluorescence intensity):

t/ns If /I0

5.0 0.45

10.0 0.21

15.0 0.11

20.0 0.05

with 500 nm light is 2.1 × 102 mol einstein−1 (1 einstein = 1 mol photons). After exposure of 300 mmol of A to the light, 2.28 mmol of B is formed. How many photons were absorbed by A?

12.36 In an experiment to measure the quantum yield of a photochemical reaction, the absorbing substance was exposed to 490 nm light from a 100 W source for 45 min. The intensity of the transmitted light was 40 per cent of the intensity of the incident light. As a result of irradiation, 0.344 mol of the absorbing substance decomposed. Determine the quantum yield. 12.37 When benzophenone is illuminated with ultraviolet radiation, it is excited into a singlet state. This singlet changes rapidly into a triplet, which phosphoresces. Triethylamine acts as a quencher for the triplet. In an experiment in methanol as solvent, the phosphorescence intensity Iphos varied with amine concentration as shown below. A time-resolved laser spectroscopy experiment had also shown that the half-life of the fluorescence in the absence of quencher is 29 ms. What is the value of kQ?

[Q]/(mol dm−3) Iphos/(arbitrary units)

0.0010 0.41

0.0050 0.25

0.0100 0.16

12.38 The fluorescence intensity If of a solution of a plant pigment illuminated by 330 nm radiation was studied in the presence of a quenching agent, with the following results

[Q]/(mmol dm−3) If /Iabs

(a) Calculate the observed fluorescence lifetime of dansyl chloride in water. (b) The fluorescence quantum yield of dansyl chloride in water is 0.70. What is the fluorescence rate constant?

1.0 0.31

2.0 0.18

3.0 0.13

4.0 0.10

5.0 0.081

In a second series of experiments, the fluorescence lifetimes of the pigment were determined by time-resolved spectroscopy: [Q]/(mmol dm−3) t/ns

1.0 76

2.0 45

3.0 32

4.0 25

5.0 20

PROJECTS

Determine the quenching rate constant and the half-life of the fluorescence. 12.39 The Förster theory of resonance energy transfer and the basis for the FRET technique can be tested by performing fluorescence measurements on a series of compounds in which an energy donor and an energy acceptor are covalently linked by a rigid molecular linker of variable and known length. L. Stryer and R.P. Haugland, Proc. Natl. Acad. Sci. USA 58, 719 (1967), collected the following data on a family of compounds with the general composition dansyl-(l-prolyl)n-naphthyl, in which the distance R between the naphthyl donor and the dansyl acceptor was varied by increasing the number of prolyl units in the linker:

R/nm 1.2 1.5 1.8 2.8 3.1 3.4 3.7 4.0 4.3 4.6 hT 0.99 0.94 0.97 0.82 0.74 0.65 0.40 0.28 0.24 0.16 Are the data described adequately by the Förster theory (eqns 12.26 and 12.27)? If so, what is the value of R0 for the naphthyl–dansyl pair? 12.40 An amino acid on the surface of a protein was labeled covalently with 1.5-I-AEDANS and another was labeled covalently with FITC. The fluorescence quantum yield of 1.5-I-AEDANS decreased by 10 per cent due to quenching by FITC. What is the distance between the amino acids? Hint: see Table 21.6. 12.41 The flux of visible photons reaching Earth from the North Star is about 4 × 103 mm−2 s−1. Of these photons, 30 per cent are absorbed or scattered by the atmosphere and 25 per cent of the surviving photons are scattered by the surface of the cornea of the eye. A further 9 per cent are absorbed inside the cornea. The area of the pupil at

511

night is about 40 mm2 and the response time of the eye is about 0.1 s. Of the photons passing through the pupil, about 43 per cent are absorbed in the ocular medium. How many photons from the North Star are focused onto the retina in 0.1 s? For a continuation of this story, see R.W. Rodieck, The first steps in seeing, Sinauer, Sunderland (1998). 12.42 In light-harvesting complexes, the fluorescence of a chlorophyll molecule is quenched by nearby chlorophyll molecules. Given that for a pair of chlorophyll a molecules R0 = 5.6 nm, by what distance should two chlorophyll a molecules be separated to shorten the fluorescence lifetime from 1 ns (a typical value for monomeric chlorophyll a in organic solvents) to 10 ps? 12.43 The light-induced electron transfer reactions in photosynthesis occur because chlorophyll molecules (whether in monomeric or dimeric forms) are better reducing agents in their electronic excited states. Justify this observation with the help of molecular orbital theory. 12.44 The emission spectrum of a porphyrin dissolved in O2-saturated water shows a strong band at 650 nm and a weak band at 1270 nm. In separate experiments, it was observed that the electronic absorption spectrum of the porphyrin sample showed bands at 420 nm and 550 nm and the electronic absorption spectrum of O2-saturated water showed no bands in the visible range of the spectrum (and therefore no emission spectrum when excited in the same range). Based on these data alone, make a preliminary assignment of the emission band at 1270 nm. Propose additional experiments that test your hypothesis.

Projects 12.45 At the current stage of your study, you have enough knowledge of physical chemistry and biochemistry to begin reading the current literature with a critical eye. Consult monographs, journal articles, and reliable internet resources, such as those listed in the web site for this text, and write a brief report (similar in length and depth of coverage to one of the many Case studies in this text) on each of the following topics.

(a) In confocal Raman microscopy, light must pass through several holes of very small diameter before reaching the detector. In this way light that is out of focus does not interfere with an image that is in focus. Prepare a brief report on the advantages and disadvantages of confocal Raman microscopy over conventional Raman microscopy in the study of biological systems. Hint: A good place to start is P. Colarusso, L.H. Lidder, I.W. Levin, E.N. Lewis, Raman and IR microspectroscopy. In Encyclopedia of spectroscopy and spectrometry (ed. J.C. Lindon, G.E. Tranter, and J.L. Holmes), 3, 1945. Academic Press, San Diego (2000). (b) We have seen throughout the text that it is possible to observe the cooperativity of biopolymer denaturation by determining the extent of denaturation as a function of some parameter that affects its stability, such as temperature or denaturant concentration. Prepare a report summarizing the use of a spectroscopic technique in the study of protein denaturation. Your report should include (i) a description of experimental methods, (ii) a discussion of the information that can be obtained from the measurements, (iii) an example from the literature of the use of the technique in protein stability work, and

(iv) a brief discussion of the advantages and disadvantages of the technique of your choice over differential scanning calorimetry (In the laboratory 1.1), a very popular technique for the study of biopolymer stability. 12.46 The protein hemerythrin (Her) is responsible for binding and carrying O2 in some invertebrates. Each protein molecule has two Fe2+ ions that are in very close proximity and work together to bind one molecule of O2. The Fe2O2 group of oxygenated hemerythrin is colored and has an electronic absorption band at 500 nm.

(a) Figure 12.55 shows the UV–visible absorption spectrum of a derivative of hemerythrin in the presence of different concentrations of CNS− ions. What may be inferred from the spectrum?

Fig. 12.55

512

12 OPTICAL SPECTROSCOPY AND PHOTOBIOLOGY

(b) The resonance Raman spectrum of oxygenated hemerythrin obtained with laser excitation at 500 nm has a band at 844 cm−1 that has been attributed to the O—O stretching mode of bound 16O2. Why is resonance Raman spectroscopy and not infrared spectroscopy the method of choice for the study of the binding of O2 to hemerythrin? (c) Proof that the 844 cm−1 band in the resonance Raman spectrum of oxygenated hemerythrin arises from a bound O2 species may be obtained by conducting experiments on samples of hemerythrin that have been mixed with 18O2 instead of 16O2. Predict the fundamental vibrational wavenumber of the 18O–18O stretching mode in a sample of hemerythrin that has been treated with 18O2. (d) The fundamental vibrational wavenumbers for the O–O stretching modes of O2, O2− (superoxide anion), and O22− (peroxide anion) are 1555, 1107, and 878 cm−1, respectively. (i) Explain this trend in terms of the electronic structures of O2, O2−, and O22−. Hint: Review Case study 10.1. (ii) What are the bond orders of O2, O2−, and O22−? (e) Based on the data given in part (d), which of the following species best describes the Fe2O2 group of hemerythrin: Fe22+O2, Fe2+Fe3+O2−, or Fe23+O22−? Explain your reasoning. (f) The resonance Raman spectrum of hemerythrin mixed with 16 18 O O has two bands that can be attributed to the O–O stretching mode of bound oxygen. Discuss how this observation may be used to exclude one or more of the four proposed schemes (12–15) for binding of O2 to the Fe2 site of hemerythrin.

Fig. 12.56

formation of the AB complex. Write an expression for dR/dt and then show that Req = Rmax

A a0 K D C a0 K + 1F

where Req is the value or R at equilibrium, Rmax is the maximum value that R can have, and a0 is the total concentration of A. To make progress with the derivation, consider that (i) in a typical SPR experiment, the flow rate of A is sufficiently high that [A] = a0 is essentially constant, (ii) we can write [B] = b0 − [AB], where b0 is the total concentration of B, (iii) the SPR signal is often observed to be proportional to [AB], and (iv) the maximum value that R can have is Rmax ∝ b0, which would be measured if all B molecules were ligated to A. (b) Discuss how a plot of a0/Req against a0 can be used to evaluate Rmax and K. (c) Show that, for the association part of the experiment in Fig. 12.56, R(t) = Req(1 − e−k t). obs

(d) Derive an expression for R(t) that applies to the dissociation part of the experiment in Fig. 12.56.

12.47 As an example of the steps taken in biosensor analysis, consider the association of two proteins, A and B. In a typical experiment, a stream of solution containing a known concentration of A flows above the sensor’s surface to which B is attached covalently. Figure 12.56 shows that the kinetics of binding of A to B may be followed by monitoring the time dependence of the surface plasmon resonance (SPR) signal, denoted by R, which is typically the shift in resonance angle. Typically, the system is first allowed to reach equilibrium, which is denoted by the plateau in Fig. 12.56. Then a solution containing no A is flowed above the surface and the AB complex dissociates. Now we see that analysis of the decay of the SPR signal reveals the kinetics of dissociation of the AB complex.

(a) First, show that the equilibrium constant for formation of the AB complex can be measured directly from data of the type displayed in Fig. 12.56. Consider the equilibrium A + B 7 AB

K = kon/koff

where kon and koff are, respectively, the rate constants for formation and dissociation of the AB complex and K is the equilibrium constant for

12.48 The Beer–Lambert law states that the absorbance of a sample at a wavenumber is proportional to the molar concentration [J] of the absorbing species J and to the length L of the sample (eqn 12.5). In this problem you will show that the intensity of fluorescence emission from a sample of J is also proportional to [J] and L. Consider a sample of J that is illuminated with a beam of intensity I0(6) at the wavenumber 6. Before fluorescence can occur, a fraction of I0(6) must be absorbed and an intensity I(6) will be transmitted. However, not all the absorbed intensity is emitted, and the intensity of fluorescence depends on the fluorescence quantum yield, fF, the efficiency of photon emission. The fluorescence quantum yield ranges from 0 to 1 and is proportional to the ratio of the integral of the fluorescence spectrum over the integrated absorption coefficient. Because of a shift of magnitude D6, fluorescence occurs at a wavenumber 6.F, with 6.F + D6. = 6. It follows that the fluorescence intensity at nF, IF(6F), is proportional to fF and to the intensity of exciting radiation that is absorbed by J, Iabs(6) = I0(6) − I(6).

(a) Use the Beer–Lambert law to express Iabs(6) in terms of I0(6), [J], L, and ε(6), the molar absorption coefficient of J at 6. (b) Use your result from part (a) to show that IF(6) ∝ I0(6)ε(6)fF[J]L. (c) In fluorescence excitation spectroscopy, the intensity of emitted radiation at a constant emission wavelength (typically the wavelength at which emission is maximal) is monitored while the excitation wavelength is scanned. Use your results from parts (a) and (b) to

PROJECTS

justify the statement that for a system consisting of a single species, the resulting excitation spectrum is identical to the absorption spectrum of the emitting species. (d) Discuss how fluorescence excitation spectroscopy may be used to provide evidence for resonance energy transfer between a donor and acceptor molecule. The following projects require the use of molecular modeling software. 12.49 We saw in Example 12.3 that water, carbon dioxide, and

methane are able to absorb some of the Earth’s infrared emissions, whereas nitrogen and oxygen cannot. The semiempirical, ab initio, and DFT methods discussed in Chapter 10 can also be used to simulate vibrational spectra, and from the results of the calculation it is possible to determine the correspondence between a vibrational frequency and the atomic displacements that give rise to a normal mode. (a) Using molecular modeling software and the computational method of your instructor’s choice, visualize the vibrational normal modes of CH4, CO2, and H2O in the gas phase. (b) Which vibrational modes of CH4, CO2, and H2O are responsible for absorption of infrared radiation?

513

12.50 Use molecule (16) as a model of the trans conformation of the chromophore found in rhodopsin. In this model, the methyl group bound to the nitrogen atom of the protonated Schiff ’s base replaces the protein.

(a) Using molecular modeling software and the computational method of your instructor’s choice, calculate the energy separation between the hom*o and LUMO of (16). (b) Repeat the calculation for the 11-cis form of (16). (c) Based on your results from parts (a) and (b), do you expect the experimental frequency for the p-to-p* visible absorption of the trans form of (16) to be higher or lower than that for the 11-cis form of (16)?

13 Principles of magnetic resonance

Magnetic resonance

514

13.1

Electrons and nuclei in magnetic fields 515

13.2

The intensities of NMR and EPR transitions

The information in NMR spectra

517 519

13.3

The chemical shift

519

13.4

The fine structure

522

13.5

Conformational conversion and chemical exchange

527

Pulse techniques in NMR

528

13.6

Time- and frequencydomain signals 528

13.7

Spin relaxation

530

In the laboratory 13.1

Magnetic resonance imaging

531

13.8

Proton decoupling

533

13.9

The nuclear Overhauser effect

533

In the laboratory 13.2

Two-dimensional NMR

One of the most widely used and helpful forms of spectroscopy, and a technique that has transformed the practice of chemistry, biochemistry, and medicine, makes use of an effect that is familiar from classical physics. When two pendulums are joined by the same slightly flexible support and one is set in motion, the other is forced into oscillation by the motion of the common axle, and energy flows between the two. The energy transfer occurs most efficiently when the frequencies of the two oscillators are identical. The condition of strong effective coupling when the frequencies are identical is called resonance, and the excitation energy is said to ‘resonate’ between the coupled oscillators. Resonance is the basis of a number of everyday phenomena, including the response of radios to the weak oscillations of the electromagnetic field generated by a distant transmitter. Historically, spectroscopic techniques that measure transitions between nuclear and electron spin states have carried the term ‘resonance’ in their names because they have depended on matching a set of energy levels to a source of monochromatic radiation and observing the strong absorption that occurs at resonance. In this chapter we explore magnetic resonance, a form of spectroscopy that when originally developed (and in some cases still) depends on matching a set of energy levels to a source of monochromatic radiation in the radiofrequency and microwave ranges and observing the strong absorption by magnetic nuclei in nuclear magnetic resonance (NMR) or by unpaired electrons in electron paramagnetic resonance (EPR) that occurs at resonance. Nuclear magnetic resonance is a radiofrequency technique; EPR is a microwave technique. A growing number of structures of biopolymers are now determined by NMR. So powerful is the technique that a clever variation, known as magnetic resonance imaging (MRI), makes possible the spectroscopic characterization of living tissue and has become a major diagnostic tool in medicine.

535

Case study 13.1 The COSY spectrum of isoleucine 536

Principles of magnetic resonance

The information in EPR spectra

537

13.10 The g-value

538

The application of resonance that we describe here depends on the fact that electrons and many nuclei possess spin angular momentum (Table 13.1). An electron in a magnetic field can take two orientations, corresponding to ms = + 12 (denoted a or ↑) and ms = − 12 (denoted b or ↓). A nucleus with nuclear spin quantum number I (the analog of s for electrons and that can be an integer or a half-integer) may take 2I + 1 different orientations relative to an arbitrary axis. These orientations are distinguished by the quantum number mI, which can take on the values mI = I, I − 1, . . . , −I. A proton has I = 12 (the same spin as an electron) and can adopt either of two orientations (mI = + 12 and − 12). A 14N nucleus has I = 1 and can adopt any of

13.11 Hyperfine structure 539 In the laboratory 13.3

Spin probes

540

Checklist of key concepts

541

Checklist of key equations

542

Discussion questions

543

Exercises

543

Projects

545

13.1 ELECTRONS AND NUCLEI IN MAGNETIC FIELDS

Table 13.1

Nuclear constitution and the nuclear spin quantum number

Number of protons

Number of neutrons

I

Even

Even

Odd

Odd

Integer (1, 2, 3, . . .)

Even

Odd

Half-integer ( 12 , 32 , 52 , . . .)

Odd

Even

Half-integer ( 12 , 32 , 52 , . . .)

Table 13.2

Nuclear spin properties

Nucleus

1

H

Natural abundance/percent 99.98

2

H (D)

0.0156

Spin, I

g N /(107 T −1 s−1)

gI

1 2

5.5857

1

0.857 44

26.752 4.1067

12

98.99

13

1.11

1 2

1.4046

6.7272

14

99.64

1

0.403 56

1.9328

16

99.96

C C N O

17

O

0.037

19

100

31

100

F P

35

75.4

37

24.6

Cl Cl

5 2 1 2 1 2 3 2 3 2

— −0.7572

3.627

5.2567

25.177

2.2634

10.840

0.5479

2.624

0.4561

2.184

three orientations (mI = +1, 0, −1). Spin-12 nuclei include protons (1H) and 13C, 19F, and 31P nuclei (Table 13.2). As for electrons, the state with mI = + 12 (↑) is denoted a and that with mI = − 12 (↓) is denoted b. 13.1 Electrons and nuclei in magnetic fields To understand the principles of EPR and NMR we need to understand the magnetic properties of electrons and nuclei.

