457454 physical chemistry quantum chemistry

352
Chemistry 351 and 352 Physical Chemistry I and II Darin J. Ulness Fall 2006 – 2007

Upload: elcachero

Post on 18-Nov-2014

1.224 views

Category:

Documents


30 download

TRANSCRIPT

Page 1: 457454 Physical Chemistry Quantum Chemistry

Chemistry 351 and 352

Physical Chemistry I and II

Darin J. Ulness

Fall 2006 – 2007

Page 2: 457454 Physical Chemistry Quantum Chemistry

Contents

I Basic Quantum Mechanics 15

1 Quantum Theory 161.1 The “Fall” of Classical Physics . . . . . . . . . . . . . . . . . . . . 16

1.2 Bohr’s Atomic Theory . . . . . . . . . . . . . . . . . . . . . . . . 17

1.2.1 First Attempts at the Structure of the Atom . . . . . . . . 17

2 The Postulates of Quantum Mechanics 222.1 Postulate I . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

2.2 How to normalize a wavefunction . . . . . . . . . . . . . . . . . . 23

2.3 Postulates II and II . . . . . . . . . . . . . . . . . . . . . . . . . . 24

3 The Setup of a Quantum Mechanical Problem 273.1 The Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

3.2 The Quantum Mechanical Problem . . . . . . . . . . . . . . . . . 27

3.3 The Average Value Theorem . . . . . . . . . . . . . . . . . . . . . 29

3.4 The Heisenberg Uncertainty Principle . . . . . . . . . . . . . . . . 30

4 Particle in a Box 314.1 The 1D Particle in a Box Problem . . . . . . . . . . . . . . . . . . 31

4.2 Implications of the Particle in a Box problem . . . . . . . . . . . 34

5 The Harmonic Oscillator 385.1 Interesting Aspects of the Quantum Harmonic Oscillator . . . . . 40

i

Page 3: 457454 Physical Chemistry Quantum Chemistry

5.2 Spectroscopy (An Introduction) . . . . . . . . . . . . . . . . . . . 42

II Quantum Mechanics of Atoms and Molecules 45

6 Hydrogenic Systems 466.1 Hydrogenic systems . . . . . . . . . . . . . . . . . . . . . . . . . . 46

6.2 Discussion of the Wavefunctions . . . . . . . . . . . . . . . . . . . 49

6.3 Spin of the electron . . . . . . . . . . . . . . . . . . . . . . . . . . 51

6.4 Summary: the Complete Hydrogenic Wavefunction . . . . . . . . 52

7 Multi-electron atoms 557.1 Two Electron Atoms: Helium . . . . . . . . . . . . . . . . . . . . 55

7.2 The Pauli Exclusion Principle . . . . . . . . . . . . . . . . . . . . 56

7.3 Many Electron Atoms . . . . . . . . . . . . . . . . . . . . . . . . 58

7.3.1 The Total Hamiltonian . . . . . . . . . . . . . . . . . . . . 59

8 Diatomic Molecules and the Born Oppenheimer Approximation 608.1 Molecular Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

8.1.1 The Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . 61

8.1.2 The Born—Oppenheimer Approximation . . . . . . . . . . 62

8.2 Molecular Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 63

8.2.1 The Morse Oscillator . . . . . . . . . . . . . . . . . . . . . 64

8.2.2 Vibrational Spectroscopy . . . . . . . . . . . . . . . . . . . 66

9 Molecular Orbital Theory and Symmetry 679.1 Molecular Orbital Theory . . . . . . . . . . . . . . . . . . . . . . 67

9.2 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

10 Molecular Orbital Diagrams 7210.1 LCAO–Linear Combinations of Atomic Orbitals . . . . . . . . . 72

10.1.1 Classification of Molecular Orbitals . . . . . . . . . . . . . 73

Page 4: 457454 Physical Chemistry Quantum Chemistry

10.2 The Hydrogen Molecule . . . . . . . . . . . . . . . . . . . . . . . 74

10.3 Molecular Orbital Diagrams . . . . . . . . . . . . . . . . . . . . . 76

10.4 The Complete Molecular Hamiltonian and Wavefunction . . . . . 78

11 An Aside: Light Scattering–Why the Sky is Blue 7911.1 The Classical Electrodynamics Treatment of Light Scattering . . . 79

11.2 The Blue Sky . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

11.2.1 Sunsets . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82

11.2.2 White Clouds . . . . . . . . . . . . . . . . . . . . . . . . . 83

III Statistical Mechanics and The Laws of Thermody-namics 88

12 Rudiments of Statistical Mechanics 8912.1 Statistics and Entropy . . . . . . . . . . . . . . . . . . . . . . . . 89

12.1.1 Combinations and Permutations . . . . . . . . . . . . . . . 90

12.2 Fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92

13 The Boltzmann Distribution 9413.1 Partition Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 96

13.1.1 Relation between the Q and W . . . . . . . . . . . . . . . 97

13.2 The Molecular Partition Function . . . . . . . . . . . . . . . . . . 99

14 Statistical Thermodynamics 103

15 Work 10715.1 Properties of Partial Derivatives . . . . . . . . . . . . . . . . . . . 107

15.1.1 Summary of Relations . . . . . . . . . . . . . . . . . . . . 107

15.2 Definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108

15.2.1 Types of Systems . . . . . . . . . . . . . . . . . . . . . . . 108

15.2.2 System Parameters . . . . . . . . . . . . . . . . . . . . . . 109

Page 5: 457454 Physical Chemistry Quantum Chemistry

15.3 Work and Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109

15.3.1 Generalized Forces and Displacements . . . . . . . . . . . 110

15.3.2 PV work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

16 Maximum Work and Reversible changes 11316.1 Maximal Work: Reversible versus Irreversible changes . . . . . . . 113

16.2 Heat Capacity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115

16.3 Equations of State . . . . . . . . . . . . . . . . . . . . . . . . . . 116

16.3.1 Example 1: The Ideal Gas Law . . . . . . . . . . . . . . . 116

16.3.2 Example 2: The van der Waals Equation of State . . . . . 117

16.3.3 Other Equations of State . . . . . . . . . . . . . . . . . . . 118

17 The Zeroth and First Laws of Thermodynamics 11917.1 Temperature and the Zeroth Law of Thermodynamics . . . . . . . 119

17.2 The First Law of Thermodynamics . . . . . . . . . . . . . . . . . 121

17.2.1 The internal energy state function . . . . . . . . . . . . . . 121

18 The Second and Third Laws of Thermodynamics 12418.1 Entropy and the Second Law of Thermodynamics . . . . . . . . . 124