An electron possesses a magnetic moment due to its spin, and this moment interacts with an external magnetic field. That is, an electron behaves like a tiny bar magnet. The orientation of this magnet is determined by the value of ms, and in a magnetic field B 0 the two orientations have different energies. These energies are given by Em = −gegħB0ms s

Energy of an electron in a magnetic field

(13.1)

Magnetogyric ratio

(13.2)

where g is the magnetogyric ratio of the electron, g=−

e 2me

515

516

13 MAGNETIC RESONANCE

and ge is a factor, the g-value of the electron, which is close to 2.0023 for a free electron.1 The energies are sometimes expressed in terms of the Bohr magneton mB =

eħ 2me

mB = 9.274 × 10−24 J T−1

Bohr magneton

(13.3)

a fundamental unit of magnetism. The symbol T, for tesla, is the unit for reporting the intensity of a magnetic field (1 T = 1 kg s−2A−1). It follows from eqns 13.1 and 13.3 that Alternative expression for the energy of an electron in a magnetic field

Em = ge mBB0ms s

(13.4)

For an electron, the b state lies below the a state. A nucleus with nonzero spin also has a magnetic moment and behaves like a tiny magnet. The orientation of this magnet is determined by the value of mI, and in a magnetic field B 0 the 2I + 1 orientations of the nucleus have different energies. These energies are given by Energy of a nucleus in a magnetic field

Em = −gNħB0mI I

(13.5)

where gN is the nuclear magnetogyric ratio. For spin-12 nuclei with positive magnetogyric ratios (such as 1H), the a state lies below the b state. The energy is sometimes written in terms of the nuclear magneton, mN, mN =

eħ 2mp

mN = 5.051 × 10−27 J T −1

Nuclear magneton

(13.6)

and an empirical constant called the nuclear g-factor, gI, when it becomes Alternative expression for the energy of a nucleus in a magnetic field

Em = −gI mN B0mI I

(13.7)

Nuclear g-factors are experimentally determined dimensionless quantities that vary between −6 and +6 (see Table 13.2). Positive values of gN (and gI) indicate that the nuclear magnet lies in the same direction as the nuclear spin (this is the case for protons). Negative values indicate that the magnet points in the opposite direction. A nuclear magnet is about 2000 times weaker than the magnet associated with electron spin. Two very common nuclei, 12C and 16O, have zero spin and hence are not affected by external magnetic fields. The energy separation of the two spin states of an electron (Fig. 13.1) is DE = Ea − Eb = g m B 0 − (− g m B 0) = ge mBB 0 1 2 e B

1 2 e B

Energy difference between the spin states of an electron in a magnetic field

We infer from the Boltzmann distribution (Fundamentals F.3) that the populations of the a and b states, Na and Nb, are proportional to e−E /kT and e−E /kT, respectively, so the ratio of populations at equilibrium is a

The energy levels of an electron in a magnetic field. Resonance occurs when the energy separation of the levels matches the energy of the photons in the electromagnetic field. Fig. 13.1

(13.8)

Na −(E −E )/kT =e Nb a

b

b

(13.9)

1 The 2 comes from Dirac’s relativistic theory of the electron; the 0.0023 comes from additional correction terms.

13.2 THE INTENSITIES OF NMR AND EPR TRANSITIONS

517

Because Ea − Eb > 0 (the b state lies below the a state), Na /Nb < 1 and there are slightly more b spins than a spins. If the sample is exposed to radiation of frequency n, the energy separations come into resonance with the radiation when the frequency satisfies the resonance condition: hn = ge mBB0 or

n=

ge mBB 0 h

Resonance condition for an electron

(13.10)

At resonance there is strong coupling between the electron spin and the radiation, and strong absorption occurs as the spins flip from b (low energy) to a (high energy). We refer to these transitions as electron paramagnetic resonance (EPR), or electron spin resonance (ESR), transitions. The behavior of nuclei is very similar. The energy separation of the two states of a spin-12 nucleus (Fig. 13.2) is DE = Eb − Ea = g ħB 0 − (− g ħB0) = gNħB 0 1 2 N

1 2 N

Energy difference between the spin states of a nucleus in a magnetic field

(13.11)

Because for nuclei with positive gN the a state lies below the b state, Eb − Ea > 0 and it follows from eqn 13.9 that Nb /Na < 1: there are slightly more a spins than b spins (the opposite of an electron). If the sample is exposed to radiation of frequency n, the energy separations come into resonance with the radiation when the frequency satisfies the resonance condition: hn = gNħB0

or

n=

gNB0 2p

Resonance condition for a nucleus

(13.12)

At resonance there is strong coupling between the nuclear spins and the radiation, and strong absorption occurs as the spins flip from a (low energy) to b (high energy). We refer to these transitions as nuclear magnetic resonance (NMR) transitions.

Self-test 13.1 Calculate the frequency at which radiation comes into reson-

ance with proton spins in a 12 T magnetic field. Answer: 510 MHz

13.2 The intensities of NMR and EPR transitions To appreciate the power of NMR and EPR for investigating biochemical structures and reactions, we need to understand the factors that control the intensities of spin-flipping transitions.

The intensity of an NMR transition depends on a number of factors. We show in the following Justification that intensity ∝ (Na − Nb)B 0

(13.13)

where Na − Nb ≈

NgN ħB 0 2kT

(13.14)

The energy levels of a spin- 12 nucleus (for example, 1 H or 13C) in a magnetic field. Resonance occurs when the energy separation of the levels matches the energy of the photons in the electromagnetic field. Fig. 13.2

518

13 MAGNETIC RESONANCE with N the total number of spins (N = Na + Nb). It follows that decreasing the temperature increases the intensity by increasing the population difference. By combining eqns 13.13 and 13.14, we see that the intensity is proportional to B 02 so NMR transitions can be enhanced significantly by increasing the strength of the applied magnetic field. Similar arguments apply to EPR transitions. We also conclude that absorptions of nuclei with large magnetogyric ratios (1H, for instance) are more intense than those with small magnetogyric ratios (13C, for instance).

Justification 13.1 Intensities in NMR spectra

From the general considerations of transition intensities in Section 12.2, we know that the rate of absorption of electromagnetic radiation is proportional to the population of the lower energy state (Na in the case of a proton NMR transition) and the rate of stimulated emission is proportional to the population of the upper state (Nb). At the low frequencies typical of magnetic resonance, we can neglect spontaneous emission as it is very slow. Therefore, the net rate of absorption is proportional to the difference in populations, and we can write rate of absorption ∝ Na − Nb The intensity of absorption, the rate at which energy is absorbed, is proportional to the product of the rate of absorption (the rate at which photons are absorbed) and the energy of each photon, and the latter is proportional to the frequency n of the incident radiation (through E = hn). At resonance, this frequency is proportional to the applied magnetic field (through n = gNB 0/2p), so we can write intensity of absorption ∝ (Na − Nb)B 0 To write an expression for the population difference, we begin with eqn 13.9, written as A brief comment

The Taylor expansion (Mathematical toolkit 3.2) of an exponential function used in Justification 13.1 is e−x = 1 − x + 12 x 2 − · · · . If x JAX < JMX.

13.21 Calculate the magnetic field needed to satisfy the resonance condition for unshielded protons in a 500.0 MHz radiofrequency field.

13.32 Formulate the version of Pascal’s triangle that you would expect in an NMR spectrum for a collection of N spin-1 nuclei, with N up to 5.

13.22 What is the shift of the resonance from TMS of a group of protons with d = 6.33 in a polypeptide in a spectrometer operating at 420 MHz?

13.33 Show that the coupling constant as expressed by the Karplus equation passes through a minimum when cos f = B/4C. Hint: Evaluate the first derivative with respect to f and set the result equal

544

13 MAGNETIC RESONANCE

to 0. To confirm that the extremum is a minimum, go on to evaluate the second derivative and show that it is positive. 13.34 A proton jumps between two sites with d = 2.7 and d = 4.8. At what rate of interconversion will the two signals collapse to a single line in a spectrometer operating at 500 MHz? 13.35 NMR spectroscopy may be used to determine the equilibrium constant for dissociation of a complex between a small molecule, such as an enzyme inhibitor I, and a protein, such as an enzyme E:

EI 7 E + I

KI = [E][I]/[EI]

In the limit of slow chemical exchange, the NMR spectrum of a proton in I would consist of two resonances: one at nI for free I and another at nEI for bound I. When chemical exchange is fast, the NMR spectrum of the same proton in I consists of a single peak with a resonance frequency n given by n = fInI + fEInEI where fI = [I]/([I] + [EI]) and fEI = [EI]/([I] + [EI]) are, respectively, the fractions of free I and bound I. For the purposes of analyzing the data, it is also useful to define the frequency differences dn = n − nI and Dn = nEI − nI. Show that when the initial concentration of I, [I]0, is much greater than the initial concentration of E, [E]0, a plot of [I]0 versus 1/dn is a straight line with slope [E]0Dn and y-intercept −KI. 13.36 The duration of a 90° pulse depends on the strength of the B1

field. If a 90° pulse requires 10 ms, what is the strength of the B 1 field? 13.37 Interpret the following features of the NMR spectra of hen lysozyme: (a) saturation of a proton resonance assigned to the side chain of methionine-105 changes the intensities of proton resonances assigned to the side chains of tryptophan-28 and tyrosine-23; (b) saturation of proton resonances assigned to tryptophan-28 did not affect the spectrum of tyrosine-23. 13.38 You are designing an MRI spectrometer. What field gradient (in microtesla per meter, mT m−1) is required to produce a separation of 100 Hz between two protons separated by the long diameter of a human kidney (taken as 8 cm) given that they are in environments with d = 3.4? The radiofrequency field of the spectrometer is at 400 MHz and the applied field is 9.4 T. 13.39 Suppose that a uniform disk-shaped organ is in a linear field

gradient and that the MRI signal is proportional to the number of protons in a slice of width dx at each horizontal distance x from the center of the disk. Sketch the shape of the absorption intensity for the MRI image of the disk before any computer manipulation has been carried out. 13.40 Figure 13.45 shows the proton COSY spectrum of 1-nitropropane. Account for the appearance of off-diagonal peaks in the spectrum. 13.41 The proton chemical shifts for the NH, CaH, and CbH groups of alanine are 8.25, 4.35, and 1.39, respectively. Sketch the COSY spectrum of alanine between d = 1.00 and 8.50.

Proton COSY spectrum of 1-nitropropane. The circles show enhanced views of the spectral features. (Spectrum provided by Prof. G. Morris.)

Fig. 13.45

13.42 The center of the EPR spectrum of atomic hydrogen lies at 329.12 mT in a spectrometer operating at 9.2231 GHz. What is the g-value of the electron in the atom? 13.43 A radical containing two equivalent protons shows a three-line spectrum with an intensity distribution 1:2:1. The lines occur at 330.2 mT, 332.5 mT, and 334.8 mT. What is the hyperfine coupling constant for each proton? What is the g-value of the radical given that the spectrometer is operating at 9.319 GHz? 13.44 Predict the intensity distribution in the hyperfine lines of the EPR spectra of (a) ·CH3 and (b) ·CD3. 13.45 The benzene radical anion has g = 2.0025. At what field should you search for resonance in a spectrometer operating at (a) 9.302 GHz and (b) 33.67 GHz? 13.46 The EPR spectrum of a radical with two equivalent nuclei of a particular kind is split into five lines of intensity ratio 1:2:3:2:1. What is the spin of the nuclei? 13.47 Formulate the version of Pascal’s triangle that you would expect to represent the hyperfine structure in an EPR spectrum for a collection of N spin- 32 nuclei, with N up to 5. 13.48 (a) Sketch the EPR spectra of the di-tert-butyl nitroxide radical (5) at 292 K in the limits of very low concentration (at which electron exchange is negligible), moderate concentration (at which electron exchange effects begin to be observed), and high concentration (at which electron exchange effects predominate). (b) Discuss how the observation of electron exchange between nitroxide spin probes can inform the study of lateral mobility of lipids in a biological membrane.

PROJECTS

545

Projects 13.49 Consult library and reliable internet resources, such as those listed on the website for this text, and write a brief report (similar in length and depth of coverage to one of the many Case studies in this text) summarizing the use of NMR or EPR spectroscopy in the study of protein denaturation. Your report should include (a) a description of experimental methods, (b) a discussion of the information that can be obtained from the measurements, (c) an example from the chemical or biological literature of the use of the technique in protein stability work, and (d) a brief discussion of the advantages and disadvantages of the technique of your choice over differential scanning calorimetry (In the laboratory 1.1) and the techniques you described in Exercise 12.45b. 13.50 The following pulse sequence is used in the inversion recovery technique: a 180° pulse is followed by a time interval t, then a 90° pulse, acquisition of a FID curve, and Fourier transformation. A 180° pulse is achieved by applying a B 1 field for twice as long as for a 90° pulse, so the magnetization vector precesses through 180° and points in the −z-direction.

(a) If a 180° pulse requires 12.5 ms, what is the strength of the B 1 field? (b) Draw a series of diagrams showing the effect of the pulse sequence described in part (a) on a sample of equivalent nuclei. The first diagram can be drawn with ease because we already know that the 180° pulse tips the magnetization vector toward the −z-direction. The second diagram should show the effect of spin–lattice relaxation on the magnitude of the magnetization vector after a time interval 0 < t < T1 has elapsed. The third diagram should show the effect of the 90° pulse on the magnetization vector. (c) Why is an FID signal generated after application of the 90° pulse? (d) How does the intensity of the spectrum (obtained by Fourier transformation of the FID curve) vary with the time interval τ, with 0 < t < T1?

(e) Use your results from parts (a)–(d) to show that the inversion recovery technique can be used to measure spin–lattice relaxation times. The following project requires the use of molecular modeling software. The website for this text contains links to freeware and to other sites where you may perform molecular orbital calculations directly from your web browser. 13.51 The molecular electronic structure methods described in Chapter 10 may be used to predict the spin density distribution in a radical. Recent EPR studies have shown that the amino acid tyrosine participates in a number of biological electron transfer reactions, including the processes of water oxidation to O2 in plant photosystem II and of O2 reduction to water in cytochrome c oxidase. During the course of these electron transfer reactions a tyrosine radical forms, with spin density delocalized over the side chain of the amino acid.

(a) The phenoxy radical shown in (7) is a suitable model of the tyrosine radical. Using molecular modeling software and the computational method of your instructor’s choice, calculate the spin densities at the O atom and at all of the C atoms in (7). (b) Predict the form of the EPR spectrum of (7).

Resource section 1: Atlas of structures In this section are displayed structures of biologically significant molecules that occur throughout the text. They are arranged as follows: Section A Section B Section C Section E Section L Section M Section N Section P Section R Section S Section T

Amino acids Bases Carboxylic acids Polyenes Lipids Miscellaneous Nucleotides Proteins Porphyrin-based ring complexes Saccharides Nucleic acids

For proteins, we give the appropriate Protein Data Bank reference.

RESOURCE SECTION 1: ATLAS OF STRUCTURES

A Amino acids

547

548

RESOURCE SECTION 1: ATLAS OF STRUCTURES

B Bases

C Carboxylic acids

RESOURCE SECTION 1: ATLAS OF STRUCTURES

E Polyenes

549

550

RESOURCE SECTION 1: ATLAS OF STRUCTURES

L Lipids

RESOURCE SECTION 1: ATLAS OF STRUCTURES

M Miscellaneous

551

552

RESOURCE SECTION 1: ATLAS OF STRUCTURES

N Nucleotides

RESOURCE SECTION 1: ATLAS OF STRUCTURES

N Nucleotides continued

553

554

RESOURCE SECTION 1: ATLAS OF STRUCTURES

P Proteins

RESOURCE SECTION 1: ATLAS OF STRUCTURES

P Proteins continued

555

556

RESOURCE SECTION 1: ATLAS OF STRUCTURES

R Porphyrin-based ring complexes

RESOURCE SECTION 1: ATLAS OF STRUCTURES

S Saccharides

T Nucleic acids

557

Resource section 2: Units Table 1

The SI base units

Physical quantity

Symbol for quantity

Base unit

Length

l

meter, m kilogram, kg

Mass

m

Time

t

second, s

Electric current

I

ampere, A

Thermodynamic temperature

T

kelvin, K

Amount of substance

n

mole, mol

Luminous intensity

Iv

candela, cd

Table 2

A selection of derived units

Physical quantity

Derived unit*

Name of derived unit

Force

1 kg m s−2

newton, N

−1 −2

Pressure

pascal, Pa

1 kg m s 1 N m−2

1 kg m2 s−2

Energy

joule, J

1Nm 1 Pa m3 1 kg m2 s−3

Power

watt, W

−1

1Js

*Equivalent definitions in terms of derived units are given following the definition in terms of base units.

Table 3

Common SI prefixes

Prefix z

a

Name zepto atto Factor 10−21

f

p

n

μ

m

c

d

k

M

G

T

femto pico nano micro milli centi deci kilo mega giga tera

10−18 10−15

10−12 10−9

10−6

10−3

10−2

10−1 103

106

109

1012

RESOURCE SECTION 2: UNITS

Table 4

Some common units

Physical quantity

Name of unit

Symbol for unit

Value*

Time

minute

min

60 s

hour

h

3600 s

Length

ångström

Å

10−10 m

Volume

liter

L, l

1 dm3

Mass

tonne

t

103 kg

Pressure

bar

bar

105 Pa

Energy

atmosphere

atm

101.325 kPa

electronvolt

eV

1.602 176 × 10−19 J 96.485 31 kJ mol−1

*All values in the final column are exact, except for the definition of 1 eV.