18.1.1 Statements of the Second Law . . . . . . . . . . . . . . . . 127

18.2 The Third Law of Thermodynamics . . . . . . . . . . . . . . . . . 127

18.2.1 The Third Law . . . . . . . . . . . . . . . . . . . . . . . . 128

18.2.2 Debye’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . 129

18.3 Times Arrow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130

IV Basics of Thermodynamics 134

19 Auxillary Functions and Maxwell Relations 13519.1 The Other Important State Functions of Thermodynamics . . . . 135

19.2 Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136

19.2.1 Heuristic definition: . . . . . . . . . . . . . . . . . . . . . . 137

Page 6: 457454 Physical Chemistry Quantum Chemistry

19.3 Helmholtz Free Energy . . . . . . . . . . . . . . . . . . . . . . . . 137

19.3.1 Heuristic definition: . . . . . . . . . . . . . . . . . . . . . . 138

19.4 Gibbs Free Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 138

19.4.1 Heuristic definition: . . . . . . . . . . . . . . . . . . . . . . 139

19.5 Heat Capacity of Gases . . . . . . . . . . . . . . . . . . . . . . . . 139

19.5.1 The Relationship Between CP and CV . . . . . . . . . . . 139

19.6 The Maxwell Relations . . . . . . . . . . . . . . . . . . . . . . . . 140

20 Chemical Potential 14220.1 Spontaneity of processes . . . . . . . . . . . . . . . . . . . . . . . 142

20.2 Chemical potential . . . . . . . . . . . . . . . . . . . . . . . . . . 144

20.3 Activity and the Activity coefficient . . . . . . . . . . . . . . . . . 146

20.3.1 Reference States . . . . . . . . . . . . . . . . . . . . . . . 147

20.3.2 Activity and the Chemical Potential . . . . . . . . . . . . 148

21 Equilibrium 15121.0.3 Equilibrium constants in terms of KC . . . . . . . . . . . . 153

21.0.4 The Partition Coefficient . . . . . . . . . . . . . . . . . . . 153

22 Chemical Reactions 15622.1 Heats of Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . 156

22.1.1 Heats of Formation . . . . . . . . . . . . . . . . . . . . . . 157

22.1.2 Temperature dependence of the heat of reaction . . . . . . 157

22.2 Reversible reactions . . . . . . . . . . . . . . . . . . . . . . . . . . 158

22.3 Temperature Dependence of Ka . . . . . . . . . . . . . . . . . . . 159

22.4 Extent of Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . 160

23 Ionics 16123.1 Ionic Activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161

23.1.1 Ionic activity coefficients . . . . . . . . . . . . . . . . . . . 162

23.2 Theory of Electrolytic Solutions . . . . . . . . . . . . . . . . . . . 163

Page 7: 457454 Physical Chemistry Quantum Chemistry

23.3 Ion Mobility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

23.3.1 Ion mobility . . . . . . . . . . . . . . . . . . . . . . . . . . 165

24 Thermodynamics of Solvation 16924.1 The Born Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 170

24.1.1 Free Energy of Solvation for the Born Model . . . . . . . . 173

24.1.2 Ion Transfer Between Phases . . . . . . . . . . . . . . . . . 174

24.1.3 Enthalpy and Entropy of Solvation . . . . . . . . . . . . . 174

24.2 Corrections to the Born Model . . . . . . . . . . . . . . . . . . . . 175

25 Key Equations for Exam 4 177

V Quantum Mechanics and Dynamics 180

26 Particle in a 3D Box 18126.1 Particle in a Box . . . . . . . . . . . . . . . . . . . . . . . . . . . 181

26.2 The 3D Particle in a Box Problem . . . . . . . . . . . . . . . . . . 183

27 Operators 18727.1 Operator Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

27.2 Orthogonality, Completeness, and the Superposition Principle . . 191

28 Angular Momentum 19228.1 Classical Theory of Angular Momentum . . . . . . . . . . . . . . 192

28.2 Quantum theory of Angular Momentum . . . . . . . . . . . . . . 193

28.3 Particle on a Ring . . . . . . . . . . . . . . . . . . . . . . . . . . . 194

28.4 General Theory of Angular Momentum . . . . . . . . . . . . . . . 195

28.5 Quantum Properties of Angular Momentum . . . . . . . . . . . . 199

28.5.1 The rigid rotor . . . . . . . . . . . . . . . . . . . . . . . . 200

Page 8: 457454 Physical Chemistry Quantum Chemistry

29 Addition of Angular Momentum 20129.1 Spin Angular Momentum . . . . . . . . . . . . . . . . . . . . . . . 201

29.2 Addition of Angular Momentum . . . . . . . . . . . . . . . . . . . 202

29.2.1 The Addition of Angular Momentum: General Theory . . 202

29.2.2 An Example: Two Electrons . . . . . . . . . . . . . . . . . 203

29.2.3 Term Symbols . . . . . . . . . . . . . . . . . . . . . . . . . 204

29.2.4 Spin Orbit Coupling . . . . . . . . . . . . . . . . . . . . . 205

30 Approximation Techniques 20730.1 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . 207

30.2 Variational method . . . . . . . . . . . . . . . . . . . . . . . . . . 209

31 The Two Level System and Quantum Dynamics 21131.1 The Two Level System . . . . . . . . . . . . . . . . . . . . . . . . 211

31.2 Quantum Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 214

VI Symmetry and Spectroscopy 220

32 Symmetry and Group Theory 22132.1 Symmetry Operators . . . . . . . . . . . . . . . . . . . . . . . . . 222

32.2 Mathematical Groups . . . . . . . . . . . . . . . . . . . . . . . . . 222

32.2.1 Example: The C2v Group . . . . . . . . . . . . . . . . . . 223

32.3 Symmetry of Functions . . . . . . . . . . . . . . . . . . . . . . . . 223

32.3.1 Direct Products . . . . . . . . . . . . . . . . . . . . . . . . 225

32.4 Symmetry Breaking and Crystal Field Splitting . . . . . . . . . . 225

33 Molecules and Symmetry 22833.1 Molecular Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 228

33.1.1 Normal Modes . . . . . . . . . . . . . . . . . . . . . . . . 229

33.1.2 Normal Modes and Group Theory . . . . . . . . . . . . . . 229

Page 9: 457454 Physical Chemistry Quantum Chemistry

34 Vibrational Spectroscopy and Group Theory 23134.1 IR Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231

34.2 Raman Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . 233

35 Molecular Rotations 23535.1 Relaxing the rigid rotor . . . . . . . . . . . . . . . . . . . . . . . . 236

35.2 Rotational Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . 236

35.3 Rotation of Polyatomic Molecules . . . . . . . . . . . . . . . . . . 237

36 Electronic Spectroscopy of Molecules 24036.1 The Structure of the Electronic State . . . . . . . . . . . . . . . . 240

36.1.1 Absorption Spectra . . . . . . . . . . . . . . . . . . . . . . 241

36.1.2 Emission Spectra . . . . . . . . . . . . . . . . . . . . . . . 241

36.1.3 Fluorescence Spectra . . . . . . . . . . . . . . . . . . . . . 242

36.2 Franck—Condon activity . . . . . . . . . . . . . . . . . . . . . . . 243

36.2.1 The Franck—Condon principle . . . . . . . . . . . . . . . . 243

37 Fourier Transforms 24537.1 The Fourier transformation . . . . . . . . . . . . . . . . . . . . . 245

VII Kinetics and Gases 249

38 Physical Kinetics 25038.1 kinetic theory of gases . . . . . . . . . . . . . . . . . . . . . . . . 250

38.2 Molecular Collisions . . . . . . . . . . . . . . . . . . . . . . . . . . 252

39 The Rate Laws of Chemical Kinetics 25439.1 Rate Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254

39.2 Determination of Rate Laws . . . . . . . . . . . . . . . . . . . . . 258

39.2.1 Differential methods based on the rate law . . . . . . . . . 259

39.2.2 Integrated rate laws . . . . . . . . . . . . . . . . . . . . . . 259

Page 10: 457454 Physical Chemistry Quantum Chemistry

40 Temperature and Chemical Kinetics 26140.1 Temperature Effects on Rate Constants . . . . . . . . . . . . . . . 261

40.1.1 Temperature corrections to the Arrhenious parameters . . 262

40.2 Theory of Reaction Rates . . . . . . . . . . . . . . . . . . . . . . 262

40.3 Multistep Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . 265

40.4 Chain Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 267

41 Gases and the Virial Series 26941.1 Equations of State . . . . . . . . . . . . . . . . . . . . . . . . . . 269

41.2 The Virial Series . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

41.2.1 Relation to the van der Waals Equation of State . . . . . . 271

41.2.2 The Boyle Temperature . . . . . . . . . . . . . . . . . . . 272

41.2.3 The Virial Series in Pressure . . . . . . . . . . . . . . . . . 272

41.2.4 Estimation of Virial Coefficients . . . . . . . . . . . . . . . 273

42 Behavior of Gases 27442.1 P, V and T behavior . . . . . . . . . . . . . . . . . . . . . . . . . 274

42.1.1 α and κT for an ideal gas . . . . . . . . . . . . . . . . . . . 275

42.1.2 α and κT for liquids and solids . . . . . . . . . . . . . . . . 275

42.2 Heat Capacity of Gases Revisited . . . . . . . . . . . . . . . . . . 276

42.2.1 The Relationship Between CP and CV . . . . . . . . . . . 276

42.3 Expansion of Gases . . . . . . . . . . . . . . . . . . . . . . . . . . 279

42.3.1 Isothermal and Adiabatic expansions . . . . . . . . . . . . 279

42.3.2 Heat capacity CV for adiabatic expansions . . . . . . . . . 280

42.3.3 When P is the more convenient variable . . . . . . . . . . 281

42.3.4 Joule expansion . . . . . . . . . . . . . . . . . . . . . . . . 282

42.3.5 Joule-Thomson expansion . . . . . . . . . . . . . . . . . . 283

43 Entropy of Gases 28643.1 Calculation of Entropy . . . . . . . . . . . . . . . . . . . . . . . . 286

43.1.1 Entropy of Real Gases . . . . . . . . . . . . . . . . . . . . 288

Page 11: 457454 Physical Chemistry Quantum Chemistry

VIII More Thermodyanmics 292

44 Critical Phenomena 29344.1 Critical Behavior of fluids . . . . . . . . . . . . . . . . . . . . . . 293

44.1.1 Gas Laws in the Critical Region . . . . . . . . . . . . . . . 294

44.1.2 Gas Constants from Critical Data . . . . . . . . . . . . . . 295

44.2 The Law of Corresponding States . . . . . . . . . . . . . . . . . . 296

44.3 Phase Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . 296

44.3.1 The chemical potential and T and P . . . . . . . . . . . . 297

44.3.2 The Clapeyron Equation . . . . . . . . . . . . . . . . . . . 298

44.3.3 Vapor Equilibrium and the Clausius-Clapeyron Equation . 298

44.4 Equilibria of condensed phases . . . . . . . . . . . . . . . . . . . . 299

44.5 Triple Point and Phase Diagrams . . . . . . . . . . . . . . . . . . 300

45 Transport Properties of Fluids 30145.1 Diffusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 301

45.2 Viscosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 303

45.3 Thermal conductivity . . . . . . . . . . . . . . . . . . . . . . . . . 305

45.3.1 Thermal Conductivity of Gases and Liquids . . . . . . . . 306

45.3.2 Thermal Conductivity of Solids . . . . . . . . . . . . . . . 307

46 Solutions 30846.1 Measures of Composition . . . . . . . . . . . . . . . . . . . . . . . 308

46.2 Partial Molar Quantities . . . . . . . . . . . . . . . . . . . . . . . 308

46.2.1 Notation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 309

46.2.2 Partial Molar Volumes . . . . . . . . . . . . . . . . . . . . 310

46.3 Reference states for liquids . . . . . . . . . . . . . . . . . . . . . . 311

46.3.1 Activity (a brief review) . . . . . . . . . . . . . . . . . . . 311

46.3.2 Raoult’s Law . . . . . . . . . . . . . . . . . . . . . . . . . 312

46.3.3 Ideal Solutions (RL) . . . . . . . . . . . . . . . . . . . . . 314

46.3.4 Henry’s Law . . . . . . . . . . . . . . . . . . . . . . . . . . 316

Page 12: 457454 Physical Chemistry Quantum Chemistry

46.4 Colligative Properties . . . . . . . . . . . . . . . . . . . . . . . . . 318

46.4.1 Freezing Point Depression . . . . . . . . . . . . . . . . . . 318

46.4.2 Osmotic Pressure . . . . . . . . . . . . . . . . . . . . . . . 319

47 Entropy Production and Irreverisble Thermodynamics 32247.1 Fundamentals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322

47.2 The Second Law . . . . . . . . . . . . . . . . . . . . . . . . . . . 324

47.3 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 325

47.3.1 Entropy Production due to Heat Flow . . . . . . . . . . . 326

47.3.2 Entropy Production due to Chemical Reactions . . . . . . 328

47.4 Thermodynamic Coupling . . . . . . . . . . . . . . . . . . . . . . 330

47.5 Echo Phenonmena . . . . . . . . . . . . . . . . . . . . . . . . . . 331

Page 13: 457454 Physical Chemistry Quantum Chemistry

Chemistry 351: PhysicalChemistry I

1

1

Page 14: 457454 Physical Chemistry Quantum Chemistry

Solved Problems

I make-up most of the problems on the problems sets, so it might be helpful to

you to see some of these problems worked out.

Even though there aren’t many “book” problems assigned during the year, you can

still learn a lot be working these and looking that their solutions in the solution

manual.

Keep in mind this chapter provides some examples of how to solve problems for

both physical chemistry I and physical chemistry II. Consequently early in the

course some of the examples might seem very itimidating. Simply skip those

examples as you scan through this chapter.

Tips for solving problems

Working problem sets is the heart and sole of learning physical chemistry. The

only way that you can be sure that you understand a concept at to be able to

solve the problems associated with it.

This takes time and hard work.

But there are some things that you can do to help yourself with these problems.

Tips

2

2

Page 15: 457454 Physical Chemistry Quantum Chemistry

1. Remember nobody cares if you solve any particular problem on the problem

set. They have all been solved before, so if you solve them you will not

become famous nor will you save the world. The only reason you work them

is to learn.

2. Budget your time so that you don’t have to work on an overwhelming number

of problems at a time. Try to whip-off a few on the same day that you get

the problem set. Then work on them consistently during the week. This

will make the problem sets much more efficient at helping you learn.

3. You can do the problem. I don’t assign problems that you cannot do. If you

think you can’t do the problem then maybe you need try a different way of

thinking about it.

4. Part of the trouble is simply understanding what the problem is asking you

to do. There is a tendency to try to start solving the problem before fully

understanding the question.

• Read the question carefully

• Try to think about what topic(s) in lecture and in the notes the problemis dealing with.

• Do not worry about not knowing how to solve it yet.

• Just identify the general ideas that you think you might need.

• Determine wether you need to approach the problem mathematically

or conceptually or both.

• If the question is long, try to identify subsections of it.

5. For problems that require a mathematical approach...

• Do not be afraid. Try to figure out what mathematical techniques youneed to express the solution to the problem.

3

Page 16: 457454 Physical Chemistry Quantum Chemistry

• Do the math; either you will be able to do this or you won’t. It mighttake some review on your part.

• Always check to see if the math makes sense when you are done.

6. For problems that require a conceptual approach...

• Make sure that the physical idea that you are using in your argument iscorrect. If you are not sure, start with a related concept that is better

known by you.

• Look for self-consistency. Does you final answer jive with what youknow.

Problems Dealing With Quantum Mechanics

Problem: What is the periodicity of the following functions

• f(x) = sin2 x

• f(x) = cosx

• f(x) = e−2ix

Solution: For the first function it is easiest to see the periodicity by writing thefunction as f(x) = (sinx)(sinx). We know that this function will repeat zeros

when ever sinx = 0. This occurs at x = nπ, n = 0,±1,±2 . . ., so the periodicityis π. The second function we should remember from trig as having a period of 2π.

Finally for the last function it is best to used Euler’s identity and write

e−2ix = cos 2x+ i sin 2x (1)

The real part of this function, cos 2x, has a period of π as does the imaginary

part, sin 2x. Therefore the entire function has a period of π.

4

Page 17: 457454 Physical Chemistry Quantum Chemistry

Problem: Which of the following functions are eigenfunction of the momentumoperator, px = −i~ d

dx.

• ψ(x) = eikx

• ψ(x) = e−αx2

• ψ(x) = cos kx

Solution: We need to determine if pxψ(x) = λψ(x) where λ is a constant. If

this equation is true then the function is an eigenfunction with eigenvalue λ. For

the case of momentum all we need to do is take the derivative of each function,

multiply by −i~ and check to see if the eigenvalue equation holds.For the first function

pxψ(x) = −i~dψ(x)

dx= −i~de

ikx

dx= ~keikx = ~kψ(x), (2)

so, yes, this function is an eigenfunction of the momentum operator.

For the second function

pxψ(x) = −i~dψ(x)

dx= −i~de

−αx2

dx= 2i~αxe−αx2 = 2i~α↓xψ(x), (3)

so, no, this function is not and eigenfunction of the momentum operator.

For the last function

pxψ(x) = −i~dψ(x)

dx= −i~d cos kx

dx= −i~k

6=cos kxz | sin kx, (4)

so, no, this function is not an eigenfunction of the momentum operator.

Problem: A quantum object is described by the wavefunction ψ(x) = e−αx2.

What is the probability of finding the object further than α away from the origin

(x = 0)?

5

Page 18: 457454 Physical Chemistry Quantum Chemistry

Solution: First of all we do not know if this wavefunction is normalized, so weshould assume that it isn’t. We could normalize this wavefunction, but we won’t.

We are interested in finding the probability that the object is outside of the region

−α < x < α. To do this using an unnormalized wavefunction we must evaluate

P (|x| > α) =

R −α−∞ |ψ(x)|

2 dx+R∞α|ψ(x)|2 dxR∞

−∞ |ψ(x)|2 dx

. (5)

The first integral in the numerator gives the probability that the object is at a

position x < −α and the second integral in the numerator gives the probabilityfor x > α. The denominator accounts for the fact that the wavefunction is un-

normalized. The limits of the integral in the denominator represent all space for

the object. If you were working with a normalized wavefunction the denominator

would be equal to 1 and hence not needed. Plugging in the wavefunctions we have

P (|x| > α) =

R −α−∞ e−2αx

2dx+

R∞α

e−2αx2dxR∞

−∞ e−2αx2dx. (6)

Mathematica can assist with these integrals to give the final answer of

P (|x| > α) = erfc[√2α

32 ]. (7)

Problem: A quantum object is described by the wavefunction ψ(x) = e−γx over

the range 0 ≤ x <∞. Normalize this wavefunction.

Solution: Following our general procedure from the notes if we have some unnor-malized wavefunction, ψunnorm we know that this function must simply be some

constant N multiplied by the normalized version of this function:

ψunnorm = Nψnorm (8)

We have shown generally that N is given by

N =

sZspace

|ψunnorm(x)|2 dx. (9)

6

Page 19: 457454 Physical Chemistry Quantum Chemistry

Which for this case is

N =

sZ ∞

0

|e−γx|2 dx =sZ ∞

0

e−2γxdx =

r1

2γ(10)

So finally we get the normalized wavefunction by rearanging ψunnorm = Nψnorm:

ψnorm(x) =p2γe−γx. (11)

Problem: A quantum object is described by the wavefunction ψ(x) = e−γx over

the range 0 ≤ x <∞. What is the average position of the object?

Solution: We need to work with the normalized wavefunction that we found inthe previous problem, ψ(x) =

√2γe−γx. Generally and average is calculated as

hoi =Zspace

ψ∗(x)oψ(x), (12)

which in this case is

hxi =Z ∞

0

p2γe−γxx

p2γe−γxdx = 2γ

Z ∞

0

xe−2γxdx =1

2γ. (13)

So on average you will find the object at x = 12γ.

Problem: What is the probability of finding an electron in the 1s state of hydrogenfurther than one Bohr radius away from the nucleus?

Solution: We need to evaluate

P (r > a0) =

Z 2π

0

Z π

0

Z ∞

a0

|ψ1s|2 r2 sin θdrdθdφ. (14)

Remember the extra r2 sin θ is needed when integrating in spherical polar coordi-

nates. The normalized 1s wavefunction is

ψ1s =1pπa30

er/a0 . (15)

7

Page 20: 457454 Physical Chemistry Quantum Chemistry

We can do this integral by hand or have Mathematica help us to give

P (r > a0) =5

e2= 0.677. (16)

So, about 68% of the time the electron would be found at some distance greater

then one Bohr radius from the proton.

Problem: A free particle in three dimensions is described by the Hamiltonian,

H = −~22m∇2. Express the wavefunction (in Cartesian coordinates) as a product

state.

Solution: This problem appears hard at first since we are not studying three

dimensional systems, but all it is asking is to express the wavefunction, which is

a function of the three spatial dimensions, Ψ(x, y, z) as a product state. We know

that if the wavefunction is to be a product state then the Hamiltonian must be

made up of a sum of independent terms. To see this we write out the Laplacian

to get

H =−~22m

µ∂2

∂x2+

∂2

∂y2+

∂2

∂z2

¶. (17)

We see that indeed the Hamiltonian is a sum of term that depends only on x,

a term depending only on y and a term that depends only on z. Therefore the

appropriate product state is

Ψ(x, y, z) = ψ(x)ψ(y)ψ(z). (18)

Problem: Expand the Morse potential in a Taylor’s series about Req. Verify that

the coefficient for the linear term is zero. What is the force constant associated

with the Morse potential?

Solution: The Morse potential is

V (x) = De

£1− e−β(R−Req)

¤. (19)

8

Page 21: 457454 Physical Chemistry Quantum Chemistry

The Taylor series about Req for this function is

V (x) = V (x)|Req| z = 0

+dV (x)

dx

¯Req| z

= 0

(R−Req) +1

2!

d2V (x)

dx2

¯Req| z

= β2De

(R−Req)2 + · · · . (20)

So, yes the coefficient of the linear term (the term involving (R − Req) to the

first power) is zero. This will always be true when you perform a Taylor series

expansion about a minimum (or maximum). The force constant is given by the

coefficient of the quadratic term so in this case k = β2De.

Problem: Without performing any calculations, compare hRi as a function ofthe vibrational quantum number for a diatomic modelled as a harmonic oscillator

versus a Morse oscillator.