559

Resource section 3: Data Table 1

Thermodynamic data for organic compounds (all values relate to 298.15 K) M/g mol−1

D f H 9/ kJ mol−1 0

D f G 9/ kJ mol−1

9 C p,m / J K−1 mol−1

5.740

8.527

−393.51

+2.377

6.113

−395.40

D c H 9/ kJ mol−1

C(s) (graphite)

12.011

C(s) (diamond)

12.011

CO2(g)

44.010

−393.51

−394.36

213.74

37.11

CH4(g), methane

16.04

−74.81

−50.72

186.26

35.31

CH3(g), methyl

15.04

+145.69

+147.92

194.2

38.70

C2H2(g), ethyne

26.04

+226.73

+209.20

200.94

43.93

−1300

C2H4(g), ethene

28.05

+52.26

+68.15

219.56

43.56

−1411

C2H6(g), ethane

30.07

−84.68

−32.82

229.60

52.63

−1560

C3H6(g), propene

42.08

+20.42

+62.78

267.05

63.89

−2058

C3H6(g), cyclopropane

42.08

−103.85

−23.49

269.91

73.5

−2220

C4H8(g), 1-butene

56.11

−0.13

+71.39

305.71

85.65

−2717

C4H8(g), cis-2-butene

56.11

−6.99

+65.95

300.94

78.91

−2710

+1.895

Sm9 / J K−1 mol−1

+2.900

Hydrocarbons −890

C4H8(g), trans-2-butene

56.11

−11.17

+63.06

296.59

87.82

−2707

C4H10(g), butane

58.13

−126.15

−17.03

310.23

97.45

−2878

C5H12(g), pentane

72.15

−146.44

−8.20

348.40

120.2

C5H12(l)

72.15

−173.1

C6H6(l), benzene

78.12

+49.0

+124.3

173.3

136.1

C6H6(g)

78.12

+82.93

+129.72

269.31

C6H12(l), cyclohexane

84.16

−156

C6H14(l), hexane

86.18

−198.7

C6H5CH3(g), methylbenzene (toluene)

92.14

+50.0

100.21

−224.4

C8H18(l), octane

114.23

−249.9

C8H18(l), iso-octane

114.23

−255.1

C10H8(s), naphthalene

128.18

+78.53

C7H16(l), heptane

26.8

81.67 156.5

−3537 −3268 −3320 −3902 −4163

204.3 +122.0

320.7

103.6

+1.0

328.6

224.3

+6.4

361.1

−3953 −5471 −5461 −5157

Alcohols and phenols CH3OH(l), methanol

32.04

−238.86

−166.27

126.8

81.6

CH3OH(g)

32.04

−200.66

−166.27

239.81

43.89

−726 −764

C2H5OH(l), ethanol

46.07

−277.69

−174.78

160.7

111.46

−1368

C2H5OH(g)

46.07

−235.10

−168.49

282.70

65.44

−1409

C6H5OH(s), phenol

94.12

−165.0

−50.9

146.0

−3054

RESOURCE SECTION 3: DATA

M/g mol−1

D f H 9/ kJ mol−1

D f G 9/ kJ mol−1

Sm9 / J K−1 mol−1

9 C p,m / J K−1 mol−1

561

D c H 9/ kJ mol−1

Carboxylic acids, hydroxy acids, and esters HCOOH(l), formic

46.03

−424.72

−361.35

128.95

CH3COOH(l), ethanoic

60.05

−484.3

−389.9

159.8

CH3COOH(aq)

60.05

−485.76

−396.46

178.7

CH3CO2 (aq)

59.05

−486.01

−369.31

86.6

CH3(CO)COOH(l), pyruvic

88.06

CH3(CH2)2COOH(l), butanoic

88.10

−533.8

CH3COOC2H5(l), ethyl acetate

88.10

−479.0

(COOH)2(s), oxalic

90.04

−827.2

99.04 124.3

−875

−6.3 −950

−332.7

259.4

170.1 117

90.08

−694.0

−522.9

HOOCCH2CH2COOH(s), succinic

118.09

−940.5

−747.4

153.1

167.3

C6H5COOH(s), benzoic

122.13

−385.1

−245.3

167.6

146.8

CH3(CH2)8COOH(s), decanoic

172.27

−713.7

C6H8O6(s), ascorbic

176.12

−1164.6

HOOCCH2C(OH)(COOH) CH2COOH(s), citric

192.12

−1543.8

CH3CH(OH)COOH(s), lactic

−255

−2231 −254 −1344

−1236.4

−3227

−1985

CH3(CH2)10COOH(s), dodecanoic

200.32

−774.6

CH3(CH2)14COOH(s), hexadecanoic

256.41

−891.5

C18H36O2(s), stearic

284.48

−947.7

HCHO(g), methanal

30.03

−108.57

−102.53

218.77

CH3CHO(l), ethanal

44.05

−192.30

−128.12

160.2

CH3CHO(g)

44.05

−166.19

−128.86

250.3

57.3

−1192

CH3COCH3(l), propanone

58.08

−248.1

−155.4

200.4

124.7

−1790

−917.2

212.1

−918.8

205.4

404.3 501.5

Alkanals and alkanones 35.40

−571 −1166

Sugars C5H10O5(s), d-ribose

150.1

−1051.1

C5H10O5(s), d-xylose

150.1

−1057.8

C6H12O6(s), a-d-glucose

180.16

−1273.3

C6H12O6(s), b-d-glucose

180.16

−1268

C6H12O6(s), b-d-fructose

180.16

−1265.6

C6H12O6(s), a-d-galactose

180.16

−1286.3

C12H22O11(s), sucrose

342.30

−2226.1

−1543

360.2

C12H22O11(s), lactose

342.30

−2236.7

−1567

386.2

Amino acids

−2808 −2810 −5645

1

l-Glycine solid

75.07

−528.5

−373.4

103.5

aqueous solution

75.07

−469.8

−315.0

111.0

89.09

−604.0

−369.9 −508.8

l-Alanine

99.2

−969

129.2

122.2

−1618

149.2

135.6

−1455

164.0

151.2

l-Serine

105.09

−732.7

l-Proline

115.13

−515.2

l-Valine

117.15

−617.9

−359.0

178.9

168.8

−2922

l-Threonine

119.12

−807.2

−550.2

152.7

147.3

−2053

l-Cysteine

121.16

−534.1

−340.1

169.9

162.3

−1651

562

RESOURCE SECTION 3: DATA

M/g mol−1

D f H 9/ kJ mol−1

D f G 9/ kJ mol−1

Sm9 / J K−1 mol−1

9 C p,m / J K−1 mol−1

D c H 9/ kJ mol−1

l-Leucine

131.17

−637.4

−347.7

211.8

200.1

−3582

l-Isoleucine

131.17

−637.8

−347.3

208.0

188.3

−3581

l-Asparagine

132.12

−789.4

−530.1

174.5

160.2

−530

l-Aspartic acid

133.10

−973.3

−730.1

170.1

155.2

−1601

l-Glutamine

146.15

−826.4

−532.6

195.0

184.2

−2570

l-Glutamic acid

147.13

−1009.7

−731.4

188.2

175.0

−2244

l-Methionine

149.21

−577.5

−505.8

231.5

290.0

−2782

l-Histidine

155.16

−466.7

l-Phenylalanine

165.19

−466.9

−211.7

213.6

203.0

−4647

l-Tyrosine

181.19

−685.1

−385.8

214.0

216.4

−4442

l-Tryptophan

204.23

−415.3

−119.2

251.0

238.1

−5628

l-Cystine

240.32

−1032.7

−685.8

280.6

261.9

−3032

NH2CH2CONHCH2COOH(s), glycylglycine

132.12

747.7

NH2CH(CH3)CONHCH2COOH, alanylglycine

146.15

Peptides 487.9

180.3

164.0

1972

489.9

213.4

182.4

2619

+32.16

243.41

53.1

−1085

−197.33

104.60

93.14

Other nitrogen compounds −22.97

CH3NH2(g), methylamine

31.06

(NH2)2CO(s), urea

60.06

−333.1

C6H5NH2(l), aniline

93.13

+31.1

C4H5N3O(s), cytosine

111.10

−221.3

C4H4N2O2(s), uracil

112.09

−429.4

C5H6N2O2(s), thymine

126.11

−462.8

C5H5N5(s), adenine

135.14

+96.9

+299.6

151.1

C5H5N5O(s), guanine

151.13

−183.9

+47.4

160.3

−632 −3393

132.6 150.8 147.0

1

See the Atlas of structures, Section A, for the molecular structures of the amino acids. Unless otherwise noted, data relate to the substance in the solid state.

Table 2

Thermodynamic data (all values relate to 298.15 K)* M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

9 C p,m /(J K −1 mol−1)

Aluminum Al(s)

26.98

0 +10.56

Al(l)

26.98

Al(g)

26.98

+326.4

Al3+(g)

26.98

+5483.17

0 +7.20 +285.7

28.33

24.35

39.55

24.21

164.54

21.38

Al3+(aq)

26.98

−531

−485

Al2O3(s, )

101.96

−1675.7

−1582.3

−321.7 50.92

79.04

AlCl3(s)

133.24

−704.2

−628.8

110.67

91.84

154.84

20.786

Argon Ar(g)

39.95

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

9 C p,m /(J K −1 mol−1)

Antimony Sb(s)

121.75

45.69

25.23

SbH3(g)

153.24

+145.11

+147.75

232.78

41.05

74.92

35.1

24.64

174.21

20.79

Arsenic As(s,a) As(g)

74.92

+302.5

+261.0

As4(g)

299.69

+143.9

+92.4

314

+68.93

222.78

AsH3(g)

77.95

+66.44

38.07

Barium Ba(s)

137.34

Ba(g)

137.34

+180

+146

62.8

28.07

170.24

20.79

Ba2+(aq)

137.34

−537.64

−560.77

BaO(s)

153.34

−553.5

−525.1

70.43

47.78

BaCl2(s)

208.25

−858.6

−810.4

123.68

75.14

+9.6

9.50

16.44

136.27

20.79

56.74

25.52

187.00

20.79

152.23

75.689

Beryllium Be(s)

9.01

Be(g)

9.01

0 +324.3

0 +286.6

Bismuth Bi(s)

208.98

Bi(g)

208.98

0 +207.1

0 +168.2

Bromine Br2(l)

159.82

Br2(g)

159.82

Br(g)

79.91

+111.88

Br−(g)

79.91

−219.07

Br (aq)

79.91

−121.55

−103.96

+82.4

HBr(g)

90.92

−36.40

−53.45

198.70

0 +30.907

0 +3.110

245.46

36.02

+82.396

175.02

20.786 −141.8 29.142

Cadmium Cd(s,γ)

112.40

51.76

25.98

Cd(g)

112.40

+112.01

+77.41

167.75

20.79

−77.612

−73.2

−75.90

Cd2+(aq)

112.40

CdO(s)

128.40

−258.2

−228.4

54.8

CdCO3(s)

172.41

−750.6

−669.4

92.5

43.43

Caesium: see cesium Calcium Ca(s)

40.08

41.42

25.31

Ca(g)

40.08

+178.2

+144.3

154.88

20.786

Ca2+(aq)

40.08

−542.83

−553.58

−53.1

CaO(s) CaCO3(s) (calcite)

56.08 100.09

−635.09 −1206.9

−604.03 −1128.8

563

39.75

42.80

92.9

81.88

564

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

100.09

−1207.1

−1127.8

88.7

CaF2(s)

78.08

1219.6

−1167.3

68.87

CaCl2(s)

110.99

−795.8

−748.1

104.6

CaBr2(s)

199.90

−682.8

−663.6

130

CaCO3(s) (aragonite)

9 C p,m /(J K −1 mol−1)

81.25 67.03 72.59

Carbon ( for ‘organic’ compounds, see Table 1) C(s) (graphite)

12.011

C(s) (diamond)

12.011

C(g)

12.011

+716.68

+671.26

158.10

20.838

C2(g)

24.022

+831.90

+775.89

199.42

43.21

CO(g)

28.011

−110.53

−137.17

197.67

29.14

CO2(g)

44.010

−393.51

−394.36

213.74

37.11

CO2(aq)

44.010

−413.80

−385.98

117.6

H2CO3(aq)

62.03

−699.65

−623.08

187.4

HCO3−(aq)

61.02

−691.99

−586.77

+91.2

60.01

−677.14

−527.81

−56.9

CCl4(l)

153.82

−135.44

−65.21

216.40

131.75

CS2(l)

76.14

+89.70

+65.27

151.34

75.7

2− 3

CO (aq)

0 +1.895

0 +2.900

5.740 2.377

8.527 6.133

HCN(g)

27.03

+135.1

+124.7

201.78

35.86

HCN(l)

27.03

+108.87

+124.97

112.84

70.63

CN−(aq)

26.02

+150.6

+172.4

+94.1

Cesium Cs(s)

132.91

85.23

Cs(g)

132.91

+76.06

+49.12

175.60

Cs+(aq)

132.91

−258.28

−292.02

+133.05

32.17 20.79 −10.5

Chlorine Cl2(g)

70.91

223.07

33.91

Cl(g)

35.45

+121.68

+105.68

165.20

21.840

Cl−(g)

35.45

−233.13

Cl (aq)

35.45

−167.16

−131.23

+56.5

HCl(g)

36.46

−92.31

−95.30

186.91

HCl(aq)

36.46

−167.16

−131.23

56.5

52.00

23.77

23.35 20.79

−136.4 29.12 −136.4

Chromium Cr(s)

52.00

+396.6

+351.8

174.50

CrO42−(aq)

115.99

−881.15

−727.75

+50.21

Cr2O72−(aq)

215.99

Cr(g)

−1490.3

−1301.1

+261.9

Copper Cu(s)

63.54

Cu(g)

63.54

+338.32

+298.58

166.38

Cu+(aq)

63.54

+71.67

+49.98

+40.6

Cu2+(aq)

63.54

+64.77

+65.49

−99.6

Cu2O(s)

143.08

−168.6

−146.0

33.150

93.14

24.44 20.79

63.64

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

9 C p,m /(J K −1 mol−1)

79.54

−157.3

−129.7

CuSO4(s)

159.60

−771.36

−661.8

109

100.0

CuSO4·H2O(s)

177.62

−1085.8

−918.11

146.0

134

CuSO4·5H2O(s)

249.68

−2279.7

300.4

280

CuO(s)

−1879.7

42.63

42.30

Deuterium D2(g)

4.028

HD(g)

3.022

D2O(g)

20.028

0 +0.318 −249.20

0 −1.464 −234.54

144.96

29.20

143.80

29.196

198.34

34.27

D2O(l)

20.028

−294.60

−243.44

75.94

84.35

HDO(g)

19.022

−245.30

−233.11

199.51

33.81

HDO(l)

19.022

−289.89

−241.86

79.29

F2(g)

38.00

202.78

Fluorine F(g)

19.00

+78.99

+61.91

158.75

F −(aq)

19.00

−332.63

−278.79

−13.8

HF(g)

20.01

−271.1

−273.2

173.78

31.30 22.74 −106.7 29.13

Gold Au(s)

196.97

Au(g)

196.97

0 +366.1

0 +326.3

47.40

25.42

180.50

20.79

126.15

20.786

Helium He(g)

4.003

Hydrogen (see also deuterium) H2(g)

2.016

130.684

28.824

H(g)

1.008

+217.97

+203.25

114.71

20.784

H+(aq)

1.008

H2O(l)

18.015

−285.83

−237.13

69.91

75.291

H2O(g)

18.015

−241.82

−228.57

188.83

33.58

H2O2(l)

34.015

−187.78

−120.35

109.6

89.1

116.135

54.44

Iodine I2(s)

253.81

I2(g)

253.81

+62.44

+19.33

260.69

36.90

I(g)

126.90

+106.84

+70.25

180.79

20.786

I−(aq)

126.90

−55.19

−51.57

HI(g)

127.91

+26.48

+1.70

+111.3 206.59

−142.3 29.158

Iron Fe(s)

55.85

Fe(g)

55.85

+416.3

Fe2+(aq)

55.85

−89.1

0 +370.7

27.28

25.10

180.49

25.68

−78.90

−137.7

55.85

−48.5

−4.7

−315.9

Fe3O4(s) (magnetite)

231.54

−1184.4

−1015.4

146.4

Fe2O3(s) (hematite)

159.69

−824.2

−742.2

3+

Fe (aq)

87.40

565

143.43 103.85

566

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

9 C p,m /(J K −1 mol−1)

FeS(s,a)

87.91

−100.0

−100.4

60.29

50.54

FeS2(s)

119.98

−178.2

−166.9

52.93

62.17

164.08

20.786

Krypton Kr(g)

83.80

Lead Pb(s)

207.19

Pb(g)

207.19

+195.0

Pb2+(aq)

207.19

−1.7

0 +161.9 −24.43

64.81

26.44

175.37

20.79

+10.5

PbO(s, yellow)

223.19

−217.32

−187.89

68.70

45.77

PbO(s, red)

223.19

−218.99

−188.93

66.5

45.81

PbO2(s)

239.19

−277.4

−217.33

68.6

64.64

Lithium Li(s)

6.94

29.12

24.77

Li(g)

6.94

+159.37

+126.66

138.77

20.79

Li+(aq)

6.94

−278.49

−293.31

+13.4

+68.6

Magnesium Mg(s)

24.31

32.68

24.89

Mg(g)

24.31

+147.70

+113.10

148.65

20.786

Mg 2+(aq)

24.31

−466.85

−454.8

−601.70

−569.43

MgO(s)

40.31

MgCO3(s)

84.32

MgCl2(s)

95.22

−641.32

−591.79

MgBr2(s)

184.13

−524.3

−503.8

−1095.8

−1012.1

−138.1 26.94

37.15

65.7

75.52

89.62

71.38

117.2

Mercury Hg(l)

200.59

76.02

27.983

Hg(g)

200.59

+61.32

+31.82

174.96

20.786

200.59

+171.1

+164.40

−32.2

2+ 2

Hg (aq)

401.18

+172.4

+153.52

+84.5

HgO(s)

216.59

−90.83

−58.54

Hg2Cl2(s)

472.09

−265.22

−210.75

192.5

HgCl2(s)

271.50

−224.3

−178.6

146.0

HgS(s, black)

232.65

−53.6

−47.7

88.3

20.18

146.33

20.786

28.013

191.61

29.125

Hg 2+(aq)

70.29

44.06 102

Neon Ne(g) Nitrogen N2(g) N(g)

14.007

+472.70

+455.56

153.30

20.786

NO(g)

30.01

+90.25

+86.55

210.76

29.844

N2O(g)

44.01

+82.05

+104.20

219.85

38.45

NO2(g)

46.01

+33.18

+51.31

240.06

37.20

N2O4(g)

92.01

+9.16

+97.89

304.29

77.28

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

9 C p,m /(J K −1 mol−1)

N2O5(s)

108.01

−43.1

+113.9

178.2

143.1

N2O5(g)

108.01

+11.3

+115.1

355.7

84.5

HNO3(l)

63.01

−174.10

−80.71

155.60

109.87

HNO3(aq)

63.01

−207.36

−111.25

146.4

−86.6

NO3−(aq)

62.01

−205.0

−108.74

+146.4

−86.6

NH3(g)

17.03

−46.11

−16.45

192.45

NH3(aq)

17.03

−80.29

−26.50

113.3

NH4+(aq)

18.04

−132.51

−79.31

+113.4

NH2OH(s)

33.03

−114.2

HN3(l)

43.03

+264.0

+327.3

140.6

NH3(g)

43.03

+294.1

+328.1

238.97

35.06 +79.9

43.68

N2H4(l)

32.05

+50.63

+149.43

121.21

NH4NO3(s)

80.04

−365.56

−183.87

151.08

139.3

NH4Cl(s)

53.49

−314.43

−202.87

94.6

84.1

O2(g)

31.999

205.138

29.355 21.912

98.87

Oxygen O(g)

15.999

+249.17

+231.73

161.06

O3(g)

47.998

+142.7

+163.2

238.93

OH −(aq)

17.007

−229.99

−157.24

−10.75

39.20 −148.5

Phosphorus P(s, wh)

30.97

41.09

23.840

P(g)

30.97

+314.64

+278.25

163.19

20.786

P2(g)

61.95

+144.3

+103.7

P4(g)

123.90

+58.91

218.13

32.05

+24.44

279.98

67.15

PH3(g)

34.00

+5.4

+13.4

210.23

37.11

PCl3(g)

137.33

−287.0

−267.8

311.78

71.84

PCl3(l)

137.33

−319.7

−272.3

217.1

PCl5(g)

208.24

−374.9

−305.0

364.6

112.8

PCl5(s)

208.24

−443.5

H3PO3(s)

82.00

−964.4

H3PO3(aq)

82.00

−964.8

H3PO4(s)

94.97

−1279.0

−1119.1

110.50

106.06

H3PO4(l)

94.97

−1266.9

H3PO4(aq)

94.97

−1277.4

−1018.7

−222

PO43−(aq)

94.97

−1277.4

−1018.7

−222

P4O10(s)

283.89

−2984.0

−2697.0

P4O6(s)

219.89

−1640.1

228.86

211.71

Potassium K(s)

39.10

64.18

29.58

K(g)

39.10

+89.24

+60.59

160.336

20.786

K+(g)

39.10

+514.26

39.10

−252.38

+

K (aq)

−283.27

+102.5

567

+21.8

568

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

D f G 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

9 C p,m /(J K −1 mol−1)

KOH(s)

56.11

−424.76

−379.08

78.9

64.9

KF(s)

58.10

−576.27

−537.75

66.57

49.04

KCl(s)

74.56

−436.75

−409.14

82.59

51.30

KBr(s)

119.01

−393.80

−380.66

95.90

52.30

Kl(s)

166.01

−327.90

−324.89

106.32

52.93

Si(s)

28.09

18.83

20.00

Si(g)

28.09

+455.6

+411.3

167.97

22.25

SiO2(s,a)

60.09

−910.93

−856.64

41.84

44.43

Silicon

Silver Ag(s)

107.87

42.55

25.351

Ag(g)

107.87

+284.55

+245.65

173.00

20.79

Ag+(aq)

107.87

+105.58

+77.11

+72.68

AgBr(s)

187.78

−100.37

−96.90

107.1

52.38

AgCl(s)

143.32

−127.07

−109.79

96.2

50.79

Ag2O(s)

231.74

−31.05

−11.20

121.3

65.86

AgNO3(s)

169.88

−124.39

−33.41

140.92

93.05

28.24

+21.8

Sodium Na(s)

22.99

51.21

Na(g)

22.99

+107.32

+76.76

153.71

Na+(aq)

22.99

−240.12

−261.91

+59.0

20.79 +46.4

NaOH(s)

40.00

−425.61

−379.49

64.46

59.54

NaCl(s)

58.44

−411.15

−384.14

72.13

50.50

NaBr(s)

102.90

−361.06

−348.98

86.82

51.38

NaI(s)

149.89

−287.78

−286.06

98.53

52.09

Sulfur S(s,a) (rhombic)

32.06

S(s,b) (monoclinic)

32.06

+0.33

31.80

22.64

+0.1

32.6

23.6

S(g)

32.06

+278.81

+238.25

167.82

23.673

S2(g)

64.13

+128.37

S2−(aq)

32.06

+33.1

+79.30

228.18

32.47

+85.8

−14.6

SO2(g)

64.06

−296.83

−300.19

248.22

SO3(g)

80.06

−395.72

−371.06

256.76

H2SO4(l)

98.08

−813.99

−690.00

156.90

H2SO4(aq)

98.08

−909.27

−744.53

20.1

−293

39.87 50.67 138.9

SO (aq)

96.06

−909.27

−744.53

+20.1

−293

HSO4−(aq)

97.07

−887.34

−755.91

+131.8

−84

H2S(g)

34.08

−20.63

−33.56

H2S(aq)

34.08

−39.7

−27.83

121

HS −(aq)

33.072

−17.6

+12.08

+62.08

2− 4

SF6(g)

146.05

−1209

−1105.3

205.79

291.82

34.23

97.28

RESOURCE SECTION 3: DATA

M/(g mol−1)

D f H 9/(kJ mol−1)

Sm9 /(J K −1 mol−1)

D f G 9/(kJ mol−1)

569

9 C p,m /(J K −1 mol−1)

Tin Sn(s,b)

118.69

Sn(g)

118.69

0 +302.1

0 +267.3

51.55

26.99

168.49

20.26

Sn2+(aq)

118.69

−8.8

−27.2

SnO(s)

134.69

−285.8

−256.8

−17 56.5

44.31

SnO2(s)

150.69

−580.7

+519.6

52.3

52.59

169.68

20.786

Xenon Xe(g)

131.30

Zinc Zn(s)

65.37

41.63

25.40

Zn(g)

65.37

+130.73

+95.14

160.98

20.79

Zn2+(aq)

65.37

−153.89

−147.06

ZnO(s)

81.37

−348.28

−318.30

−112.1

+46

43.64

40.25

Reduction half-reaction

E 9/V

Cu2+ + e− → Cu+

+0.16

*Entropies and heat capacities of ions are relative to H+(aq) and are given with a sign.

Standard potentials at 298.15 K in electrochemical order

Table 3a

E 9/V

Reduction half-reaction Strongly oxidizing H4XeO6 + 2 H + 2 e → XeO3 + 3 H2O

+3.0

Sn4+ + 2 e− → Sn2+

+0.15

F2 + 2 e− → 2 F −

+2.87

AgBr + e− → Ag + Br−

+0.07

+

O3 + 2 H + 2 e → O2 + H2O

+2.07

Ti + e → Ti

S2O82− + 2 e− → 2 SO42−

+2.05

2 H+ + 2 e− → H

Ag2+ + e− → Ag+

+1.98

Fe3+ + 3 e− → Fe

+

4+

3+

0.00 0, by definition −0.04

Co + e → Co

+1.81

O2 + H2O + 2 e → HO + OH

HO2 + 2 H+ + 2 e− → 2 H2O

+1.78

Pb2+ + 2 e− → Pb

−0.13 −0.14

3+

2+

− 2

−0.08

Au + e → Au

+1.69

In + e → In

Pb4+ + 2 e− → Pb2+

+1.67

Sn2+ + 2 e− → Sn

−0.14

2 HClO + 2 H + + 2 e− → Cl2 + 2 H2O

+1.63

Agl + e− → Ag + I

−0.15

Ce4+ + e− → Ce3+

+1.61

Ni2+ + 2 e− → Ni

−0.23

2 HBrO + 2 H + 2 e → Br2 + 2 H

+1.60

Co2+ + 2 e− → Co

−0.28

MnO4− + 8 H + + 5 e− → Mn2+ + 4 H2O

+1.51

In3+ + 3 e− → In

−0.34

+

+

+

Mn + e → Mn

+1.51

Tl + e → Tl

−0.34

Au3+ + 3 e− → Au

+1.40

PbSO4 + 2 e− → Pb + SO2− 4

−0.36

Cl2 + 2 e− → 2 Cl−

+1.36

Ti3+ + e− → Ti2+

−0.37

3+

Cr2O + 14 H + 6 e → 2 Cr + 7 H2O

+1.33

Cd + 2 e → Cd

−0.40

O3 + H2O + 2 e− → O2 + 2 OH −

+1.24

In2+ + e− → In+

−0.40

O2 + 4 H + + 4 e− → 2 H2O

+1.23

Cr 3+ + e− → Cr2+

−0.41

ClO + 2 H + 2 e → ClO + H2O

+1.23

Fe + 2 e → Fe

−0.44

MnO2 + 4 H + + 2 e− → Mn2+ + 2 H2O

+1.23

In3+ + 2 e− → In+

−0.44

3+

2+

+

2− 7

− 4

+

− 3

+

2+

2+

570

RESOURCE SECTION 3: DATA

Reduction half-reaction

E 9/V

Reduction half-reaction

E 9/V

Br2 + 2 e− → 2 Br−

+1.09

S + 2 e− → S2−

−0.48

Pu + e → Pu

+0.97

In3+ + e − → In2+

−0.49

NO3− + 4 H+ + 3 e− → NO + 2 H2O

+0.96

U4+ + e− → U3+

−0.61

2 Hg + 2 e → Hg

−0.74

4+

3+

+0.92

Cr + 3 e → Cr

ClO− + H2O + 2 e− → Cl− + 2 OH −

+0.89

Zn2+ + 2 e− → Zn

2+

2+ 2

3+

−0.76

Hg + 2 e → Hg

+0.86

Cd(OH)2 + 2 e → Cd + 2 OH

NO3− + 2 H+ + e− → NO2 + H2O

+0.80

2 H2O + 2 e− → H2 + 2 OH−

−0.83

Ag+ + e− → Ag

+0.80

Cr 2+ + 2 e− → Cr

−0.91

2+

−0.81

Hg + 2 e → 2 Hg

+0.79

Mn + 2 e → Mn

−1.18

Fe3+ + e− → Fe2+

+0.77

V2+ + 2 e− → V

−1.19

BrO− + H2O + 2 e− → Br− + 2 OH−

+0.76

Ti2+ + 2 e− → Ti

−1.63

Hg2SO4 + 2 e → 2 Hg + SO

2+ 2

2+

+0.62

Al + 3 e → Al

−1.66

MnO42− + 2 H2O + 2 e− → MnO2 + 4 OH −

+0.60

U3+ + 3 e− → U

−1.79

MnO4− + e− → MnO42−

+0.56

Mg 2+ + 2 e− → Mg

−2.36

I2 + 2 e → 2 I

+0.54

Ce + 3 e → Ce

−2.48

Cu+ + e− → Cu

+0.52

La3+ + 3 e− → La

−2.52

+0.53

Na+ + e− → Na

−2.71 −2.87

2− 4

I 3− + 2 e− → 3 I− NiOOH + H2O + e → Ni(OH)2OH

3+

3+

+0.49

Ca + 2 e → Ca

IAg2CrO4 + 2 e− → 2 Ag + CrO42−

+0.45

Sr 2+ + 2 e− → Sr

−2.89

O2 + 2 H2O + 4 e− → 4 OH−

+0.40

Ba2+ + 2 e− → Ba

−2.91

ClO4− + H2O + 2 e− → ClO3− + 2 OH −

+0.36

Ra2+ + 2 e− → Ra

−2.92

[Fe(CN)6] + e → [Fe(CN)6]

+0.36

Cs+ + e− → Cs

−2.92

Cu2+ + 2 e− → Cu

+0.34

Rb+ + e− → Rb

−2.93

3−

Hg2Cl2 + 2 e → 2 Hg + 2 Cl

4−

2+

+0.27

K +e →K

−2.93

AgCl + e− → Ag + Cl−

+0.22

Li+ + e− → Li

−3.05

Bi + 3 e → Bi

+0.20

Strongly reducing

3+

Table 3b

+

Standard potentials at 298.15 K in alphabetical order

Reduction half-reaction

E 9/V

Reduction half-reaction

E 9/V

Ag+ + e− → Ag

+0.80

I2 + 2 e− → 2 I−

+0.54

Ag + e → Ag

+1.98

I +2e →3I

+0.53

AgBr + e− → Ag + Br−

+0.0713

In+ + e− → In

−0.14

+0.22

In2+ + e− → In+

−0.40

+0.45

In3+ + 2 e− → In+

−0.44

AgF + e− → Ag + F −

+0.78

In3+ + 3 e− → In

−0.34

Agl + e− → Ag + I−

−0.15

In3+ + e− → In2+

−0.49

Al + 3 e → Al

−1.66

K+ + e− → K

−2.93

Au+ + e− → Au

+1.69

La3+ + 3 e− → La

−2.52

Au + 3 e → Au

+1.40

Li + e → Li

−3.05

Ba2+ + 2 e− → Ba

−2.91

Mg 2+ + 2 e− → Mg

−2.36

2+

+

AgCl + e− → Ag + Cl− Ag2CrO4 + 2 e → 2 Ag + CrO −

3+

3+

2− 4

− 3

+

RESOURCE SECTION 3: DATA

571

Reduction half-reaction

E 9/V

Reduction half-reaction

E 9/V

Be2+ + 2 e− → Be

−1.85

Mn2+ + 2 e− → Mn

−1.18

Bi + 3 e → Bi

+0.20

Mn3+ + e− → Mn2+

+1.09

MnO2 + 4 H + 2 e → Mn + 2 H2O

+1.23

BrO− + H2O + 2 e− → Br− + 2 OH−

+0.76

MnO4− + 8 H + + 5 e− → Mn2+ + 4 H2O

+1.51

Ca2+ + 2 e− → Ca

−2.87

MnO4− + e− → MnO42−

3+

Br2 + 2 e → 2 Br −

Cd(OH)2 + 2 e → Cd + 2 OH

+

+1.51 −

2+

+0.56

−0.81

MnO + 2 H2O + 2 e → MnO2 + 4 OH

Cd2+ + 2 e− → Cd

−0.40

Na+ + e− → Na

Ce3+ + 3 e− → Ce

−2.48

Ni2+ + 2 e− → Ni

Ce + e → Ce

+1.61

NiOOH + H2O + e → Ni(OH)2 + OH

Cl2 + 2 e− → 2 Cl−

+1.36

NO3− + 2 H + + e− → NO2 + H2O

ClO− + H2O + 2 e− → Cl− + 2 OH −

+0.89

NO3− + 3 H + + 3 e− → NO + 2 H2O

4+

3+

2− 4

+0.60 −2.71 −0.23

+0.49 +0.80 +0.96

ClO + 2 H + 2 e → ClO + H2O

+1.23

NO + H2O + 2 e → NO + 2 OH

ClO4− + H2O + 2 e− → ClO3− + 2 OH −

+0.36

O2 + 2 H2O + 4 e− → 4 OH −

+0.40

Co2+ + 2 e− → Co

−0.28

O2 + 4 H + + 4 e− → 2 H2O

+1.23

Co + e → Co

+1.81

O2 + e → O

−0.56

Cr 2+ + 2 e− → Cr

−0.91

O2 + H2O + 2 e− → HO2− + OH −

−0.08

Cr2O72− + 14 H+ + 6 e− → 2 Cr3+ + 7 H2O

+1.33

O3 + 2 H+ + 2 e− → O2 + H2O

+2.07

− 4

+

3+

− 3

2+

− 3

− 2

− 2

+0.10

Cr + 3 e → Cr

−0.74

O3 + H2O + 2 e → O2 + 2 OH

Cr 3+ + e− → Cr 2+

−0.41

Pb2+ + 2 e− → Pb

Cs+ + e− → Cs

−2.92

Pb4+ + 2 e− → Pb2+

Cu + e → Cu

+0.52

PbSO4 + 2 e → Pb + SO

−0.36

Cu2+ + 2 e− → Cu

+0.34

Pt 2+ + 2 e− → Pt

+1.20

Cu + e → Cu

+0.16

Pu + e → Pu

+0.97

F2 + 2 e− → 2 F −

+2.87

Ra2+ + 2 e− → Ra

−2.92

Fe2+ + 2 e− → Fe

−0.44

Rb+ + e− → Rb

−2.93

Fe3+ + 3 e− → Fe

−0.04

S + 2 e− → S2−

−0.48

Fe + e → Fe

+0.77

S2O82− + 2 e− → SO42−

+2.05

[Fe(CN)6]3− + e− → [Fe(CN)6]4−

+0.36

Sn2+ + 2 e− → Sn

−0.14

2 H+ + 2 e− → H2

0, by definition

Sn4+ + 2 e− → Sn2+

+0.15

−0.83

Sr 2+ + 2 e− → Sr

−2.89

2 HBrO + 2 H + + 2 e− → Br2 + 2 H2O

+1.60

Ti2+ + 2 e− → Ti

−1.63

2 HClO + 2 H + + 2 e− → Cl2 + 2 H2O

+1.63

Ti3+ + e− → Ti2+

−0.37

H2O2 + 2 H + 2 e → 2 H2O

+1.78

Ti4+ + e− → Ti3+

H4XeO6 + 2 H + + 2 e− → XeO3 + 3 H2O

+3.0

Tl+ + e− → Tl

−0.34

Hg 22+ + 2 e− → 2 Hg

+0.79

U3+ + 3 e− → U

−1.79

+0.27

U4+ + e− → U3+

−0.61

+0.86

V2+ + 2 e− → V

−1.19

+0.92

V +e →V

−0.26

+0.62

Zn2+ + 2 e− → Zn

−0.76

3+

+

2+

+

3+

2+

2 H2O + 2 e → H2 + 2 OH −

+

Hg2Cl2 + 2 e → 2 Hg + 2 Cl −

Hg 2+ + 2 e− → Hg 2 Hg + 2 e → Hg 2+

2+ 2

Hg2SO4 + 2 e− → 2 Hg + SO42−

3+

3+

2+

+1.24 −0.13 +1.67

2− 4

4+

0.00

572

RESOURCE SECTION 3: DATA

Table 3c

Biological standard potentials at 298.15 K in electrochemical order E 9/V

Reduction half-reaction O2 + 4 H + + 4 e− → 2 H2O

+0.81

NO3− + 2 H+ + 2 e− → NO2− + H2O

+0.42

Fe3+(cyt f ) + e− → Fe2+ (cyt f )

+0.36

Cu (plastocyanin) + e → Cu (plastocyanin)

+0.35

Cu2+(azurin) + e− → Cu+(azurin)

+0.30

O2 + 2 H + + 2 e− → H2O2

+0.30

Fe (cyt c551) + e → Fe (cyt c551)

+0.29

2+

3+

+

2+

Fe3+(cyt c) + → Fe2+ (cyt c)

+0.25

Fe3+(cyt b) + e− → Fe2+ (cyt b)

+0.08

Dehydroascorbic acid + 2 H + + 2 e− → ascorbic acid

+0.08

Coenzyme Q + 2 H + + 2 e− → coenzyme QH2

+0.04

Fumarate2− + 2 H + + 2 e− → succinate2−

+0.03

Vitamin K1(ox) + 2 H + 2 e → vitamin K1(red)

−0.05

Oxaloacetate2− + 2 H + + 2 e− → malate2−

−0.17

Pyruvate + 2 H + 2 e → lactate

−0.18

+

+

Ethanal + 2 H + + 2 e− → ethanol

−0.20

Riboflavin(ox) + 2 H + + 2 e− → riboflavin (red)

−0.21

FAD + 2 H + + 2 e− → FADH2

−0.22

Glutathione (ox) + 2 H + + 2 e− → glutathione (red)

−0.23

Lipoic acid (ox) + 2 H + + 2 e− → lipoic acid (red)

−0.29

NAD+ + H + + 2 e− → NADH

−0.32

Cystine + 2 H + + 2 e− → 2 cysteine

−0.34

Acetyl − CoA + 2 H + + 2 e− → ethanal + CoA

−0.41

2H2O + 2 e− → H2 + 2 OH−

−0.42

Ferredoxin (ox) + e− → ferredoxin (red)

−0.43

O2 + e− → O2−

−0.4

Answers to odd-numbered exercises Fundamentals F.13 F.15 F.17 F.19 F.21 F.23 F.25 F.27 F.29 F.31 F.33