Solution: This problem requires the we think qualitatively about the wavefunc-

tions and the potentials for the harmonic oscillator and the Morse oscillator. The

potential for the harmonic oscillator is described by a parabola centered about the

equilibrium bond length. Hence no mater what the vibrational quantum number is

there is just as much of the wavefunction on either side equilibrium thus hRi = Req

for any quantum number. The Morse potential does not have this symmetry. It

is steeper on the “short” side of equilibrium and softer on the “long” side of equi-

librium and this “softness” increases with increasing quantum number. Therefore

without performing any calculations we can at least say that hRi increases as thequantum number increases.

Problems Dealing With Statistical Mechanics and Thermo-

dynamics

Problem: A vial containing 1020 benzene molecules is at 300K. How many mole-cules are in the first excited state of the ‘ring breathing’ mode (992 cm−1)? How

9

Page 22: 457454 Physical Chemistry Quantum Chemistry

many are in the first excited state of the symmetric C—H vibrational mode (3063

cm−1)?

Solution: This is a problem that deals with the Boltzmann distribution. So,

N rbv=1 =

µ2 sinh

992

2× 208

¶×³e−

3×9922×208

´× 1020 = 8.41× 1017 (21)

and

NC—Hv=1 =

µ2 sinh

3063

2× 208

¶×³e−

3×30632×208

´× 1020 = 4.02× 1013. (22)

We see that about 8.41×1017

1020× 100% = 0.841% of the benzene molecules are in the

first vibrational excited state for the ring breathing mode and 4.02×10131020

× 100% =0.0000402% of the benzene molecules are in the first excited state for the C—H

stretching mode.

Problem: Consider a linear chain of N atoms. Each of the atoms can be in one

of three states A, B or C, except that an atom in state A can not be adjacent to

an atom in state C. Find the entropy per atom for this system as N → ∞. To

solve this problem it is useful to define the set of three dimensional column vectors

V (j) such that the three elements are the total number of allowed configurations of

a j-atom chain having the jth atom in state A, B or C. For example,

V (1) =

⎡⎢⎣ 111

⎤⎥⎦ , V (2) =

⎡⎢⎣ 232

⎤⎥⎦ , V (3) =

⎡⎢⎣ 575

⎤⎥⎦ , · · · . (23)

The V (j+1) can be found from the V (j) vector using the matrix equation,

V (j+1) =MV (j), (24)

where for this example

M =

⎡⎢⎣ 1 1 0

1 1 1

0 1 1

⎤⎥⎦ . (25)

10

Page 23: 457454 Physical Chemistry Quantum Chemistry

The matrix M is the so-called transfer matrix for this system. It can be shownthat the number of configurations W = Tr[MN ]. Now for large N, Tr[MN ] ≈ λNmax,

where λmax is the largest eigenvalue of M. So

W = limN→∞

λNmax. (26)

1. 1. Use M to find V (4)

2. Verify V (3) explicitly by drawing all the allowed 3-atom configurations.

3. Verify W = Tr[MN ] for N = 1 and N = 2.

4. Use Boltzmann’s equation to find the entropy per atom for this chain

as N goes to infinity.

Solution: For part (a) we simply use the transfer matrix as directed in the

problem (we are given V (3)):

V (4) =

⎡⎢⎣ 1 1 0

1 1 1

0 1 1

⎤⎥⎦⎡⎢⎣ 575

⎤⎥⎦ =⎡⎢⎣ 121712

⎤⎥⎦ .For part (b) we need to list all states for the case of N = 3 and verify the we get

the same result as calculated using the transfer matrix. Remembering that V (3)

gives us the number of sequences that end in a given state we should organize our

list in the same manner

States ending in A States ending in B States ending in C

AAA AAB ABC

ABA ABB BBC

BAA BAB BCC

BBA BBB CBC

CBA BCB

CBB

CCB

5 states√

7 states√

5 states√

.

11

Page 24: 457454 Physical Chemistry Quantum Chemistry

States like AAC are not allowed because A and C are neighbors.

For part (c) we evaluate W = Tr[MN ] for N = 1 and 2. For N = 1, W =

Tr[M ] = 3 This corresponds to the three distinguishable microstates A, B, and

C. For N = 2,

W = Tr[M2] = Tr

⎡⎢⎣⎡⎢⎣ 1 1 0

1 1 1

0 1 1

⎤⎥⎦⎡⎢⎣ 1 1 0

1 1 1

0 1 1

⎤⎥⎦⎤⎥⎦ = Tr

⎡⎢⎣⎡⎢⎣ 2 2 1

2 3 2

1 2 2

⎤⎥⎦⎤⎥⎦ = 7 (27)

This corresponds to the seven distinguishable microstates AA, AB, BA, BB, BC,

CB and CC (Remember C and A cannot be neighbors).

For part (d) we use

S

N=

k

NlnW = lim

N→∞

k

NlnλNmax = lim

N→∞

k

NN lnλmax = k lnλmax. (28)

So, we simply need to find the maximum eigenvalue of the Transfer matrix. Using

Mathematica we find λmax = 1 +√2. Therefore the limiting entropy per atom

isS

N= k ln

³1 +√2´. (29)

Problem: Using the classical theory of light scattering, calculate the positions ofthe Rayleigh, Stokes and anti-Stokes spectral lines for benzene. Assume benzene

has only two active modes (992cm−1 and 3063cm−1) and assume the Laser light

used to do the scattering is at 20000cm−1 (this is 500nm–green light).

Solution: Since there are two vibrational modes we expect two Stokes lines tothe red of 20000cm−1, one at 20000cm−1 − 992cm−1 = 19008cm−1 and one at

20000cm−1 − 3063cm−1 = 16937cm−1. Likewise we expect two anti-Stokes lines,one at 20000cm−1 + 992cm−1 = 20992cm−1 and one at 20000cm−1 + 3063cm−1 =

23063cm−1. There is only one Rayleigh line and it is at the same frequency at the

input laser beam which, in this case, is 20000cm−1.

12

Page 25: 457454 Physical Chemistry Quantum Chemistry

Problem: A simple model for a crystal is a “gas” of harmonic oscillators. De-termine A, S, and U from the partition function for this model.

Solution: For this model the crystal is modelled as a collection of harmonicoscillators so we need the partition function for the harmonic oscillator.

Qcrystal = qNHO =

Ã1

2 sinh β~ω2

!(30)

From our formulas for statistical thermodynamics

A = −kT lnQcrystal = +NKT ln

µ2 sinh

β~ω2

¶, (31)

where we used properties of logs to pull the N out front and move the sinh term

from to the numerator,

S = −kβ∂Qcrystal

∂β+ k lnQcrystal (32)

=Nkβ~ω2

cothβ~ω2− k ln

µ2 sinh

β~ω2

¶and

U = −∂Qcrystal

∂β=

N~ω2

cothβ~ω2

. (33)

Problem: Express the equation of state for internal energy for a Berthelot gas.

Solution: The equation representing a Berthelot gas is

P =nRT

V − nb− n2a

TV 2. (34)

We are interesting in an equation of state for U(T, V ). Writing out the total

derivative of U(T, V ) we get

dU =

µ∂U

∂T

¶V

dT +

µ∂U

∂V

¶T

dV. (35)

13

Page 26: 457454 Physical Chemistry Quantum Chemistry

Now¡∂U∂T

¢Vis just heat capacity, CV , but

¡∂U∂V

¢Tis nothing convenient so we must

proceed. We employ the “useful relation”µ∂U

∂V

¶T

= T

µ∂P

∂T

¶V

− P (36)

to eliminate U in favor of P so that we can use the equation of state for a Berthelot

gas. One obtains

T

µ∂P

∂T

¶V

− P = T

µnR

V − nb+

n2a

T 2V 2

¶− nRT

V − nb+

n2a

TV 2=2n2a

TV 2. (37)

Hence the equation of state for internal energy of a Berthelot gas is

dU = CV dT +2n2a

TV 2dV (38)

Problem: Use the identities for partial derivatives to eliminate the¡∂P∂T

¢Vfactor

in

Cp = Cv + T

µ∂V

∂T

¶P

µ∂P

∂T

¶V

(39)

so that all derivatives are at constant pressure or temperature.

Solution: Here we either remember an identity or turn to our handout of partialderivative identities to employ the cyclic rule to

¡∂P∂T

¢V:µ

∂P

∂T

¶V

= −µ∂P

∂V

¶T

µ∂V

∂T

¶P

. (40)

This eliminates the constant V term and so,

Cp = Cv − T

µ∂V

∂T

¶2P

µ∂P

∂V

¶T

. (41)

14

Page 27: 457454 Physical Chemistry Quantum Chemistry

Part I

Basic Quantum Mechanics

15

15

Page 28: 457454 Physical Chemistry Quantum Chemistry

1. Quantum Theory

The goal of science is unification.

• Many phenomena described by minimal and general concepts.

1.1. The “Fall” of Classical Physics

A good theory:

• explain known experimental results

• self consistent

• predictive

• minimal number of postulates

Around the turn of the century, experiments were being performed in which the re-

sults defied explanation by means of the current understanding of physics. Among

these experiments were

1. The photoelectric effect

2. Low temperature heat capacity

3. Atomic spectral lines

4. Black body radiation and the ultraviolet catastrophe

16

16

Page 29: 457454 Physical Chemistry Quantum Chemistry

5. The two slit experiment

6. The Stern-Gerlach experiment

∗ ∗ See Handouts ∗ ∗

1.2. Bohr’s Atomic Theory

1.2.1. First Attempts at the Structure of the Atom

The “solar system” model.

• The electron orbits the nucleus with the attractive coulomb force balancedby the repulsive centrifugal force.

Flaws of the solar system model

• Newton: OK √

• Maxwell: problem √

17

Page 30: 457454 Physical Chemistry Quantum Chemistry

— As the electron orbits the nucleus, the atom acts as an oscillating dipole

• — The classical theory of electromagnetism states that oscillating dipolesemit radiation and thereby lose energy.