−459.67°F 2.52 mmol 4.18 bar 388 K 0.50 m3 26 J 2.3 kJ 31.2 V 3.26 m 0.37 (a) 638 m s−1, 1.26 km s−1, 2.30 km s−1 (b) 319 m s−1, 627 m s−1, 1.15 km s−1

Chapter 1 E1.11 E1.13 E1.15 E1.17 E1.19 E1.21 E1.23 E1.25 E1.27 E1.29 E1.31 E1.33 E1.35

E1.37 E1.39

39 J −1.0 × 102 J – × 102 kJ, 6.1 × 102 s 9.20 −13.0 J (b) 37.1 J K−1 mol−1, 28.8 J K−1 mol−1 b + 2cT bT 2 c (a) DHm(T) = aT + − − 16.1 kJ mol−1 T 2 (a) +1.9 kJ mol−1 (b) +30.6 kJ mol−1 +301 kJ 40.6– kJ mol−1, 37.5– kJ mol−1 −2346– kJ mol−1, −1051– kJ mol−1 15.2 kJ g−1, 34.0 kJ g−1 (a) −24.7– kJ (b) 7.9 m – kJ (c) +39.0 – (d) 12.4 m (a) −1560 kJ mol−1 (b) slightly less efficient (a) −23.47 kJ mol−1 (b) −93.9 kJ mol−1 (c) −2810.44 kJ mol−1 (d) +306.94 kJ mol−1

E1.41 E1.43 E1.45

(a) (i) +552– kJ mol−1 (ii) −2.9 kJ mol−1 – (b) 263.5 K 279 J K−1 mol−1, −2805 kJ mol−1, less exothermic D rU 3 (298 K) + D rC V3 × (T − 298 K)

Chapter 2 E2.9 E2.11 E2.13 E2.15 E2.17 E2.19 E2.23 E2.25 E2.27 E2.29

5.03 kJ K−1 122 J K−1, 130 J K−1, 606 J K−1, 858 J K−1 4.0 × 10−4 J K−1 mol−1 5.11 J K−1 0.95 J K−1 mol−1 (b) +34 kJ K−1 mol−1 537 J K−1 mol−1 −198.72 J K−1, −32.99 kJ 0.41 g 8.1 × 1023 molecules of ATP

Chapter 3 E3.9 E3.11 E3.13

E3.15 E3.21 E3.23 E3.25 E3.27 E3.29 E3.31 E3.33

(a) +2.03 kJ mol−1 (b) +1.50 J mol−1 (a) +1.7 kJ mol−1 (b) −20 kJ mol−1 (a) 2.4 kg (b) 32 kg (c) 2.5 g (d) 135.6 bar (b) 0.758 Pa (b) −20.5 kJ mol−1, −126 kJ mol−1, −372 J K−1 mol−1 (c) 348 K (a) 1.32 dm3 (b) 61.2 kPa 2.41 × 10−3 (a) −1.31 kJ mol−1, spontaneous (b) +4.38 J K−1 mol−1 2.30 kPa (a) 0.056 mg N2, 0.014 mg N2 (b) 0.17 mg N2 (a) 1.36 mmol dm−3 (b) 33.9 mmol dm−3

574

ANSWERS TO ODD-NUMBERED EXERCISES

E3.35 E3.37 E3.39

−0.27°C −0.09°C 4.9 × 103 mol dm−3

Chapter 5 For notational simplicity, we have used both the molality concentration expression aJ = gJbJ/b3, where b3 = 1 mol kg −1 [5.1a], and aJ = gJbJ [5.1b] where bJ is the unitless magnitude of molality. The convention of eqn 5.1b is most often used in calculations of ionic strength, while the convention of eqn 5.1a appears in Nernst equation computations. E5.9 0.90

Chapter 4 E4.9

(a) 2.9 × 10−5 (b) 1.2 × 109 (c) 1.8 × 102

E5.11

g± = (g+g −2)1/3

E5.13

B = 2.01

E5.15

hydrolysis of 1 mol of ATP does supply sufficient Gibbs energy to transport 3 mol of sodium cation and 2 mol of potassium cation

E4.11

−294 kJ mol−1

E4.13

3.5 × 103, 2.3 × 102, 36

E4.15

6.8 kJ mol−1

E4.17 E4.19

−25.1 kJ mol−1 (a) 0 (b) −61 kJ mol−1 (c) +18 kJ mol−1

E5.17

Yes

E5.21

n=2

E4.21

(a) 41% (b) 75%

P4.23

(a) KMb = 2.33 torr = 0.311 kPa, KHb = 34.7 torr = 4.62 kPa

E4.25

(a) −5798 kJ mol−1 (b) (i) −16.5 MJ (ii) −16.9 MJ

E4.27

(a) −13 kJ mol−1 (b) more exergonic

E4.29

−15.0 kJ mol−1, −38.9 J K−1 mol−1

E4.33

(b) (c) (d) (e)

E4.35

(a) 9.60 × 10−3 mol dm−3, 2.02 (b) 0.025 mol dm−3, 12.40 (c) 1.26 (a) 9.1 (b) 4.83 (c) none of the Br− is protonated

E4.39

(a) (b) (c) (d) (e)

E4.41

(a) 6.5 (b) 2.1 (c) 1.5

E4.43

E4.45

E4.49

E4.51

n = 4, +0.56 V, −216– kJ mol−1, 7.3 × 1037

E5.25

41 mV

E5.29

(a) n = 2 (b) n = 1 (c) n = 2

E5.31

(a) +0.94 V (b) Ecell/V = 1.51 − 0.094656 × pH

E5.33

(a) decreases (b) increases (c) decreases (a) −440 kJ mol−1 (b) +29.7 kJ mol−1 (c) −313 kJ mol−1 +1.15 V (a) −444 kJ mol−1, −505 kJ mol−1 (b) −442 kJ mol−1, −442 kJ mol−1 +1.15 V, +0.08 V, +0.82 V, −0.33 V reduced lipoic acid (a) +0.34 V (b) −0.09 V −131.25 kJ mol−1, −167.10 kJ mol−1, + 56.7 J K−1 mol−1 1 mol (a) −27 kJ mol−1 (b) eight

E5.35

14.870 3.67 × 10−8 mol dm−3, 3.67 × 10−8 mol dm−3 7.45 14.870

E4.37

E5.23

E5.37

E5.39 E5.41 E5.43 E5.47 E5.49 E5.51

2.0, 12.0, 0.083 3.9, 10.1, 0.87 5.0, 9.0, 6.1 × 10−5 5.0, 9.0, 6.1 × 10−5 2.6, 11.4, 0.024

Chapter 6 E6.7

1.6 per cent

(a) 6.9 × 10−2 mol dm−3, 1.4 × 10−13 mol dm−3, 8.1 × 10−2 mol dm−3, 6.5 × 10−5 mol dm−3, 0.069 mol dm−3 (b) 1.65 × 10−3 mol dm−3, 2.78

E6.9

(a) 1.5 mol dm−3 s−1, 0.73 mol dm−3 s−1, 1.5 mol dm−3 s−1 (b) mol−2 dm6 s−1

E6.13

0.92 g dm−3 h−1

(a) pI = pH = 12 (pKa1 + pKa2)

E6.17

3.19 × 10−6 Pa−1 s−1

(b) pI = pH = 12 (pKa1 + pKa2) (c) pI = pH = 12 (pKa2 + pKa3)

E6.19

3.67 × 10−3 min−1

E6.21

(a) first (b) 30.27 dm3 mol−1 s−1 1 1 = + 2krt [A]2 [A]02

Kd 1 + Kd (b) K = exp{−(D dH 3 − TDdS 3)/RT } (c) 317 K (d) 9 kJ mol−1 (a)

(a) (i) 0.142 (b) (i) 0.142

(ii) 0.858 (ii) 0.858

(iii) 0.716

E6.23 E6.25

(iv) 0.68

(a) krt = (b)

2x(A0 − x) A20(A0 − 2x)2

2x D A 1 D A A0 − 2x D A + ln C A20(A0 − 2x)F C A20F C A0 − x F

ANSWERS TO ODD-NUMBERED EXERCISES

3067 a 1.44 × 10−9t1/2, 1.64 min (a) 0.043 mol dm−3, 0.138 mol dm−3 (b) 0.0001 mol dm−3, 0.0951 mol dm−3

E8.27

t = x 2/2D or x = (2Dt)1/2 (a) 27 h (b) 2.7 × 103 h (c) 3.0 × 103 a

E8.29

r 2 = 6Dt

E6.35

2n−1 − 1 ( 43 )n−1 − 1 85.6 kJ mol−1, 3.65 × 1011 mol dm−3 s−1

E6.37

Ea = 52 kJ mol−1

E6.39

30.1 kJ mol−1

E8.31 E8.33 E8.35 E8.37

E6.41

47.8 kJ mol−1

0.234 1 × 106 steps 62.3 mm s−1 (a) 1.33 × 10−9 m2 s−1, 184 pm (b) 1 (a) 16.5– nm−1

E6.27 E6.29 E6.31

E6.33

E8.39

Chapter 7 E7.11

k′([A]0 + [B]0) + {kr[A]0 − k′[B]0}e−(kr+k′)t [A] = , kr + k′

E7.17

(kr [A]0 − k′[B]0)(1 + e−(k+k′)t) kr + k′ t −1 = 4kr[A]eq + k r′ 1.7– × 107 s−1, 2.8 × 109 mol−1 dm3 s−1, 1.6– × 102 1.6– × 102

E7.19

first-order in H2O2 and in Br −, second-order overall

E7.21

(ii) Both the pre-equilibrium approximation and the steady-state approximation predict that the reaction is first-order in A, first-order in B, and second-order overall ka[HA][B] kakb[HA]2[B] [A−] = , + + k′[BH ] + kb[HA] k′[BH ] + kb[HA] a a (b−d)×(t−1750 y) , the Malthus model does seem to describe N = N0e the data as an exponential growth, kr = b − d = 0.0095 y−1

[B] = E7.13 E7.15

E7.23 E7.25

E9.11

E9.13

E9.15 E9.17 E9.19 E9.21

E9.23

(a) (i) 0.18 (ii) 0.30 (b) (i) 3.9 × 10−18 (ii) 6.0 × 10−6

E7.29

1.5 × 1015

E7.31

E9.27

E7.33

−3 kJ mol−1 126– kJ mol−1

E7.35

−33.8 J K−1 mol−1, +27.6 kJ mol−1, 37.7 kJ mol−1

E9.31

E7.37

(b) +61.4 kJ mol−1

E7.39

1.08 dm6 mol−2 min−1

kb[E]0[S] , k′a >> kb 1 [S] + K

rate of formation of P = kb[ES] =

E8.13 E8.15

[S] = KM. 2.31 mmol dm−3 s−1, 1.11 mmol dm−3, 1.16– × 102 s−1, 1.0 × 102 mmol−1 dm3 s−1 [S] [S] K (a) + M = v vmax vmax (c) 279 pmol dm−3 s−1, 86.4 mmol dm−3

E8.17

E9.25

E9.33

E8.9

E8.19

(b) 3.0, 0.91 fmol dm−3

E8.21

Sequential mechanism, 5.10 mol s−1 (kg protein)−1, 0.259 mmol dm−3, 0.0189 mol dm−3, 0.0173 mol dm−3

E8.23

Phenylbutyrate ion is a competitive inhibitor of carboxypeptidase. Benzoate ion is an uncompetitive inhibitor of carboxypeptidase.

t = r 2/6D, 1.7 × 10−2 s

Chapter 9

E7.27

Chapter 8

or

575

E9.35 E9.37 E9.39 E9.41

E9.43

(a) 6.6 × 10−19 J, 4.0 × 102 kJ mol−1 (b) 3.3 × 10−20 J, 20 kJ mol−1 (c) 1.3 × 10−33 J, 7.8 × 10−13 kJ mol−1 (a) 6.6 × 10−31 m (b) 6.6 × 10−39 m (c) 99.7 pm (d) 3.5 × 10−36 m (a) 5.35 pm (b) 2.2 × 10−24 m s−1 0.90 nm (a) 1.7—7 × 10−4 (b) 5.9—2 × 10−5 (a) 0.196 (b) 0.609 (c) 0.196 (a) 3.30 × 10−19 J (b) 4.95 × 10−14 s−1 (a) 1.1– × 1010 s−1 (b) 0.24, 4 (a) 5.275 × 10−34 J s, 7.89 × 10−19 J (b) 5.2 × 1014 Hz (a) 6.89 × 1013 s−1 (b) 4.35 mm (a) 6.432 × 1013 s−1 (b) 2146 cm−1 (c) 2099 cm−1, 612C18O = 2094 cm−1, 613C18O = 2046 cm−1 16 orbitals 1 2

(a) (b) (a) (b) (c) (d) (e) (b)

1.9 a0 and 7.1 a0 1.87 a0, 6.61 a0, and 15.5 a0 ang. mom. = 0 ang. mom. = 0 ang. mom. = 6 ħ ang. mom. = 2 ħ ang. mom. = 2 ħ Mg, Ti

Chapter 10 E10.19 N(0.644A − 0.245B) E10.25 C2 and CN are stabilized by anion formation. NO, O2, and F2 are stabilized by cation formation E10.27 (a) g, u, g

576

ANSWERS TO ODD-NUMBERED EXERCISES

E10.29 E10.35 E10.37 E10.39 E10.41

N2 2a + 2b, 4a + 4.48b, lower (b) 1.518b, 8.913 eV (b) 5, 1 square planar arrangement

(c) 48 (d) 54 E12.27 (a) 7 (b) structure 5 is inconsistent with these absorptions. E12.31 (a) 6.54– ns (b) 0.11 ns−1 E12.33 0.4 ns

Chapter 11

E12.35 3.3 × 1018

E11.15 3.40 × 10 kg mol E11.17 31 kg mol−1 E11.19 (a) plot of nbp against t2 is linear (b) 167 ms E11.25 66.1 pm E11.27 (a) 47.9– kJ mol−1 (b) 24 kJ mol−1 (c) 0.60 kJ mol−1 3

−1

E11.29 (a) 1.45 D (b) mortho /mmeta = 3 E11.31 2mO–H cos(f/2) (a) 2.13 D (b) 2 arccos(mH–O–O–H /3.02 D) E11.35 196 pm E11.37 −4.2 × 10−3 J mol−1 R = 21/6 s 24 nm 1.3 × 104 serum albumin and bushy stunt virus resemble solid spheres, but DNA does not E11.51 −0.042 J K−1 mol−1 P11.39 E11.45 E11.47 E11.49

E11.53 (a)

A 0.957 D b = 0.362 C 3.59 F

b0 b1 b2

(b) W = 1.362

Chapter 12 E12.9

0.307 m−1 3.26 m 1.01 × 104 dm3 mol−1 cm−1 0.951% A −r A r A − A2 [B] = 2 A 1 , [A] = B 1 (De2)L (De2)L 99.5 mmol dm−3, 96.3 mmol dm−3 (a) 6.37, 2.12 (b) 1.74 × 106 dm3 mol−1 cm−2 (a) d6 = 53 cm−1. (b) d6 = 0.27 cm−1. (a) 967.0, 515.6, 411.8, 314.2 (b) 3002.2, 2143.7, 1885.8, 1640.2 (a) 3 (b) 4

(a) (b) E12.11 (a) (b) E12.13 E12.15 E12.17 E12.21 E12.23 E12.25

E12.37

1 1 k [Q] = + Q , 5.2 × 106 dm3 mol−1 s−1 Iphos Iabs kphos Iabs

E12.39 3.5 nm E12.41 3 × 103 Req aK = 0 Rmax a0K + 1 (b) Rmax = 1/slope and K = slope/intercept (c) R(t) = Req(1 − e−kobst ), where kobs = kona0 + k off

E12.47 (a)

(d) R(t) = Rmax e−kobst, where kobs = k off E12.49 (b) All modes

Chapter 13 E13.11 5.57 × 10−24 J E13.13 (a) T −1Hz (b) A s kg−1 E13.15 (a) 6.72 × 10−4 (b) 2.47 × 10−3 E13.17 (a) 3.4 × 10−5 (b) 8.6 × 10−6 E13.19 328.5 MHz E13.21 11.74 T E13.23 (a) independent (b) 13 E13.25 (a) 9.5 mT (b) 46 mT E13.27 1:7:21:35:35:21:7:1 E13.33 cos f = B/4C E13.35 [I]0 =

[E]0Dn − KI dv

E13.43 B 3 − B 2 = (334.8 − 332.5)mT = 2.3 mT # a = 2.3 mT B 2 − B 1 = (332.5 − 330.2)mT = 2.3 mT $ hn 9.319 × 109 Hz [13.23] = (7.14478 × 10−11 T Hz−1) × mBB 0 332.5 × 10−3 T – = 2.0025 E13.45 (a) 331.9 mT (b) 1.201 T E13.51 (b) seven lines separated by 12 × (0.675 mT), 1:2:3:4:3:2:1 g=

Index of Tables Fundamentals F.1

Pressure units and conversion factors

6

F.2

The gas constant in various units

8

F.3

A summary of standard conditions

10

Biochemical thermodynamics 1.1

Heat capacities of selected substances

33

1.2

Standard enthalpies of transition at the transition temperature

47

1.3

Selected bond enthalpies, H(A−B)/(kJ mol−1)

50 51

1.4

Mean bond enthalpies, ΔHB/(kJ mol )

1.5

Standard enthalpies of combustion

53

1.6

Thermochemical properties of some fuels

54

1.7

Reference states of some elements at 298.15 K

59

1.8

Standard enthalpies of formation at 298.15 K

60

2.1

Entropies of vaporization at 1 atm and the normal boiling point

76

2.2

Standard molar entropies of some substances at 298.15 K

3.1

Critical constants

3.2

Henry’s law constants for gases dissolved in water at 25°C

115

3.3

Activities and standard states

119

3.4

Cryoscopic and ebulioscopic constants

123

4.1

Thermodynamic criteria of stability

142

4.2

Standard Gibbs energies of formation at 298.15 K

148

4.3

Transfer potentials at 298.15 K

153

4.4

Acidity and basicity constants at 298.15 K

160

4.5

Successive acidity constants of polyprotic acids at 298.15 K

165

4.6

Acidity constants of amino acids at 298.15 K

167

5.1

Standard potentials at 25°C

199

5.2

Biological standard potentials at 25°C

201

−1

79 104

The kinetics of life processes 6.1

Kinetic techniques

220

7.1

Collision cross-sections of atoms and molecules

261

8.1

Diffusion coefficients in water, D/(10−9 m2 s−1)

8.2

−8

2 −1

286 −1

Ionic mobilities in water at 298 K, u/(10 m s V )

291

Biochemical structure 9.1

Atomic radii of main-group elements, r/pm

352

9.2

Ionic radii of selected main-group elements

353

9.3

First ionization energies of main-group elements, I/eV

354

578

INDEX OF TABLES

9.4

Electron affinities of main-group elements, Eea/eV

355

10.1

Hybrid orbitals

371

10.2

Electronegativities of the main-group elements

384

10.3

Summary of ab initio calculations and spectroscopic data for four linear polyenes

402

11.1

The essential symmetries of the seven crystal systems

416

11.2

Partial charges in polypeptides

425

11.3

Dipole moments and mean polarizability volumes

426

11.4

Interaction potential energies

435

11.5

Lennard-Jones parameters for the (12,6) potential

436

11.6

Radii of gyration of biological macromolecules and assemblies

441

11.7

Relative frequencies of amino acid residues in helices and sheets

446

11.8

Variation of micelle shape with the surfactant parameter

449

11.9

Gibbs energies of transfer of amino acid residues in a helix from the interior of a membrane to water

450

Biochemical spectroscopy 12.1

Typical vibrational wavenumbers

481

12.2

Typical vibrational wavenumbers for the amide I and II bands in polypeptides

482

12.3

Color, frequency, and energy of light

485

12.4

Electronic absorption properties of amino acids, purine, and pyrimidine bases in water at pH = 7

488

12.5

Values of R0 for some donor−acceptor pairs

500

13.1

Nuclear constitution and the nuclear spin quantum number

515

13.2

Nuclear spin properties

515

Index 3D QSAR, 454 90º pulse, 528

A A and B forms (DNA), 447 Ab initio method, 398 Aberration, 501 Absolute zero, 7, 27 Absorbance, 466 Absorption spectroscopy, 463 Abundant-spin species, 532 Acceleration, 10 Acceleration of free fall, 12 Acetic acid, 2 Acetylene, VB description, 370 Acid, 156 Acid buffer, 171 Acidity constant, 158 electrochemical measurement, 199 Acidosis, 173 Actinoid, 351 Action potential, 188, 189 Activated complex, 238, 261 Activated complex theory, 261 Activation barrier, 259 Activation energy, 236 interpretation, 260 Activation Gibbs energy, 262 Activation-controlled limit, 257 Active site, 273 Active transport, 187, 285 Activity, 118 summary, 119, 137 Activity coefficient, 119 electrolyte solution, 182 mean, 183 Adiabatic, 25 Adiabatic bomb calorimeter, 42 Adiabatic flame calorimeter, 44 ADP, 208 hydrolysis, 140 AEDANS (1.5-I-AEDANS), 501 Aerobic metabolism, 153 AFM, 329, 436 AIDS, 437 Airy radius, 484 Alkalosis, 173 Allosteric effect, 145, 398 Allosteric enzyme, 306

Allowed transition, 470 Alpha electron, 347 Alpha helix, 442 Alveoli, 118 Alzheimer’s disease, 386 AM1, 399 Amide I and II bands, 482 Amide III region, 489 Amino acid solution speciation, 169 Amount of substance, 5 Ampere, 42 Amphipathic, 86, 449 Amphiprotic anion pH calculation, 176 Amphiprotic species, 169 Amphoteric anion, 169 Amphoteric substance, 169 Amplitude, 12 Amyloid plaque, 446 Amylotrophic lateral sclerosis, 386 Anabolism, 28 Anaerobic metabolism, 153 Ångström, 558 Angular momentum, 331 quantization, 333, 335 Angular wavefunction, 342 Anharmonic vibration, 477 Anion, 1 Anion configuration, 351 Anode, 192 Antibonding orbital, 375, 385 Antifreeze, 124 Antioxidant, 214, 384 Antiparallel beta sheet, 445 Anti-Stokes radiation, 464 Antisymmetric stretch, 479 Approximation Born–Oppenheimer, 364 Hückel, 388 orbital, 346 steady-state, 252 Aquatic life, 116 Arctic fish, 124 Arrhenius equation, 236 Arrhenius parameters, 236 interpretation, 237 Arrhenius, S., 235 Artist’s color wheel, 485 Ascorbic acid, 384

Atmosphere, 6, 558 ionic, 184 reactions in, 494 temperature profile, 494 Atmospheric pressure, 6 Atomic force microscopy, 329, 436 Atomic number, 1 Atomic orbital, 340 Atomic radius, 352 Atomic structure, 1 ATP, 28, 208 biosynthesis, 152 Aufbau principle, 349 Austin Method 1, 399 Autocatalysis, 308 Autoionization, 157 Autoionization constant, 159 Autoprotolysis contribution to pH, 175 Autoprotolysis constant, 159 Autoprotolysis equilibrium, 157 Average rate, 221 Average speed, 16 Avogadro’s constant, 6 AX spectrum, 522 AX2 spectrum, 523 AX3 spectrum, 523 Azimuthal quantum number, 340

B Bar, 6, 558 Base, 156 Base buffer, 171 Base pair, 107 Base stacking, 447 Basicity constant, 158 Beam combiner, 465 Beer’s law, 220, 466 Beer–Lambert law, 466 Bending mode, 479 Bends (diving), 133 Benzene elpot surface, 401 isodensity surface, 400 MO description, 390 VB description, 372 Benzene radical anion, 537 Beta barrel, 445 Beta electron, 347

580

INDEX

Beta sheet, 442 Beta-blocker, 234 Bilayer, 3, 450 Bimolecular reaction, 248 Binary mixture, 130 Biochemical cascade, 503 Bioenergetics, 23 Biological membrane phase transition, 108 Biological standard potential, 198, 199 Biological standard state, 139 Biopolymer crystallization, 421 melting temperature, 107, 180 Biosensor analysis, 473 Biradical, 383 Blood, buffer action, 173 Bohr effect, 174 Bohr frequency condition, 314, 464 Bohr magneton, 514 Bohr radius, 342 Bohr, N., 342 Boiling, 103 Boiling point, elevation of, 123 Boiling temperature, 103 Boiling-point constant, 123 Boltzmann distribution, 13, 15, 27 chemical equilibrium, 144 spectroscopic intensity, 471 spin states, 514 Boltzmann formula, 80 Boltzmann’s constant, 15 Bomb calorimeter, 42 Bond, 1 covalent, 364 ionic, 364 Bond enthalpy, 50 Bond length, 365 Bond order, 382 Bonding orbital, 374, 385 Born interpretation, 320 Born, M., 320 Born–Oppenheimer approximation, 364 Boson, 359 Boundary condition, 230, 319 Boundary surface, 342 Bovine serum albumin, 293, 411 Bragg, W. and L., 415 Bragg’s law, 419 Bravais lattice, 416 Breathing, 117 Bremsstrahlung, 415 Brønsted–Lowry theory, 156 BSA, 293, 411 Buffer action, 171 blood, 173 Buffer solution, 170 Building-up principle, 349 Bulk matter, 4 Buoyancy correction, 408 Butadiene, 391

C Cage effect, 256 Calorimeter, 42 Calorimeter constant, 42 Calorimetry, 42 Camping gas, 67 Candela, 557 Capillary electrophoresis, 293 Carbohydrate, food, 56 Carbon hybridization, 369 role in biochemistry, 391 Carbon dioxide normal modes, 480 polarity, 427 vibration, 476 Carbonic acid, 165 Carbonic anhydrase, 83, 356 Carotene, 401 electronic structure, 327 Cartesian coordinates, 376 Catabolism, 28 Catalase, 383 Catalyst effect on activation energy, 238 effect on equilibrium, 150 Catalytic antibody, 284 Catalytic constant, 279 Catalytic efficiency, 279, 280 Catalytic triad, 284 Cathode, 192 Cation, 1 Cation configuration, 351 CBG, 455 CCD, 466 ccDNA, 447 Cell death, 484 Cell notation, 194 Cell potential, 196 equilibrium constant, 203 thermodynamic function determination, 202 variation with pH, 199 variation with temperature, 206 Celsius scale, 7 Cesium atoms STM, 330 Cetyl trimethylammonium bromide, 449 Chain rule, 37 Channel former, 188 Charge balance, 175 Charge-coupled device, 466 Charge–dipole interaction, 429 Charge–transfer transition, 487 Chemical equilibrium approach to, 245 effect of temperature, 150 Gibbs energy, 135 molecular interpretation, 144, 151 thermodynamic criterion, 137 Chemical exchange, 526 Chemical potential activity, 118 gas, 112 introduced, 111

ion in solution, 183 reacting species, 137 solute, 116 solution, 125 solvent, 114 uniformity of, 111 variation with pressure, 112 Chemical quench flow method, 221 Chemical reactivity, 402 Chemical shift, 519 Chemisosmotic theory, 209 Chiral, 488 Chlorophyll, 210, 485, 486, 503 spectrum, 314 Chloroplast, 209, 503 Chromatic aberration, 501 Chromophore, 487 Chymotrypsin, 276, 284 Circular dichroism, 488 Circular polarization, 488 Citric acid cycle, 155 Clamp, 294 Clapeyron equation, 101 Classical mechanics, 10 Classical physics, 313 Classical thermodynamics, 23 Clausius–Clapeyron equation, 102, 132 Climate change, 476 Closed shell, 347 Closed system, 24 CMC, 449 CNDO, 399 Coefficient activity, 119, 182 diffusion, 286, 287 Einstein, 470 frictional, 408 Hill, 177 integrated absorption, 468 interaction, 306 mean activity, 183 molar absorption, 220, 466 partition, 289 transmission, 262 viscosity, 288 Cold denaturation, 246 Cold pack, 25 Collagen, 445 Colligative property, 123 thermodynamic origin, 124 Collision, 267 Collision frequency, 259 Collision theory, 259 Collisional deactivation, 472, 499 Color wheel, 485 Combustion, 53 Common logarithm, 182 Competitive inhibition, 281 Complementary color, 485 Complete neglect of differential overlap, 399 Complete shell, 347 Composition at equilibrium, 143 Computational chemistry, 61 Computational technique, 398

INDEX

Concentration determination of, 220, 468 measure of, 130 Concentration gradient, 409 Condensation, 48 Configuration atom, 347 ion, 351 macromolecule, 438 system, 26 weight of, 80 Confocal Raman microscopy, 511 Conformation, 438 Conformational conversion, 526 Conformational energy, 442 Conformational entropy, 441 Conjugate acid, 157 Conjugate base, 157 Conjugated molecule, 327 Conjugation, 327 Connectivity, 2 Consecutive reaction, 249 Conservation of energy, 12, 24 Constant acidity, 158, 199 autoionization, 159 autoprotolysis, 159 Avogadro’s, 6 basicity, 158 boiling point, 123 Boltzmann’s, 15 calorimeter, 42 catalytic, 279 cryoscopic, 123 diffusion, 286, 408 dissociation (acid and base), 158 ebullioscopic, 123 equilibrium, 140 Faraday’s, 186 force, 474 freezing point, 123 gas, 8 gravitational, 19 Henry’s law, 115 hydrophobic, 87 hyperfine coupling, 538 ionization (acid and base), 158 Michaelis, 274 normalization, 324 Planck’s, 13, 314 rate, 223 sedimentation, 408 spin–spin coupling, 521 Constant-current mode, 329 Constant-force mode, 329 Constant-z mode, 329 Contact interaction, 525 Contact mode, 329 Continuous wave spectrometer, 520 Contour length, 441 Contrast agent, 531 Convection, 286 Conventional temperature, 46 Cooperative binding, 145, 398 Cooperative process, 107

Corey, R., 442 Corey–Pauling rules, 442 Cornea, 501 Correlation spectroscopy, 534 Corticosteroid-binding globulin, 455 Cosmic rays, 13 COSY, 534 Coulomb, 11, 42 Coulomb integral, 388 Coulomb potential, 18 Coulomb potential energy, 11, 338, 425 Counter ion, 184 Coupled reactions, 151 Covalent bond, 1, 364 Creatine phosphate, 153 Critical micelle concentration, 449 Critical point, 104 Critical pressure, 104 Critical temperature, 104 Crixivan, 438 Cross peaks, 534 Cryoscopic constant, 123 Crystal field theory, 392 Crystal plane, 416 Crystal system, 415 Crystal-field splitting, 393 Crystallization, 110 Crystallization, 421 CTAB, 449 Cubic system, 416 Curvature, 287 CW spectrometer, 520 CW-EPR, 536 Cyclic boundary conditions, 332 Cytochrome, 208 Cytosol, 153

D d Block, 351 d Electron, 341 d Orbital, 345 d Subshell, 341 Dalton, 410 Dalton’s law, 10 Daniell cell, 193 Dansyl chloride, 509 Davisson, C., 315 Davisson–Germer experiment, 315 d–d Transition, 487 De Broglie relation, 316 De Broglie, L., 316 Deactivation, 496, 499 Debye, 426 3 Debye T -law, 80 Debye, P., 184, 422 Debye–Hückel limiting law, 185 Debye–Hückel theory, 184 Debye–Scherrer technique, 422 Decay, 490, 528 Definite integral, 97 Degeneracy, 331 Delocalization energy, 391 Delocalized orbital, 391 Delta orbital, 378

Delta scale, 519 Denaturation, 45 Density functional theory, 399 Deoxygenated heme, 394 Deoxyribose, 3 Depression of freezing point, 123 Deprotonation, 158 Derivative, 37 Deshielded, 519 Detector, 465 Determinant, 389 Deuteration, effect of, 363 DFT, 399 Diagonal peaks, 534 Dialysis, 126, 134, 422 Diamagnetic species, 383 Diathermic, 25 Dielectric constant, 425 Diethyl ether, NMR spectrum, 524 Differential equation, 230, 244 Differential overlap, 399 Differential scanning calorimeter, 44 Differentiation, 37 Diffraction, 315, 415 Diffraction grating, 465 Diffraction pattern, 415 Diffractometer, 422 Diffusion, 285, 304 across membranes, 288 Diffusion coefficient, 286, 408 variation with temperature, 287 Diffusion equation, 286, 305 Diffusion-controlled limit, 257 Dihelium, 380 Dihydroxypropanone phosphate, 154 Dilute-spin species, 532 Dipole, 426 Dipole moment calculation, 428 formaldehyde, 429 induced, 431 magnitude, 428 peptide group, 428 transition, 469 Dipole–dipole interaction, 430 Dipole–induced-dipole interaction, 432 Disease, 386 Dismutation, 205 Dispersal in disorder, 71 Dispersion interaction, 432 Disproportionation, 205 Dissociation, 490 Dissociation constant, 158 Dissociation limit, 490 Dissolving, thermodynamics of, 121 Distribution, end-to-end, 440 Disulfide link, 445 d-Metal complex, 392 DNA A, B, and Z forms, 447 closed-circular, 447 freely jointed chain, 68 melting temperature, 107 stability, 133 STM image, 330

581

582

INDEX

structure from X-rays, 423 supercoiled, 447 UV damage, 504 X-ray pattern, 415 Dobson unit, 508 Donnan equilibrium, 119 Double bond, VB description, 370 Drift speed, 290 Drift velocity, 290 DSC, 44 DU, 508 Duality, 315, 316 Dynamic equilibrium, 100 Dynamic light scattering, 414

E Eadie–Hofstee plot, 306 Ebullioscopic constant, 123 Effect allosteric, 145, 398 Bohr, 174 cage, 256 kinetic isotope, 363 kinetic salt, 264 nuclear Overhauser, 532 photoelectric, 315 relativistic, 360 Effective concentration, 118 Effective mass, 362, 474 Effective nuclear charge, 348 Effective rate constant, 226 Effector, 306 Effector molecule, 188 Efficiency, 500 eg Orbital, 393 EHT, 399 Einstein coefficient spontaneous emission, 471 stimulated absorption, 470 stimulated emission, 471 Einstein relation, 307 Einstein, A., 315 Electric charge, interaction of, 11 Electric dipole, 426 Electric dipole moment, 426 Electric heating, 42 Electric permittivity, 425 Electrical work, 195 Electrochemical cell, 189 Electrochemical properties, 401 Electrochemical series, 207 Electrode, 192 Electrode compartment, 192 Electrode concentration cell, 193 Electrolyte concentration cell, 193 Electrolyte solution, 110 activity coefficient, 182 Electrolytic cell, 192 Electromagnetic radiation, 12 Electromagnetic spectrum, 12 Electromotive force, see cell potential, 196 Electron affinity, 355 Electron cryomicroscopy, 318

Electron diffraction, 315 Electron microscopy, 317 Electron pair formation, MO theory, 379 Electron paramagnetic resonance, 513, 536 Electron spin resonance, see Electron paramagnetic resonance, 513 Electron transfer, 296, 499 Electron transfer reaction, 208 Electronegativity, 384 dipole moment, 426 Electron-gain enthalpy, 355 Electronvolt, 558 Electrophoresis, 291 Electrospray ionization, 410 Electrostatic potential surface, 400 Elementary reaction, 247 Elevation of boiling point, 123 Elpot surface, 400 EMF, see cell potential, 196 Emission spectroscopy, 463 Enantiomer, 488 Encounter pair, 256 Endergonic, 149 Endergonic reaction, 93 Endocytosis, 450 Endothermic, 25 Endothermic compound, 61 Endothermic process, 40 End-to-end separation, 440 Energy conservation of, 12 flow in organisms, 28 thermal, 15 zero-point, 326 Energy levels FEMO theory, 405 harmonic oscillator, 336, 474 hydrogenic atom, 338 particle in a 2D box, 330 particle in a box, 325 particle on a ring, 331 particle on a sphere, 334 Energy transfer, 500 Engrailed homeodomain protein, 255 enhancement factor, 534 En-HD, 255 Enthalpy, 39 activation, 263 dissolving, 122 electron-gain, 355 heat transfer at constant pressure, 40 internal energy change relation, 53 state function, 39 temperature dependence, 41 Enthalpy change composite process, 49 reverse process, 49 Enthalpy density, 54 Enthalpy of reaction, combination, 57 Entropy Entropy, 71 activation, 263 and life, 85 at T = 0, 78 Boltzmann formula, 80

close to T = 0, 80 conformational, 441 determination of, 74, 78 dissolving, 122 fusion, 75 molecular interpretation, 80 phase transition, 75 random coil, 441 residual, 82 state function, 73 Third Law, 78 units, 73 vaporization, 76 Entropy change definition, 72 heating, 73 surroundings, 77 total, 84 Enzyme inhibition, 280 Enzyme kinetics, 273 Epifluorescence microscope, 493 EPR, 513, 536 EPR spectrometer, 536 Equation Arrhenius, 236 Clapeyron, 101 Clausius–Clapeyron, 102, 132 differential, 230, 244 diffusion, 286, 305 Einstein, 307 Eyring, 262 Goldman, 188 Henderson–Hasselbalch, 171 Karplus, 524 Kohn–Sham, 399 McConnell, 539 Michaelis–Menten, 274 Nernst, 197 quadratic, 162 Scatchard, 134 Schrödinger, 319, 358 secular, 388 simultaneous, 389 Stern–Volmer, 497 Stokes–Einstein, 409 Stokes–Einstein relation, 288 thermochemical, 47 van ‘t Hoff (equilibrium), 150 van ‘t Hoff (osmosis), 125 Equation of state, 8 Equilibrium dynamic, 100 mechanical, 6 thermal, 6 see also chemical equilibrium Equilibrium bond length, 365 Equilibrium composition, 143 Equilibrium constant cell potential, 203 defined, 140 relation to rate constants, 243 significance, 142 standard Gibbs energy, 141 Equipartition theorem, 68 Equivalence of heat and work, 34