— The system is not stable and the electron spirals into the nucleus. Theatom collapses!

Bohr’s model: Niels Bohr (1885—1962)

18

Page 31: 457454 Physical Chemistry Quantum Chemistry

• Atoms don’t collapse =⇒ what are the consequences

Experimental clues

• Atomic gases have discrete spectral lines.

• If the orbital radius was continuous the gas would have a continuous spec-trum.

• Therefore atomic orbitals must be quantized.

r =4π 0N

2~2

Zmee2(1.1)

where Z is the atomic number, me and e are the mass and charge of the

electron respectively and 0 is the permittivity of free space. N is a positive

real integer called the quantum number. ~ = h/2π is Planck’s constant

divided by 2π.

The constant quantity 4π 0~2mee2

appears often and is given the special symbol a0 ≡4π 0~2mee2

= 0.52918 Å and is called the Bohr radius.

The total energy of the Bohr atom is related to its quantum number

EN = −Z2µ

e2

2a0

¶1

N2. (1.2)

Tests of the Bohr atom

• Ionization energy of Hydrogen atoms

— The Ionization energy for Hydrogen atoms (Z = 1) is the minium

energy required to completely remove an electron form it ground state,

i.e., N = 1→ N =∞

Eionize = E∞ −E1 =−Z2e22a0

µ1

∞2− 1

12

¶=

e2

2a0(1.3)

19

Page 32: 457454 Physical Chemistry Quantum Chemistry

— Eionize =e2

2a0= 13.606 eV= 109,667 cm−1 =R. R is called the Rydberg

constant.

— Eionize experimentally observed from spectroscopy is 13.605 eV (very

good agreement)

• Spectroscopic lines fromHydrogen represent the difference in energy betweenthe quantum states

— Bohr theory: Difference energies

Ej −Ek =e2

2a0

µ1

N2j

− 1

N2k

¶= R

µ1

N2j

− 1

N2k

¶(1.4)

Initial state Nk Final States Nj Series Name

1 2,3,4,· · · Lyman

2 3,4,5,· · · Balmer

3 4,5,6,· · · Pachen

4 5,6,7,· · · Brackett

5 6,7,8,· · · Pfund

• — Since the orbitals are quantized, the atom may only change its orbital

radius by discrete amounts.

— Doing this results in the emission or absorption of a photon with energy

v =4E

hc(1.5)

Failure of the Bohr model

• No fine structure predicted (electron-electron coupling)

• No hyperfine structure predicted (electron-nucleus coupling)

• No Zeeman effect predicted (response of spectrum to magnetic field)

20

Page 33: 457454 Physical Chemistry Quantum Chemistry

• Spin is not included in theory

The Bohr quantization idea points to a wavelike behavior for the electron.

The wave must satisfy periodic boundary conditions much like a vibrating ring

∗ ∗ ∗ See Fig. 11.9 Laidler&Meiser ∗ ∗∗

The must be continuous and single valued

Particles have wave-like characteristics

The Bohr atom was an important step towards the formulation of quantum theory

• Erwin Schrödinger (1887—1961): Wave mechanics

• Werner Heisenberg (1902—1976): Matrix mechanics

• Paul Dirac (1902—1984): Abstract vector space approach

21

Page 34: 457454 Physical Chemistry Quantum Chemistry

2. The Postulates of Quantum

Mechanics

2.1. Postulate I

Postulate I: The state of a system is defined by a wavefunction, ψ, which con-

tains all the information that can be known about the system.

We will normally take ψ to be a complex valued function of time and coordi-

nates: ψ(t, x, y, z) and, in fact, we will most often deal with time independent

“stationary” states ψ(x, y, z)

Note: In general the wavefunction need not be expressed as a function of coordi-

nate. It may, for example, be a function of momentum.

The wavefunction ψ represents a probability amplitude and is not directly observ-

able.

However the mod-square of the wavefunction, ψ∗ψ = |ψ|2 , represents a probabilitydistribution which is directly observable.

That is, the probability of finding a particle which is described by ψ(x, y, z) at the

position between x and x+dx, y and y+dy and z and z+dz is |ψ(x, y, z)|2 dxdydz(or |ψ(r, θ, φ)|2 r2 sin θdrdθdφ in spherical coordinates).

22

22

Page 35: 457454 Physical Chemistry Quantum Chemistry

Properties of the wavefunction

• Single valueness

• continuous and finite

• continuous and finite first derivative

•Rspace |ψ(x, y, z)|

2 dxdydz <∞

Normalization of the wavefunction

In order for |ψ(x, y, z)|2 to be exactly interpreted as a probability dis-tribution, ψ(x, y, z) must be normalizable.

That is, ψunnorm = Nψnorm, whereN =qR

space |ψunnorm(x, y, z)|2 dxdydz

This assures thatRspace |ψnorm|

2 dxdydz = 1 as expected for a proba-

bility distribution

From now on we will always normalize our wavefunctions.

2.2. How to normalize a wavefunction

If we have some unnormalized wavefunction, ψunnorm we know that this function

must simply be some constant N multiplied by the normalized version of this

function:

ψunnorm = Nψnorm. (2.1)

Now, we take the mod-square of both sides and then integrate both sides of this

equation over all spaceZspace

|ψunnorm|2 dxdydz =Zspace

|Nψnorm|2 dxdydz. (2.2)

23

Page 36: 457454 Physical Chemistry Quantum Chemistry

The N is just a constant so it can be pulled out of both the mod-square and the

integral Zspace

|ψunnorm|2 dxdydz = N2

Zspace

|ψnorm|2 dxdydz, (2.3)

but Zspace

|ψnorm|2 dxdydz = 1 (2.4)

because that is the very definition of a normalized wavefunction. Thus wherever

we seeRspace |ψnorm|

2 dxdydz we can replace it with 1. So,Zspace

|ψunnorm|2 dxdydz = N2 × 1 = N2. (2.5)

This gives us an expression for N. Taking the square root of both sides gives.

N =

sZspace

|ψunnorm(x, y, z)|2 dxdydz. (2.6)

So finally we get the normalized wavefunction by reagranging ψunnorm = Nψnorm:

ψnorm =1

Nψunnorm. (2.7)

Notice that no where did we ever specify what ψunnorm or ψnorm actually were,

therefore this is a general procedure that will work for any wavefunction.

To find the probability for the particle to be in a finite region of space we simple

evaluate (here a 1D case)

P (x1 < x < x2) =

R x2x1|ψ(x)|2 dxR∞

−∞ |ψ(x)|2 dx

if ψ(x)=⇒

normalized

Z x2

x1

|ψ(x)|2 dx (2.8)

2.3. Postulates II and II

Postulate II: Every physical observable is represented by a linear (Hermitian)operator.

24

Page 37: 457454 Physical Chemistry Quantum Chemistry

An operator takes a function and turns it into another function

Of(x) = g(x) (2.9)

This is just like how a function takes a number and turns it into another number.

So in quantum mechanics operators act on the wavefunction to produce a new

wavefunction

The two most important operators as far as we are concerned are

• x = x

• px = −i~ ∂∂x

and of course the analogous operators for the other coordinates (y, z) and coordi-

nate systems (spherical, cylindrical, etc.).

Nearly all operators we will need are algebraic combinations of the above.

Postulate III: The measurement of a physical observable will give a result thatis one of the eigenvalues of the corresponding operator.

There is a special operator equation called the eigenvalue equation which is

Of(x) = λf(x) (2.10)

where λ is just a number.

For a given operator only a special set of function satisfy this equation. These

functions are called eigenfunctions.

25

Page 38: 457454 Physical Chemistry Quantum Chemistry

The number that goes with each function is called the eigenvalue.

So solution of the eigenvalue equation gives a set of eigenfunctions and a set of

eigenvalues.

Example

Let O in the eignevalue equation be the operator that takes the derivative: O =

d = ddx.

So we want a solution to

df(x) = λf(x) (2.11)df(x)

dx= λf(x)

So, we ask ourselves what function is proportional to its own derivative? ⇒f(x) = eλx.

So the eigenfunctions are the set of functions f(x) = eλx and the eigenvalues are

the numbers λ

26

Page 39: 457454 Physical Chemistry Quantum Chemistry

3. The Setup of a Quantum

Mechanical Problem

3.1. The Hamiltonian

The most important physical observable is that of the total energy E.

The operator associated with the total energy is called the Hamiltonian operator

(or simply the Hamiltonian) and is given the symbol H.

The eigenvalue equation for the Hamiltonian is

Hψ = Eψ. (3.1)

This equation is the (time independent) Schrödinger equation.

This equation is the most important equation of the course and we will use it many

times throughout our discussion of quantum mechanics and statistical mechanics.

3.2. The Quantum Mechanical Problem

Nearly every problem one is faced with in elementary quantum mechanics is han-

dled by the same procedure as given in the following steps.

1. Define the classical Hamiltonian for the system.

27

27

Page 40: 457454 Physical Chemistry Quantum Chemistry

• The total energy for a classical system is

Ecl = T + V, (3.2)

where T is the kinetic energy and V is the potential energy.

• The kinetic energy is always of the form

T =1

2m

¡p2x + p2y + p2z

¢(3.3)

• The potential energy is almost always a function of coordinates only

V = V (x, y, z) (3.4)

• Note: Some quantum systems don’t have classical analogs so the Hamil-tonian operator must be hypothesized.