INDEX

ESR, see EPR, 513 Essential symmetry, 416 Ethanol, NMR spectrum, 520 Ethene, 2 MO description, 387 VB description, 370 Ethylene, see ethene, 370 Ethyne, VB description, 370 Evaporation, 47 Exciton coupling, 487 Exclusion rule, 481 Exergonic, 149 Exergonic reaction, 93 Exothermic, 25 Exothermic compound, 61 Exothermic process, 40 Expansion work, 30 Exponential decay, 229 Exponential function, 14, 182 Extended Debye–Hückel law, 186 Extended Hückel theory, 399 Extensive property, 7 Eyring equation, 262

F f Block, 351 f Subshell, 341 Facilitated transport, 289 Fractional composition, lysine, 165 Factorial, 124 FAD, 155, 208 Fahrenheit scale, 18 Far infrared, 13, 465 Faraday’s constant, 186 Far-field confocal microscopy, 493 Fat, 55 FDP, 154 FEMO theory, 405 Femtosecond observations, 264 Fermi contact interaction, 525 Fermion, 359 Ferredoxin, 210 Fick’s first law, 286 Fick’s law, 304 FID, 528 Fine structure NMR, 521 origin, 524 Fingerprint region, 481 First ionization energy, 353 First law Fick’s, 286, 304 thermodynamics, 38 First-order rate law half-life, 230 integrated, 229 Flash photolysis, 221 Flow method, 220 Fluid mosaic model, 450 Fluorescence, 490 quantum yield, 496 Fluorescence lifetime, 496 Fluorescence microscopy, 492 Fluorescence quenching, 497

Fluorescence resonance energy transfer, 500, 501 Flux, 286, 304 fMRI, 531 Food, thermochemical properties, 55 Forbidden transition, 470 Force between molecules, 436 Force constant, 474 Formaldehyde, dipole moment, 429 Förster efficiency, 500 Förster mechanism, 504 Förster theory, 500 Förster, T., 500 Four-circle diffractometer, 423 Four-helix bundle, 446 Fourier synthesis, 420 Fourier transform spectroscopy, 465 Fourier-transform NMR, 527 Fraction deprotonated, 163 Fractional composition amino acid, 169 histidine, 168 Fractional saturation, 144 Framework model, 255 Franck–Condon principle, 486 Free energy, see Gibbs energy, 84 Free expansion, 31 Free-induction decay, 528 Freely jointed chain, 68, 440 Freeze quench method, 221 Freezing, 48 Freezing point, depression of, 123 Freezing temperature, 104 Freezing-point constant, 123 Frequency, 12, 464 FRET, 501 Frictional coefficient, 408 Fructose-6-phosphate, 136 FT-EPR, 536 FT-NMR, 527 Fuel cell, 192 Fuel, thermochemical properties, 52 Functional, 399 Functional MRI, 531 Fusion entropy of, 75 standard enthalpy, 48 Fusion, 48

G g,u Symmetry, 375 Galvanic cell, 192 Gamma-ray region, 13 Gas, 4, 438 Gas constant, 8 Gas electrode, 194 Gas exchange in lung, 118 Gas solubility, 117 Gaussian function, 14 Gaussian-type orbital, 400 Gel electrophoresis, 291 Gerade symmetry, 375 Germer, L., 315 GFP, 493

Gibbs energy activation, 262, 299 cell potential, 196 chemical equilibrium, 135 chemiosmotic theory, 209 defined, 84 dissolving, 121 electrical work, 195 ion transport, 186 partial molar, 110 perfect gas, 97 phase transition, 94 variation with pressure, 95 variation with temperature, 98 work, 88 Glancing angle, 419 Glass electrode, 202 Global minimum, 451 Globar, 465 Glucopyranose, 3 Glucose, 3 alpha and beta, 448 Glucose oxidation, 153, 208 Glucose-6-phosphate, 136, 154 Glutamate ion, 63 Glutamine, 63 Glutathione peroxidise, 384 Glyceraldehyde-3-phosphate, 154 Glycine, 61 Glycogen, 449 Glycol, 124 Glycolysis, 153, 252 Glycoside chain, 448 Glycosidic bond, 448 GMP, 503 Goldman equation, 188 Gramicidin A, 188 Graph, drawing without units, 128 Grating, 465 Gravitational constant, 19 Gravitational potential energy, 12 Green fluorescent protein, 493 Greenhouse gas, 476 Gross energy content, 43 Gross selection rule, 470 Grotthus mechanism, 291 GTO, 400 GTP synthesis, 155 Gunn oscillator, 536 g-Value, 514, 537 Gyromagnetic ratio, see magnetogyric ratio

H Half-life, 230 Half-reaction, 190 reaction quotient, 191 Halley’s comet, 106 Hamburger, 56 Hamiltonian, 319 Hanes plot, 306 Harmonic oscillator, 335, 474 Heat, 25 measurement of, 32 molecular interpretation, 26

583

584

INDEX

Heat capacity, 33 constant pressure, 33, 41 constant volume, 33, 36 molecular interpretation, 34 perfect gas difference, 41 variation with temperature, 66 Heating, 25 Heisenberg, W., 321 Helix, characteristic diffraction pattern, 424 Helix–coil transition, 254 Hematoporphyrin, 506 Heme, 394 Heme group, 145, 446 Hemoglobin, 145, 394, 446 oxygen binding, 136, 144, 397 Hen white lysozyme, 67 Henderson–Hasselbalch equation, 171 Henry, W., 115 Henry’s law, 115 Henry’s law constant, 115 Hess’s law, 57 Heteronuclear diatomic molecules, 385 Hexagonal system, 416 High energy phosphate bond, 152 Highest occupied molecular orbital, 386 High-field end, 520 High-spin complex, 393 Hill coefficient, 177 HIV-AIDS, 437 Homeostasis, 57, 173 hom*o, 386, 401 hom*ogeneous mixture, 110 hom*o–LUMO energy gap, 401 hom*onuclear diatomic molecule, 375 hom*onuclear diatomic molecules, bonding, 381 Hooke’s law, 68, 335 Host–guest complex, 437 Hückel approximation, 388 Hückel, E., 184 Hull, A., 422 Hund’s rule, 350 Hybrid orbital, 369 Hybridization, 368 carbon atom, 369 variation with angle, 371 Hydrodynamic radius, 290 Hydrogen atom, 337 Hydrogen bond, 1, 433 Hydrogen electrode, 194 Hydrogen molecule, MO description, 379 Hydrogen molecule-ion, 373 Hydrogenic atom, 337 Hydrolysis reaction, 153 Hydrolytic enzyme action, 284 Hydronium ion, 156 Hydrophobic interaction, 86 Hydrophobicity constant, 87 Hyperbaric oxygen chamber, 118 Hyperfine coupling constant, 538 Hyperfine structure, 538

I Ice phases, 106 residual entropy, 82 structure, 106, 440 ICP, 489 Ideal gas, see perfect gas Ideal solution, 113 Ideal–dilute solution, 115 Immunoglobulin, 489 Incident circularly polarized technique, 489 Indefinite integral, 97 INDO, 399 Indole, 361 Induced dipole moment, 431 Induced fit model, 273 Induction period, 252 Infectious disease, 308 Infrared active, 476 Infrared spectroscopy, 476 Infrared transition, 480 Inhibition, 280 Initial condition, 230 Initial rate technique, 226 Instant cold pack, 25 Instantaneous rate, 222 Integral, 97 Integral protein, 450 Integrated absorption (NMR), 520 Integrated absorption coefficient, 468 Integrated rate law, 228 Integration, 97 partial fractions, 233 Intensity, NMR and EPR transitions, 515 Intensive property, 7 Interaction coefficient, 306 Interference, 316, 377 Interferometer, 465 Intermediate, 249 Intermediate neglect of differential overlap, 399 Intermolecular interaction, 435 Internal energy, 35 constant volume heat transfer, 36 perfect gas, 35 state function, 37 Internationial System of Units, 5 Intersystem crossing, 492 Inversion symmetry, 375 Ion channel, 188, 295 Ion pump, 188, 295 Ion transfer, 186 Ion transport, 181 Ionic atmosphere, 184 Ionic bond, 1, 364 Ionic mobility, 290 Ionic radius, 353 Ionic strength, 185 Ionization constant, acid and base, 158 Ionization energy, 339, 353 Ion-selective electrode, 202 Isobaric calorimeter, 44 Isochore, 150 Isodensity surface, 400 Isoelectronic, 353

Isolated system, 24 Isolation method, 225 Isoleucine, COSY, 535 Isomorphous replacement, 421 Iso-octane, 55 Isosbestic point, 469 Isosbestic wavelength, 469 Isothermal expansion, 35 Isotonic solution, 126 Isotope, 1 Isotope labeling, NOESY, 535 Isotope substitution, 475 Isotopolog, 475

J Jablonski diagram, 490 Joule, 11, 557 Joule, J., 11

K K shell, 341 Karplus equation, 524 Kekulé structure, 372 Kelvin, 7 Kelvin scale, 7, 105 Kinetic control, 258 Kinetic energy, 11 Kinetic isotope effect, 363 Kinetic model of gases, 16 Kinetic molecular theory, 267 Kinetic salt effect, 264 Kirchhoff ’s law, 62 Klystron, 536 KMT, 16, 267 Kohn–Sham equation, 399 Krafft temperature, 449

L L shell, 341 Lactate ion, 155 Lanthanide contraction, 353 Large calorie, 43 Larmor precession frequency, 527 Laser, 471 Latent heat, 49 Laue method, 422 Law Beer–Lambert, 466 Beer’s, 220, 466 Bragg’s, 419 conservation of energy, 12, 24 Dalton’s, 10 3 Debye T , 80 Debye–Hückel limiting, 185 extended Debye–Hückel, 186 Fick’s, 286, 304 First, of thermodynamics, 38 Henry’s, 115 Hess’s, 57 Hooke’s, 68, 335 Kirchhoff ’s, 62 limiting, 8

INDEX

Raoult’s, 113 rate, 223 Second, of thermodynamics, 71 Stokes’s, 290 Third, of thermodynamics, 78 LCAO, 373 LCAO-MO description, 374 Le Chatelier’s principle, 150 Lead compound, 453 Leaflet of bilayer, 3 Lennard-Jones (12.6) potential, 436 Lewis structure acetic acid, 2 ethene, 2 retinal, 2 water, 2 Life, and Second Law, 85 Lifetime, 472 Lifetime broadening, 472 Ligand field theory, 392, 394 Ligand-gated channel, 188 Light, polarized, 488 Light scattering, 412, 464 Light-harvesting complex, 503 Limiting law, 8, 185 Linear combination of atomic orbitals, 373 Lineweaver–Burk plot, 275 Linewidth, 472 Lipid, 3 Lipid bilayer, 3 melting, 108 Lipid raft model, 450 Liposome, 449 Liquid, 4 Liquid crystal, 108 Liquid junction, 192 Liquid junction potential, 193 Liter, 558 Local contribution, 520 Local magnetic field, 518 Local minimum, 451 Lock-and-key model, 273 Logarithm, 182 London formula, 433 London interaction, 432 Lone pair, 2 Long period, 351 Long-range order, 439 Lou Gehrig’s disease, 386 Lowest unoccupied molecular orbital, 386 Low-field end, 520 Low-spin complex, 393 Lumiflavin, 443 Luminous intensity, 557 LUMO, 386, 401 Lung, 118 Lysozyme, 67 spectrum, 483

M M shell, 341 Macular pigment, 502 Magnetic quantum number, 335, 340 Magnetic resonance, 513

Magnetic resonance imaging, 513, 530 Magnetization, 527 Magnetogyric ratio, 514 MALDI, 410 MALDI–TOF mass spectrometry, 410 Many-electron atom, 337, 346 Marcus cross-relation, 301 Marcus theory, 298, 499 Mass, 5 Mass number, 1 Mass spectrometry, 410 Mass-to-charge ratio, 411 Material balance, 175 Matrix-assisted laser desorption/ionization, 410 Matter wave, 316 Maximum velocity, 274 Maximum work, 31 Maxwell distribution of speeds, 16, 260, 267 Maxwell–Boltzmann distribution, 16 McConnell equation, 539 MCT detector, 466 Mean activity coefficient, 183 Mean bond enthalpy, 51 Mean free path, 268 Mean speed, 16 Mechanical equilibrium, 6 Mechanism of reaction, 224 Melting, 48 thermodynamic basis, 99 Melting temperature, 104 biopolymer, 45, 107, 180 polypeptide, 132 Membrane potential, 187 Mesopause, 494 Mesosphere, 494 Metabolic acidosis, 173 Metabolic alkalosis, 173 Metabolism, 27 Metarhodopsin II, 503 Methanal, see formaldehyde Methylcyclohexane, 61 Micelle, 449 Michaelis constant, 274 Michaelis–Menten equation, 274 Michaelis–Menten mechanism, 274 Michelson interferometer, 465 Microscopy atomic force, 329, 436 confocal Raman, 511 electron, 317 far-field confocal, 493 fluorescence, 492 near-field scanning optical, 493 Raman, 484 scanning electron, 318 scanning tunneling, 329 scanning-probe, 329 vibrational, 483 Microstate, 80 Microwave region, 13 Miller indices, 417 MINDO, 399 Mitchell, P., 209 Mitochondrion, 209 Mixed inhibition, 281

585

Mixture, 110 MNDO, 399 MO theory, 364, 373 Mobility, 290 Model fluid mosaic, 450 framework, 255 induced-fit, 273 KMT, 16 lipid raft, 450 lock-and-key, 273 nucleation–condensation, 255 SIR, 308 VSEPR, 364 Modified neglect of differential overlap, 399 Molality, 131 Molar absorption coefficient, 220, 466 Molar concentration, 131 Molar enthalpy, 39 Molar heat capacity, 33 Molar internal energy, 35 Molar mass determination, 292 macromolecule, 408 osmometry, 128 Molar volume, 9 Molarity, 131 Mole, 5 Mole fraction, 130 Molecular collision, 267 Molecular descriptor, 454 Molecular dynamics, 451 Molecular interpretation chemical equilibrium, 144, 151 entropy, 80 heat, 26 heat capacity, 34 temperature, 26 work, 26 Molecular mechanics, 451 Molecular motor, 296 Molecular orbital, 373 Molecular orbital theory, 364, 373 Molecular potential energy curve, 365 Molecular recognition, 437 Molecularity, 248 Molten globule phase, 109 Moment of inertia, 332 Momentum, 316 angular, 331 linear, 316 Monochromator, 465 Monoclinic system, 416 Monte Carlo method, 453 Mouse cell, 484 MRI, 513, 530 Mulliken, R., 384 Multiple sclerosis, 386 Myglobin, oxygen binding, 144, 397

N N shell, 341 NADH, 28, 208 NADP, 210, 504

586

INDEX

NADPH, 28 Native phase, 109 Natural linewidth, 472 Natural logarithm, 182 Near infrared region, 13 Near-field scanning optical microscopy, 493 Neighboring group contribution, 520 Nernst equation, 197 Nernst filament, 465 Newton, 11, 557 Newton, I., 10, 313 Nicotinamide adenine dinucleotide, 28 Nitric oxide, 386 Nitrogen, biochemical reactivity, 382 Nitrogen monoxide, see nitric oxide, 386 Nitrogen narcosis, 133 Nitroxide radical, 540 NMR, 513 2D, 534 NMR spectrometer, 518 Nodal plane, 344 Node, radial, 343 NOE, 532 NOE enhancement factor, 534 NOESY, 535 Non-competitive inhibition, 281 Noncontact mode, 329 Nondegenerate, 333 Nonelectrolyte solution, 110 Non-expansion work, 88 Nonpolar molecule, 426 Normal boiling point, 103, 104 Normal freezing point, 104, 105 Normal melting point, 105 Normal mode, 479 tetrahedral molecule, 481 Normalization constant, 324 NSOM, 493 n-to-π* Transition, 487 Nuclear charge, 348 Nuclear g-factor, 514 Nuclear magnetic resonance, 513 Nuclear magnetogyric ratio, 514 Nuclear magneton, 514 Nuclear Overhauser effect, 532 Nuclear Overhauser effect spectroscopy, 535 Nuclear spin quantum number, 513 Nucleation center, 99 Nucleation–condensation model, 255 Nucleic acid, 446 Nucleon number, 1 Number components, 129 degrees of freedom, 129 vibrational modes, 478 Nutritional calorie, 43

O Ocean freezing, 124 Ocular fluid, 501 Off-diagonal peaks, 534 Open system, 24 Operator, 319

Opsin, 502 Optical activity, 488 Orbital antibonding, 375, 385 atomic, 340 bonding, 374, 385 delocalized, 391 Gaussian type, 400 hybrid, 369 molecular, 373 Orbital angular momentum quantum number, 335, 340 Orbital approximation, 346 Orbital overlap, 376 Order differential equation, 230 elementary reaction, 248 reaction, 224 Orthorhombic system, 416 Oscillator, 335 Osmometry, 127 Osmosis, 125 cell structure, 126 Osmotic pressure, 125 Overall order, 224 Overall quantum yield, 495 Overhauser effect, 532 Overlap, 376 Overlap integral, 376 Overtone, 477 Oxidation number, 190 Oxidative phosphorylation, 209 Oxoanion hole, 284 Oxygen, biochemical reactivity, 382 Oxygen attachment, 394 Oxygen binding, 144 hemoglobin, 397 Oxygen reduction, 191 Ozone, 504 polarity, 427

P p Electron, 341 p Orbital, 341, 345 p Subshell, 341 PAGE, 292 Paired spins, 347 Parabolic potential energy, 335 Parallel beta sheet, 445 Paramagnetic species, 383 Partial charge, 385 polypeptide, 425 Partial derivative, 37, 287 Partial fraction, 233 Partial molar Gibbs energy, 110 Partial molar property, 110 Partial pressure, 10, 130 Partial vapor pressure, 112 Particle in a box, 324 Particle on a ring, 331 Particle on a sphere, 334 Partition coefficient, 289 Pascal, 6, 557 Pascal’s triangle, 523

Passive transport, 187, 285 Patch clamp technique, 294 Patch electrode, 294 Pauli exclusion principle, 347, 359 Pauli principle, 359, 366 Pauling, L., 384, 442 PDT, 505 Penetration, 348 Peptide group dipole moment, 428 VB description, 371 Peptide link, 2, 442, 482 Peptide link cleavage, 284 Perfect gas equation of state, 8 Gibbs energy, 97 heat capacity difference, 41 internal energy, 35 molar enthalpy, 39 Period, 351 Periodicity, 350 Peripheral protein, 450 Permittivity, 11, 425 Peroxynitrite ion, 386 Perpetual motion machine, 38 Persistence length, 68 pH amphoteric anion, 169 autoprotolysis contribution, 175 calculation, 161 definition, 157 Pharmaco*kinetics, 234 Phase, 46 Phase boundary, 99 Phase diagram, 99 protein, 109 Phase problem, 420 Phase rule, 129 Phase transition, 47, 94 entropy of, 75 membrane, 108 Phenoxy radical, 544 Phenylalanine, electronic structure, 334 Pheophytin, 504 Phosphate bond, 152 Phosphate-ester bond, 2 Phosphodiester bond, 2 Phospholipid, 3, 450 Phosphonate transition state, 284 Phosphorescence, 490, 491 Photobiology, 494 Photobleaching, 510 Photocatalyst, 505 Photodimerization, 504 Photodiode, 465 Photodynamic therapy, 505 Photoelectric effect, 315 Photon, 13, 315 Photon scattering, 464 Photophosphorylation, 210 Photosensitization, 505 Photosynthesis, 209, 503 general scheme, 211 Photosystem I and II, 210, 504 Photovoltaic device, 466

INDEX

Physical state, 5 Physiological buffer, 171 Pi (π) bond, 367, 376 in complexes, 396 Pi (π) orbital, 376 Pi/2 (π/2) pulse, 528 Pi (π)-donor ligand, 405 Ping-pong reaction, 278 pi-to-pi*(π-to-π*) Transition, 487 Planar bilayer, 450 Planck, M., 314 Planck’s constant, 314 Plane polarized, 488 Planes, separation of, 417 Plasma, 473 Plasmids, STM image, 330 Plasmon, 473 Plasmon resonance, 473 Plastocyanin, 210 Plastoquinone, 210 Pleated sheet, 442 Plot Eadie–Hofstee, 306 Haines, 306 Lineweaver–Burk, 275 Ramachandran, 444 Stern–Volmer, 497 Polar bond, 384 polar molecule, 426 Polarizability, 431, 478 Polarizability volume, 432 Polarizable molecule, 431 Polarization mechanism, 524 Polarized light, 488 Polyacrylamide gel electrophoresis, 292 Polyatomic molecule, 367 MO description, 387 Polychromator, 465 Polyelectrolyte, 119 Polyene, spectroscopic transitions, 402 Polymorph, 106 Polynucleotide, 2 Polypeptide, 2, 443 melting temperature, 132 Polyprotic acid, 165 Polysaccharide, 2, 448 Population, 26 states and intensity, 471 Porphine ring, 361 Potential, 19 standard cell, 197 Potential energy, 11 parabolic, 335 Powder diffractometer, 422 Power, 44 Power series, 124 Prebiotic reaction, 242 Precession, 527 Pre-equilibrium, 253 Pre-exponential factor, 236 interpretation, 260 Pressure, 6 KMT, 267 Primary kinetic isotope effect, 363 Primary quantum yield, 495

Primary structure, 3 Principal quantum number, 338, 340 Principle Aufbau, 349 building-up, 349 Franck–Condon, 486 Le Chatelier’s, 150 Pauli, 359, 366 Pauli exclusion, 347, 359 uncertainty, 321 Probabilistic interpretation, 320 Probability density, 320 Probe pulse, 264 Product rule, 37 Proflavin, 249 Projective reconstruction, 531 Promotion, 368 Protease, 437 Protein food, 56 phase diagram, 109 vibrational spectroscopy, 482 Protein biosynthesis, 152 Protein crystallization, 110 Protein folding, 254, 445 Protein structure, 442 Protein unfolding, 107, 254 Proton decoupling, 532 Proton magnetic resonance, 518 Proton mobility, 291 Proton pump, 295 Proton transfer, 156 Protonation, 158 Pseudo-first order reaction, 226 Pseudo-second order reaction, 226 Pulse techniques, 527 Pulse-field electrophoresis, 292 Pyridine, elpot surface, 401 Pyridone, 270 Pyruvate ion, 153

Q QSAR, 454 Quadratic contribution to energy, 68 Quadratic equation, 162 Quantitative structure–activity relationship, 454 Quantization angular momentum, 333 energy, 315 Quantum number azimuthal, 340 introduced, 325 magnetic, 335, 340 nuclear spin, 513 orbital angular momentum, 335, 340 particle ina box, 325 principal, 338, 340 spin, 347 spin magnetic, 347 vibrational, 336 Quantum theory, experimental foundation, 314 Quantum yield, 495 fluorescence, 496

Quaternary structure, 4 Quenching, 497 Quenching method, 221 Quotient rule, 37

R Radial distribution function, 342 liquid, 439 Radial node, 343 Radial wavefunction, 341 Radiative decay, 490 Radical, 50 Radio region, 13 Radius of gyration, 441 Ramachandran plot, 444 Raman imaging, 484 Raman microscopy, 484, 511 Raman optical activity, 489 Raman spectrometer, 466 Raman spectroscopy, 464 Raman transitions, 481 Random coil, 440 Random walk, 285 Raoult, F., 112 Raoult’s law, 113 Rate constant, 223 electron transfer, 297 relation to equilibrium constant, 243 variation with temperature, 237 viscosity dependence, 257 Rate law determination, 225 integrated, 228 introduced, 223 Rate-determining step, 251 Rayleigh radiation, 464 Rayleigh ratio, 412 Rayleigh scattering, 412 Reaction center (photosynthesis), 503 Reaction coordinate, 261 Reaction dynamics, 259 Reaction enthalpy, variation with temperature, 62 Reaction Gibbs energy, 136 Reaction mechanism, 224 Reaction order, 224 Reaction profile, 259 Reaction quotient, 138 half-reaction, 191 Reaction rate, 221 variation with temperature, 235 Reactions that ‘go’, 136 Reactive oxygen series, 384 Real gas, 8 Real solution, 114 Real-time analysis, 220 Recognition, molecular, 437 Redox couple, 190 Redox electrode, 194 Redox reaction, 181, 189 Reduced mass, 338, 474 Reference state, 59 Reflection (X-ray), 419 Refractive index, 473

587

588

INDEX

Relation between pH and pOH, 159 pKa and pKb, 159 Relativistic effect, 360 Relaxation (NMR), 529 Relaxation technique, 221, 245 Relaxation time, 246, 529 Relaxed state, 397 Reorganization energy, 299 Residual entropy, 82 Residue, 2 Resolution, 317 Resonance, 372, 513 Resonance condition, 515 Resonance energy transfer, 499 Resonance hybrid, 372 Resonance integral, 388 Resonance Raman spectroscopy, 482 Respiratory acidosis, 173 Respiratory alkalosis, 173 Respiratory chain, 208 Resting potential, 188 Retina, 502 Retinal, 2, 248, 404, 501 Reverse micelle, 449 Reversible process, 32 Rhodopsin, 502 Rhombohedral system, 416 Riboflavin equilibrium, 204 Ribonuclease, melting, 108 Ribonucleic acid, 273 Ribosome, 273 Ribozyme, 273 Ring current, 521 RNA, 273, 447 ROA, 489 Root mean square deviation, 322 Root mean square separation, 440 Root-mean-square speed, 267 ROS, 384 Rotating frame, 527 Rotation, 331 Rule chain, 37 Corey–Pauling, 442 exclusion, 481 Hund’s, 350 phase, 129 product, 37 quotient, 37 selection, 470 vibrational selection, 477

S s Electron, 341 s Orbital, 341 s Subshell, 341 Salt bridge, 192 Salt solution, pH, 164 SAR, 92 SATP, 9, 10 Saturation, 529 Scanning electron microscopy, 318 Scanning probe microscopy, 329

Scanning tunneling microscopy, 329 Scatchard equation, 134 Scattering, 464 SCF, 398 Schrödinger equation, 319 justification, 358 Schrödinger, E., 319 SCUBA diving, 133 SDS, 449 SDS-PAGE, 292 Second ionization energy, 353 Second law Fick’s, 286, 304 thermodynamics, 71 Secondary kinetic isotope effect, 363 Secondary structure, 3, 442 Second-order rate law half-life, 232 integrated, 231 Secular determinant, 389 Secular equation, 388 Sedimentation, 407 Sedimentation constant, 408 Sedimentation equilibrium, 409 Selection rule, 470 Selectivity filter, 295 Self-assembly, 407 Self-consistent field procedure, 398 SEM, 318 Semi-empirical method, 398 Semipermeable membrane, 125 Separation of variables, 330, 359 Sequential reactions, 277 SHE, 198 Shell, 341, 347 Shielded, 519 Shielded nuclear charge, 348 Shielding, 348 Shielding constant (NMR), 518 Short-range order, 439 SI base units, 557 SI prefixes, 557 SI units, 5 Sigma (σ) bond, 367, 379 Sigma (σ) electron, 374 Sigma (σ) orbital, 374 Sigma (σ)-donor ligand, 405 Sign convention, work and heat, 29 Simultaneous equations, 389 Single molecule spectroscopy, 493 Singlet state, 491 SIR model, 308 Slice selection, 531 Sneeze analogy, 72, 74 SOD, 205 Sodium dodecyl sulfate, 449 Solid, 4 Solute, 110 Solute activity, 118 Solvent, 110 Solvent activity, 118 Solvent contribution, 521 Solvent-accessible surface, 400 Speciation, 168 Specific enthalpy, 54

Specific heat capacity, 33 Specific selection rule, 470 Spectrometer, 465 Spectroscopy, 463 correlation, 534 infrared, 476 nuclear Overhauser effect, 535 resonance Raman, 482 single-molecule, 493 time-resolved, 263, 499 vibrational Raman, 478 Spectrum, 314 electromagnetic, 12 Spherical coordinates, 376 Spherically symmetrical, 342 Spin, 347 Spin correlation, 350 Spin density, 539 Spin label, 540 Spin magnetic quantum number, 347 Spin pairing, 366 MO theory, 379 Spin probe, 540 spin quantum number, 347 Spin–lattice relaxation time, 529 Spin–orbit coupling, 492 Spin–spin coupling constant, 521 Spin–spin relaxation time, 529 SPM, 329 n sp Hybrid orbital, 369 Spontaneous change, 69 Spontaneous chemical reaction, 83, 135 Spontaneous emission, 471 Spontaneous process, 85 Spontaneous reaction, criterion, 142 Stability condition, 94 Stable compound, 149 Standard cell potential, 197 from standard potentials, 203 Standard chemical potential, 112 Standard conditions, 10 Standard enthalpy combustion, 53 formation, 59 fusion, 48 sublimation, 49 vaporization, 47 Standard Gibbs energy of formation, 147 Standard hydrogen electrode, 198 Standard molar concentration, 117 Standard molar entropy, 78 Standard oxidation potential, 198 Standard potential, 198 from two others, 205 Standard reaction enthalpy, 58 from cell potential, 206 Standard reaction entropy, 82 from cell potential, 206 Standard reaction Gibbs energy, 137, 146 variation with composition, 138 Standard reduction potential, 198 Standard state, 9, 10, 46 biological, 139 summary, 119 State of matter, 4

INDEX

State function, 37 Steady-state approximation, 252 Steric factor, 261 Stern–Volmer equation, 497 Stern–Volmer plot, 497 Stimulated absorption, 470 Stimulated emission, 470 STM, 329 Stokes radiation, 464 Stokes’s law, 290 Stokes–Einstein relation, 288, 409 Stopped-flow technique, 220 STP, 10 Stratopause, 494 Stratosphere, 494, 504 Strong acid, 158 Strong base, 159 Structure factor light scattering, 412 X-ray, 420 Structure-activity relation, 92 Structure-based design, 453 Sublimation, 49 Sublimation vapor pressure, 101 Subshell, 341 Substrate, 273 Sucrose, acid hydrolysis, 236 Sulfuric acid, 158, 165 Supercoiled DNA, 447 Supercritical fluid, 104 Superheated liquid, 99 Superoxide dismutase, 205, 383 Superposition, 321 Surface plasmon resonance, 473 Surfactant, 449 Surfactant parameter, 449 Surroundings, 24 Svedberg, 408 Symmetric stretch, 479 System, 24 Système International, 5

T T1-weighted image, 531 t2g Orbital, 393 T2-weighted image, 531 Tapping mode, 329 Taylor series, 124 TEM, 317 Temperature, 6 conventional, 46 molecular interpretation, 26 Temperature dependence enthalpy, 41 reaction enthalpy, 62 Temperature jump, 246 Tense state, 397 Tertiary structure, 4 Tetragonal system, 416 Tetrahedral hybridization, 370 Theorem, equipartition, 68 Theory activated complex, 261 Brønsted–Lowry, 156

chemiosmotic, 209 collision, 259 crystal field, 392 Debye–Hückel, 184 extended Hückel, 399 FEMO, 405 Förster, 500 kinetic molecular, 267 ligand field, 392, 394 Marcus, 298, 499 molecular orbital, 364, 373 transition state, 261 valence, 364 valence-bond, 364 valence-shell electron repulsion, 2 Thermal analysis, 101 Thermal denaturation, 45 Thermal energy, 14 Thermal equilibrium, 6 Thermal motion, 14 Thermochemical equation, 47 Thermodynamic control, 258 Thermodynamic stability, 149 Thermodynamic temperature, 6 Thermodynamics, 21 Thermogram, 44 Thermosphere, 494 Third Law of thermodynamics, 78 Third-Law entropy, 78 TIBO, 459 Time constant, 231 Time-of-flight spectrometer, 411 Time-resolved spectroscopy, 263, 499 Tonne, 558 Torr, 6 Total energy, 12 Total entropy change, 84 Trajectory, 313 Transducin, 503 Transfer potential, 153 Transfer RNA, 448 Transition, 470 electronic, 487 Transition dipole moment, 469 Transition metal, 351 Transition state, 238, 261 Transition state theory, 261 Transition temperature, 99 Translation, 324 Transmission coefficient, 262 Transmission electron microscopy, 317 Transmission probability, 328 Transmittance, 466 Transport across membranes, 285 Trigonal planar hybridization, 370 Triple point, 104, 105 Triplet state, 491 Tristearin, 56 tRNA, 448 Tropopause, 494 Troposphere, 494 Tungsten–iodine lamp, 465 Tunneling, 328 Turning point, 486 Turnover frequency, 279

589

Two-dimensional electrophoresis, 293 Two-dimensional NMR, 534 Tyrosine radical, 537

U Ubiquitin, 45 Ultra centrifuge, 408 Ultracentrifugation, 408 Ultraviolet damage, 504 Uncertainty broadening, see lifetime broadening, 472 Uncertainty principle, 321 Uncompetitive inhibition, 281 Ungerade symmetry, 375 Unilamellar vesicle, 450 Unimolecular reaction, 248 Unique reaction rate, 222 Unit cell, 415 Unstable compound, 149 UVA and UVB, 504

V Vacuum permittivity, 11 Vacuum ultraviolet region, 13 Valence bond theory, 364 summary of terms, 372 Valence electron, 348 Valence theory, 364 Valence-shell electron pair repulsion model, 364 Valence-shell electron repulsion theory, 2 van ‘t Hoff equation chemical equilibrium, 150 osmotic pressure, 125 van ‘t Hoff isochore, 150 van der Waals interaction, 2, 425 Vapor deposition, 49 Vapor diffusion, 422 Vapor pressure, 100 Vaporization entropy of, 76 standard enthalpy, 47 thermodynamic basis, 99 Variation with temperature cell potential, 206 chemical equilibrium, 150 diffusion coefficient, 287 Gibbs energy, 98 heat capacity, 66 rate constant, 237 reaction rate, 235 viscosity, 288 VB theory, 364 Vector, 332 addition and subtraction, 427 Vertical transition, 486 Vesicle, 450 VESPR, 2 Vibration, 335 Vibrational frequency, 336 diatomic molecule, 362 Vibrational microscopy, 483 Vibrational modes, number of, 478 Vibrational quantum number, 336

590

INDEX

Vibrational Raman spectroscopy, 478 Vibrational selection rule, 477 Vibrational spectra, 474 Vibrational structure, 486 Vibrational transition, 476 Viscosity, variation with temperature, 288 Viscous drag, 290 Visible region, 12, 13 Vision, 501 Voltage-gated channel, 188 Voltaic cell, 192 Volume, 5 Volume element, 376 VSEPR model, 364

W Water, 2 MO description, 387 phase diagram, 105, 106 polarity, 427 radial distribution function, 439 VB description, 369 viscosity, 288

Watt, 44, 557 Wave, 12 Wavefunction angular, 342 harmonic oscillator, 336 introduced, 318 particle in a 2D box, 330 particle in a box, 324 radial, 341 VB, 365 Wavelength, 12, 464 Wavenumber, 464 Wave–particle duality, 315, 316 Weak acid, 158 pH calculation, 161 Weak base, 159 calculation of pH, 163 Weight of configuration, 80 White light, 485 Wide-field epifluorescence method, 493 Work, 10, 23 expansion, 30 Gibbs energy, 88 maximum, 31

molecular interpretation, 26 non-expansion, 88 raising a weight, 29 Work function, 315 Wrinkle, Nature’s abhorrence of, 287

X Xanthophyll, 502 X-ray crystallography, 414 X-ray diffraction, 414 X-ray diffractometer, 422 X-ray generation, 415 X-ray region, 13

Z Z form (DNA), 447 Zero-current cell potential, 196 Zero-point energy, 326 Zeroth-order rate law, 228 Zeroth-order reaction, 226 Zinc, biological role, 356 Zwitterionic form, 169

Physical Chemistry for the Life Sciences 2e [PDF] | Online Book Share (2024)

References

Top Articles
Latest Posts
Article information

Author: Amb. Frankie Simonis

Last Updated:

Views: 6421

Rating: 4.6 / 5 (76 voted)

Reviews: 83% of readers found this page helpful

Author information

Name: Amb. Frankie Simonis

Birthday: 1998-02-19

Address: 64841 Delmar Isle, North Wiley, OR 74073

Phone: +17844167847676

Job: Forward IT Agent

Hobby: LARPing, Kitesurfing, Sewing, Digital arts, Sand art, Gardening, Dance

Introduction: My name is Amb. Frankie Simonis, I am a hilarious, enchanting, energetic, cooperative, innocent, cute, joyous person who loves writing and wants to share my knowledge and understanding with you.