2. Use Postulate II to replace the classical variables, x, px etc., with their

appropriate operators. Thus,

T =−~22m∇2 = −~

2

2m∇2, (3.5)

where ∇2 ≡ ∂2

∂x2+ ∂2

∂y2+ ∂2

∂z2, and

V = V (x, y, z) = V (x, y, z). (3.6)

So,

H = T + V =−~22m∇2 + V (x, y, z) (3.7)

3. Solve the Schrödinger equation, Hψ = Eψ, which is now a second order

differential equation of the form∙−~22m∇2 + V (x, y, z)

¸ψ = Eψ

⇒ −~22m∇2ψ + (V (x, y, z)−E)ψ = 0 (3.8)

28

Page 41: 457454 Physical Chemistry Quantum Chemistry

• Note: It is solely the form of V (x, y, z) which determines whether this

is easy or hard to do.

• For one-dimensional problems

−~22m

d2

dx2ψ + (V (x)−E)ψ = 0 (3.9)

3.3. The Average Value Theorem

Postulate III implies that if ψ is an eigenfunction of a particular operator rep-

resenting a physical observable, then all measurements of that physical property

will yield the associated eigenvalue.

However, If ψ is not an eigenfunction of a particular operator, then all measure-

ments of that physical property will still yield an eigenvalue, but we cannot predict

for certain which one.

We can, however, give an expectation, or average, value for the measurement.

This is given by

hαi =Zspace

ψ∗αψdxdydz (3.10)

For example,

hxi =Zspace

ψ∗xψdxdydz =

Zspace

x |ψ|2 dxdydz (3.11)

and

hpxi =Zspace

ψ∗pxψdxdydz = −i~Zspace

ψ∗∂ψ

∂xdxdydz (3.12)

29

Page 42: 457454 Physical Chemistry Quantum Chemistry

3.4. The Heisenberg Uncertainty Principle

In quantum mechanics certain pairs of variables can not, even in principle, be

simultaneously known to arbitrary precision. Such variables are called compli-

mentary.

This idea is the Heisenberg uncertainty principle and is of profound im-portance.

The general statement of the Heisenberg uncertainty principle is

δαδβ ≥ 12

¯Dhα, β

iE¯, (3.13)

where the notationhα, β

imeans the commutator of α and β. The commutator is

defined as hα, β

i≡ αβ − βα. (3.14)

The most important example of complimentary variables is position and momen-

tum. We see

δpxδx ≥1

2|h[px, x]i| =

1

2|hpxx− xpxi| (3.15)

=1

2

¯Zψ∗~i

µ∂

∂xx− x

∂x

¶ψdx

¯=

¯~2i

¯=~2.

So, at the very best we can only hope to simultaneously know position and momen-

tum such that the product of the uncertainty in each is ~2. (n.b., δpxδy = 0, we can

know, for example, the y position and the x momentum to arbitrary precision.)

Suppose we know the position of a particle perfectly, what can we say about its

momentum?

30

Page 43: 457454 Physical Chemistry Quantum Chemistry

4. Particle in a Box

We now will apply the general program for solving a quantum mechanical problem

to our first system: the particle in a box.

This system is very simple which is one reason for beginning with it. It also can

be used as a “zeroth order” model for certain physical systems.

We shall soon see that the particle in a box is a physically unrealistic system and,

as a consequence, we must violate one of our criteria for a good wavefunction.

Nevertheless it is of great pedagogical and practical value.

4.1. The 1D Particle in a Box Problem

Consider the potential, V (x), shown in the figure and given by

V (x) =

⎧⎪⎨⎪⎩∞ x ≤ 00 0 < x < a

∞ x ≥ a

. (4.1)

Because of the infinities at x = 0 and x = a, we need to partition the x-axis into

the three regions shown in the figure.

31

31

Page 44: 457454 Physical Chemistry Quantum Chemistry

Now, in region I and III, where the potential is infinite, the particle can never

exist so, ψ must equal zero in these regions.

The particle must be found only in region II.

The Schrödinger equation in region II is (V (x) = 0)

Hψ = Eψ =⇒ −~2

2m

d2ψ(x)

dx2= Eψ, (4.2)

which can be rearranged into the form

d2ψ(x)

dx2+2mE

~2ψ(x) = 0. (4.3)

The general solution of this differential equation is

ψ(x) = A sin kx+B cos kx, (4.4)

where k =q

2mE~2 .

Now ψ must be continuous for all x. Therefore it must satisfy the boundary

conditions (b.c.): ψ(0) = 0 and ψ(a) = 0.

32

Page 45: 457454 Physical Chemistry Quantum Chemistry

From the ψ(0) = 0 b.c. we see that the constant B must be zero because

cos kx|x=0 = 1.

So we are left with ψ(x) = A sin kx for our wavefunction.

As can be inferred from the following figure, the second b.c., ψ(a) = 0, places

certain restrictions on k.

In particular,

kn =nπ

a, n = 1, 2, 3, · · · . (4.5)

The values of k are quantized. So, now we have

ψn(x) = A sinnπx

a. (4.6)

The constant A is the normalization constant. We obtain A fromZ ∞

−∞ψ∗n(x)ψn(x) = 1 =

Z a

0

A2 sinnπx

asin

nπx

adx. (4.7)

Letting u = πxa, du = π

adx, this becomes

1 = A2a

π

Z π

0

sin2 nudu = A2a

π/π/2=

A2a

2. (4.8)

33

Page 46: 457454 Physical Chemistry Quantum Chemistry

Solving for A gives

A =

r2

a. (4.9)

Thus our normalized wavefunctions for a particle in a box are

ψn(x) =

⎧⎪⎪⎨⎪⎪⎩0 Iq

2asin nπx

aII

0 III

. (4.10)

Is this wavefunction OK?

We can get the energy levels from kn =q

2mEn~2 and kn =

nπa:

En =n2π2~2

2ma2~= h

2π=n2h2

8ma2. (4.11)

4.2. Implications of the Particle in a Box problem

Zero Point Energy

34

Page 47: 457454 Physical Chemistry Quantum Chemistry

The smallest value for n is 1 which corresponds to an energy of

E1 =h2

8ma26= 0. (4.12)

That is, the lowest energy state, or ground state, has nonzero energy. This residual

energy is called the zero point energy and is a consequence of the uncertainty

principle.

If the energy was zero then we would conclude that momentum was exactly zero,

δp = 0. But we also know that the particle is located within a finite region of

space, so δx 6=∞.

Hence, δxδp = 0 which violates the uncertainty principle.

Features of the Particle in a Box Energy Levels

• The energy level spacing is

4E = En+1 − En =(n+ 1)2h2

8ma2− n2h2

8ma2= (n2/ + 2n+ 1− n2/ ) h2

8ma2

4E = (2n+ 1)h2

8ma2(4.13)

• This spacing increases linearly with quantum level n

• This spacing decreases with increasing mass

• This spacing decreases with increasing a

• It is this level spacing that is what is measured experimentally

The Curvature of the Wavefunction

35

Page 48: 457454 Physical Chemistry Quantum Chemistry

The operator for kinetic energy is T = −~22m

d2

dx2. The important part of this is d2

dx2.

From freshman calculus we know that the second derivative of a function describes

its curvature so, a wavefunction with more curvature will have a larger second

derivative and hence it will posses more kinetic energy.

This is an important concept for the qualitative understanding of wavefunctions

for any quantum system.

Applying this idea to the particle in a box we an anticipate both zero point energy

and the behavior of the energy levels with increasing a.

• We know the wavefunction is zero in regions I and III. We also know thatthe wave function is not zero everywhere. Therefore it must do something

between x = 0 and x = a. It must have some curvature and hence some zero

point energy.

• As a is increased, the wavefunction is less confined and so the curvature doesnot need to be as great to satisfy the boundary conditions. Therefore the

energy levels decrease in energy as does their difference.

The particle in a box problem illustrates some of the many strange features of

quantum mechanics.

We have already seen such nonclassical behavior as quantized energy and zero

point energy.

As another example consider the expectation value of position for a particle in

the second quantum level:

hxi =Z ∞

−∞ψ∗2(x)xψ2(x)dx =

2

a

Z a

0

x sin2[2π

ax]dx =

a

2(4.14)

36

Page 49: 457454 Physical Chemistry Quantum Chemistry

yet the probability of finding the particle at x = a2is zero: ψ2(

a2) = 0. There is

a node at x = a2. So even though the particle may be found anywhere else in the

box and it may get from the left side of the node to the right side, it can never

be found at the node.

37

Page 50: 457454 Physical Chemistry Quantum Chemistry

5. The Harmonic Oscillator

The harmonic oscillator model which is simply a mass undergoing simple harmonic

motion. The classical example is a ball on a spring

The harmonic oscillator is arguably the single most important model in all of

physics.

We shall begin by reviewing the classical harmonic oscillator and than we will

turn our attention to the quantum oscillator.

The force exerted by the spring in the above figure is F = −k(R−Req), where k

is the spring constant and Req is the equilibrium position of the ball.

Setting x = R − Req we can measure the displacement about the equilibrium

position.

38

38

Page 51: 457454 Physical Chemistry Quantum Chemistry

From Newton’s law of motion F = ma = md2xdt2

, we get

md2x

dt2= −kx⇒ d2x

dt2+

k

mx = 0 (5.1)

This is second order differential equation which we already know the solutions to:

x = A sinωt+B cosωt, (5.2)

where ω =q

kmand A and B are constants which are determined by the initial

conditions.

For quantum mechanics it is much more convenient to talk about energy rather

than forces, so in going to the quantum oscillator, we need to express the force of

the spring in terms of potential energy V . We know

V = −Z

Fdx =1

2kx2 + C. (5.3)

Since energy is on an arbitrary scale we can set C = 0. Thus V = 12kx2.

By postulate III the Schrödinger equation becomes

Hψ = Eψ ⇒

⎛⎜⎝−~22m

d2

dx2| z K.E.

+1

2kx2| z P.E.

⎞⎟⎠ψ = Eψ. (5.4)

This can be rearrange into the form

−~22m

d2ψ

dx2+

µ1

2kx2 − E

¶ψ = 0 (5.5)

This differential equation is not easy to solve (you can wait to solve it in graduate

school).

39

Page 52: 457454 Physical Chemistry Quantum Chemistry

The equation is very close to the form of a know differential equation called Her-

mite’s differential equation the solutions of which are called the Hermite polynom-

inals.

As it turns out, the solutions (the eigenfunctions) to the Schrödinger equation for

the harmonic oscillator are

ψn(y) = AnHn(y)e− y2

2 , y =

µkm

~2

¶ 14

x, An =1p

2nn!√π, (5.6)

where An is the normalization constant for the nth eigenfunction and Hn(y) are

the Hermite polynomials.

The eigenvalues (the energy levels) are

En = (n+1

2)~ω, (5.7)

where again ω =q

km.

Note the energy levels are often written as

En = (n+1

2)hν0, (5.8)

where ν0 = 12π

qkmand is called the vibrational constant.

∗ ∗ ∗ See Fig. 11.12 Laidler&Meiser ∗ ∗∗

5.1. Interesting Aspects of the Quantum Harmonic Oscilla-

tor

It is interesting to investigate some of the unintuitive properties of the oscillator

as we have gone quantum mechanical

40

Page 53: 457454 Physical Chemistry Quantum Chemistry

1. Consider the ground state (the lowest energy level)

• There is residual energy in the ground state because

E0 = (0 +1

2)~ω.

• Just like for the particle in a box, this energy is called the zero pointenergy.

• It is a consequence of uncertainty principle

— If the ground state energy was really zero, then we would concludethat the momentum of the oscillator was zero.

— On the other hand, we would conclude the particle was located atthe bottom of the potential well (at x = 0)

— Thus we would have δp = 0, δx = 0, so δpδx = 0 Not allowed!

— The uncertainty principle forces there to be some residual zeropoint energy.

2. Consider the wavefunctions.

• The wavefunctions penetrate into the region where the classical particleis forbidden to go

— The wavefunction is nonzero past the classical turning point.

• The probability distribution |ψ|2 becomes more and more like what isexpected for the classical oscillator when v →∞.

— This is a manifestation of the correspondence principle whichstates that for large quantum numbers, the quantum system must

behave like a classical system. In other words the quantum me-

chanics must contain classical mechanics as a limit.

3. Interpretation of the wavefunctions and energy levels

41

Page 54: 457454 Physical Chemistry Quantum Chemistry

• Remember the wavefunctions are time independent and the energy lev-els are stationary

• If a molecule is in a particular vibrational state it is NOT vibrating.

5.2. Spectroscopy (An Introduction)

The primary method of measuring the energy levels of a material is through the

use of electromagnetic radiation.

Experiments involving electromagnetic radiation—matter interaction are called

spectroscopies.

Atoms and molecules absorb or emit light only at specific (quantized) energies.

These specific values correspond to the energy level difference between the initial

and final states.

42

Page 55: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 1

Listed here are some of the key equations for Exam 1. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• The short cut for getting the normalization constant (1D, see above for 3D).

N =

sZspace

|ψunnorm(x)|2 dx. (5.9)

• The normalized wavefunction:

ψnorm =1

Nψunnorm. (5.10)

• The Schrödinger equation (which should be posted on your refrigerator),

Hψ = Eψ. (5.11)

43

43

Page 56: 457454 Physical Chemistry Quantum Chemistry

• The Schrödinger equation for 1D problems as a differential equation,

−~22m

d2

dx2ψ + (V (x)−E)ψ = 0. (5.12)

• How to get the average value for some property (1D version),

hαi =Zspace

ψ∗αψdx. (5.13)

• The momentum operator

px = −i~∂

∂x. (5.14)

• Normalized wavefunctions for the 1D particle in a box,

ψn(x) =

r2

asin

nπx

a. (5.15)

• The energy levels for the 1D particle in a box,

En =n2π2~2

2ma2~= h

2π=n2h2

8ma2. (5.16)

• The energy level spacing for the 1D particle in a box,

4E = (2n+ 1)h2

8ma2(5.17)

• The wavefunctions for the harmonic oscillator are

ψn(y) = AnHn(y)e−y2

2 , y =

µkm

~2

¶ 14

x, An =1p

2nn!√π, (5.18)

where An is the normalization constant for the nth eigenfunction and Hn(y)

are the Hermite polynomials.

• The energy levels are

En = (n+1

2)~ω, ω =

rk

m(5.19)

44

Page 57: 457454 Physical Chemistry Quantum Chemistry

Part II

Quantum Mechanics of Atomsand Molecules

45

45

Page 58: 457454 Physical Chemistry Quantum Chemistry

6. Hydrogenic Systems

Now that we have developed the formalism of quantum theory and have discussed

several important systems, we move onto the quantum mechanical treatment of

atoms.

Hydrogen is the only atom for which we can exactly solve the Schrödinger equation

for. So this will be the first atomic system we discuss.

The Schrödinger equation for all the other atoms on the periodic table must be

solved by approximate methods.

6.1. Hydrogenic systems

Hydrogenic systems are those atomic systems which consist of a nucleus and one

electron. The Hydrogen atom (one proton and one electron) is the obvious exam-

ple

Ions such as He+ and Li2+ are also hydrogenic systems.

These system are centrosymmetric. That is they are completely symmetric about

the nucleus.

The obvious choice for the coordinate system is to use spherical polar coordinates

46

46

Page 59: 457454 Physical Chemistry Quantum Chemistry

with the origin located on the nucleus.

The classical potential energy for these hydrogenic systems is

V (r) =−Ze2(4π 0)r

. (6.1)

So the Hamiltonian is

H =−~22me∇2 + −Ze2

(4π 0)r. (6.2)

Schrödinger’s equation (in spherical polar coordinates) becomes

Eψ = Hψ (6.3)

Eψ =

µ−~22me∇2 + −Ze2

(4π 0)r

¶ψ

Eψ =

µ−~22me

∙1

r2∂

∂rr2

∂r+1

r2

µ1

sin θ

∂θsin θ

∂θ+

1

sin2 θ

∂2

∂φ2

¶¸+−Ze2(4π 0)r

¶ψ

The Hamiltonian is (almost) the sum of a radial part (only a function of r) and

an angular part (only a function of θ and φ):

H = Hrad +1

r2Hang, (6.4)

Hrad =−~22me

∙1

r2∂

∂rr2

∂r− Ze2

(4π 0)r

¸(6.5)

and

Hang =−~22me

µ1

sin θ

∂θsin θ

∂θ+

1

sin2 θ

∂2

∂φ2

¶(6.6)

Since the Hamiltonian is the sum of two terms, ψ must be a product state.

ψ(r, θ, φ) = ψrad(r)ψang(θ, φ) (6.7)

It turns out that solving the Schrödinger equation,

Hangψang(θ, φ) = Eψang(θ, φ), (6.8)

47

Page 60: 457454 Physical Chemistry Quantum Chemistry

yields

ψang(θ, φ) = Ylm(θ, φ), (6.9)

where the Ylm(θ, φ)’s are the spherical harmonic functions characterized by quan-

tum numbers l andm. The spherical harmonics are known functions. (Mathematica

knows them and you can use them just like any other built-in function like sine

or cosine.)

We shall use the spherical harmonics more next semester when we develop the

quantum theory of angular momentum.

It also turns out that the energy associated with Hang is found to be

E = El =l(l + 1)~2

2me. (6.10)

So,

Hangψang(θ, φ) =l(l + 1)~2

2meψang(θ, φ) (6.11)

Now let’s denote the radial part of the wavefunction as ψrad(r) = R(r).

The full Schrödinger equation becomes

Hψ(r, θ, φ) = Eψ(r, θ, φ) (6.12)

HR(r)Ylm(θ, φ) = ER(r)Ylm(θ, φ)µHrad +

1

r2Hang

¶R(r)Ylm(θ, φ) = ER(r)Ylm(θ, φ),

Operating with Hang we getµHrad +

l(l + 1)~2

2mer2

¶R(r)Ylm(θ, φ) = ER(r)Ylm(θ, φ) (6.13)

48

Page 61: 457454 Physical Chemistry Quantum Chemistry

The Ylm(θ, φ) can now be cancelled to leave a one dimensional differential equation:

−~22me

µ1

r2∂

∂rr2

∂r− Ze2

4π 0r− l(l + 1)

r2

¶R(r) = ER(r). (6.14)

This differential equation is very similar to a known equation called Laguerre’s

differential equation which has as solutions the Laguerre polynomials Lln(x).

In fact, the solutions to our differential equation are closely related to the Laguerre

polynomials.

Rnl(σ) = Anl

µ2σ

n

¶l

e−σ/nL2l+1n+1

µ2σ

n

¶, (6.15)

where the normalization constant, Anl, depends on the n and l quantum numbers

as

Anl = −

sµ2Z

na0

¶3(n− l − 1)!2n[(n+ l)!]3

(6.16)

The energy eigenvalues, i.e., the energy levels are given by

En = −Z2Rn2

(6.17)

Note: The energy levels are determined by n alone–l drops out.

Also Note: the energy levels are the same as for the Bohr model.

So, the total wavefunction that describes a hydrogenic system (ignoring the spin

of the electron, which will be briefly discussed later) is

ψnlm(r, θ, φ) = Rnl(r)Ylm(θ, φ) (6.18)

6.2. Discussion of the Wavefunctions

We are now very close to having the atomic orbitals familiar from freshman chem-

istry.

49

Page 62: 457454 Physical Chemistry Quantum Chemistry

We have explicitly derived the “physicists” picture of the atomic orbitals

orbital n l m wavefunctions (σ = r/a0)

1s 1 0 0 ψ1s = ψ100 = e−σ

2s 2 0 0 ψ2s = ψ200 =¡1− σ

2

¢e−σ/2

2p 2 1 0 ψ2p0 = ψ210 = σe−σ/2 cos θ

2 1 ±1 ψ2p±1 = ψ21±1 = σe−σ/2 sin θe±iφ

3d 3 2 0 ψ3d0 = ψ320 = σ2e−σ/3 (3 cos2 θ − 1)3 2 ±1 ψ3d±1 = ψ32±1 = R32(r) cos θ sin θe

±iφ

3 2 ±2 ψ3d±2 = ψ32±2 = R32(r) sin2 θe±i2φ

The wavefunctions in the “physicists” picture are complex (they have real and

imaginary components). The wavefunctions that chemists like are pure real. So

one needs to form linear combinations of these orbitals such that these combina-

tions are pure real.

The atomic orbital you are used to from freshman chemistry are the “chemists”

picture of atomic orbitals

In the above table ψ1s, ψ2s, ψ2p0, ψ3d0 are pure real and so these are the same in

the “chemists” picture as in the “physicists” picture.

The table below lists the atomic orbitals in the “chemists” picture as linear com-

binations of the “physicists” picture wave functions.

50

Page 63: 457454 Physical Chemistry Quantum Chemistry

orbital n l m wavefunctions (σ = r/a0)

1s 1 0 0 ψ1s = ψ1s

2s 2 0 0 ψ2s = ψ2s

2p 2 1 0 ψpz = ψ2p02 1 ±1 ψ2px =

1√2

£ψ2p1 + ψ2p−1

¤2 1 ±1 ψ2py =

1i√2

£ψ2p1 − ψ2p−1

¤3d 3 2 0 ψ3dz2 = ψ3d0

3 2 ±1 ψ3dxz =1√2

£ψ3d1 + ψ3d−1

¤3 2 ±1 ψ3dyz =

1i√2

£ψ3d1 − ψ3d−1

¤3 2 ±2 ψ3dxy =

1√2

£ψ3d2 + ψ3d−2

¤ψ3dx2−y2 =

1i√2

£ψ3d2 − ψ3d−2

¤6.3. Spin of the electron

As we know from freshman chemistry, electrons also posses an intrinsic quantity

called spin.

Spin is actually rather peculiar so we will put off a more detailed discussion until

next semester.

For now we must be satisfied with the following:

• There are two quantum numbers associated with spin: s and ms

• s is the spin quantum number and for an electron s = 1/2 (always).

• ms is the spin orientation quantum number and ms = ±1/2 for electrons.

The spin wavefunction is a function in spin space not the usual coordinate space,

so we can not write down an explicit function of the coordinate space variables.

51

Page 64: 457454 Physical Chemistry Quantum Chemistry

We simply denote the spin wavefunction generally as χs,msand “tack it on” as

another factor of the complete wavefunction.

When a particular spin state is needed a further notation is commonly used:

α ≡ χ 12, 12(the “spin-up” state) and β ≡ χ 1

2,−1

2(the “spin-down” state)

6.4. Summary: the Complete Hydrogenic Wavefunction

We are now in position to fully describe all properties of hydrogenic systems

(except for relativistic effects)

The full wave function is

Ψn,l,m,s,ms = ψn,l,mχs,ms(6.19)

= Rnl(r)Yl,m(θ, φ)χ

The energy is given by

En = −Z2Rn2

, (6.20)

where recall. Again note that for a free hydrogenic system the total energy depends

only on the principle quantum number n.

The quantum numbers of the hydrogenic system

• The principle quantum number, n: determines the total energy of the sys-

tems and the atomic shells.

— The principle quantum number, n, can take on values of 1,2,3. . .

• The angular momentum quantum numbers, l: determines the total angular

momentum of the system. It also determines the atomic sub-shells

52

Page 65: 457454 Physical Chemistry Quantum Chemistry

— The angular momentum quantum number, l, can take on values of 0,

1, . . . (n− 1)

— For historical reasons l = 0 is called s, l = 1 is called p, l = 2 is called

d, l = 3 is called f etc.

• The orientation quantum number, m: determine the projection of the an-

gular momentum onto the z-axis. It also determines the orientation of the

atomic sub-shells

— The magnetic quantum number, m, can take on values of 0, ±1, . . .± l.

• The spin quantum number, s: determines the total spin angular momentum.

— For electrons s = 1/2.

• The spin orientation quantum number, ms: determines the projection of the

spin angular momentum onto the z-axis (i.e., spin-up or spin-down).

— For electrons ms = ±1/2

We have accomplished quite a bit. We have determined all that we can about the

hydrogen atom within Schrödinger’s theory of quantum mechanics.

This is not the full story however. The Schrödinger theory is a non-relativistic

one; that is, it can not account for relativistic effects which show up in spectral

data. We also had to add spin in an ad hoc manner to account for what we know

experimentally–spin did not fall out of the theory naturally.

Dirac, in the late 1920’s, developed a relativistic quantum theory in which the

well established phenomenon of spin arose naturally. His theory also made the

53

Page 66: 457454 Physical Chemistry Quantum Chemistry

bold prediction of the existence anti-matter that has now been verified time and

again.

The Dirac theory was still not fully complete, because there still existed exper-

imental phenomena that was not properly described. In 1948 Richard Feynman

developed the beginnings of quantum electrodynamics (QED). QED is the best

theory ever developed in terms of matching with experimental data.

Both the relativistic Dirac theory and QED are beyond our reach, so we limit

ourselves to the non-relativistic Schrödinger theory.

54

Page 67: 457454 Physical Chemistry Quantum Chemistry

7. Multi-electron atoms

7.1. Two Electron Atoms: Helium

We now consider a system consisting of two electrons and a nucleus; for example,

helium.

Although the extension from hydrogen to helium seems simple it is actually ex-

tremely complicated. In fact, it is so complicated that it can’t be solved exactly.

The helium atom is an example of the “three-body-problem”–difficult to handle

even in classical mechanics–one can not get a closed form solution.

The Hamiltonian for helium is

H = − ~2

2me∇21| z

K.E of electron 1

− ~2

2me∇22| z

K.E of electron 2

− Ze2

4π 0r1| z P.E of electron 1

− Ze2

4π 0r2| z P.E of eletcron 2

+e2

4π 0r12| z elec.—elec. repulsion

, (7.1)

where r12 = |r1 − r2| is the distance between the electrons.

The electron—electron repulsion term is responsible for the difficulty of the prob-

lem. It makes a closed form solution impossible.

The problem must be solved by one of the following methods

• Numerical solutions (we will not discuss this)

55

55

Page 68: 457454 Physical Chemistry Quantum Chemistry

• Perturbation theory (next semester)

• Variational theory (next semester)

• Ignore the electron—electron repulsion (good for qualitative work only)

7.2. The Pauli Exclusion Principle

Electron are fundamentally indistinguishable. They can not truly be la-belled.

All physical properties of a system where we have labelled the electrons as, say, 1

and 2 must be exactly the same as when the electrons are labelled 2 and 1.

Now, only |ψ|2 is directly measurable–not ψ itself.

All this implies that

ψ(1, 2) =

⎧⎪⎨⎪⎩+ψ(2, 1), symmetric

or

−ψ(2, 1) antisymmetric

(7.2)

The Pauli exclusion principle states: The total wavefunctions for fermions(e.g., electrons) must be antisymmetric under the exchange of indistinguishable

fermions.

Note: a similar statement exists for bosons (e.g., photons): The total wavefunction

for bosons must be symmetric under exchange of indistinguishable bosons.

Let us consider the two electron atom, helium

56

Page 69: 457454 Physical Chemistry Quantum Chemistry

The total wavefunction is

Ψ = ψ(1, 2)χ(1, 2) (7.3)

Since a complete solution for helium is not possible we must use approximate

wavefunctions. Since we are doing this, we may as well simplify matters and use

product state wavefunctions (products of the hydrogenic wavefunctions).

Ψ = ψ(1)ψ(2)| z spatial part

χ(1)χ(2)| z spin part

, (7.4)

where the single particle wavefunctions are that of the hydrogenic system.

The Pauli exclusion principle implies that if the spatial part is even with respect

to exchange then the spin part must be odd. Likewise if the spatial part is odd

then the spin part must be even.

Now let’s blindly list all possibilities for the ground state wave function of helium

Ψa = ψ1s(1)α(1)ψ1s(2)α(2) (7.5)

Ψb = ψ1s(1)α(1)ψ1s(2)β(2)

Ψc = ψ1s(1)β(1)ψ1s(2)α(2)

Ψd = ψ1s(1)β(1)ψ1s(2)β(2)

These appear to be four reasonable ground state wavefunctions which would im-

ply a four-fold degeneracy. However considering the symmetry with respect to

exchange we see the following

• Ψa has symmetric spatial and spin parts and is there for symmetric. It must

be excluded.

• Similarly for Ψd.

• Ψb and Ψc have symmetric spatial parts, but the spin part is neither sym-

metric or antisymmetric. So, one must make an antisymmetric linear com-

bination of the spin parts.

57

Page 70: 457454 Physical Chemistry Quantum Chemistry

The appropriate linear combination is

α(1)β(2)− α(2)β(1). (7.6)

So the ground state wave function for helium is

Ψg = ψ1s(1)ψ1s(2) [α(1)β(2)− α(2)β(1)] . (7.7)

Consequences of the Pauli exclusion principle

• No two electrons can have the same five quantum numbers

• Electrons occupying that same subshell must have opposite spins

7.3. Many Electron Atoms

The remaining atoms on the periodic table are handled in a manner similar to

helium.

Namely the wavefunction is product state that must be antisymmeterized in ac-

cordance with the Pauli exclusion principle.

The product wavefunction for the ground state is determined by applying the

aufbau principle. The aufbau principle states that the ground state wavefunction

is built-up of hydrogenic wavefunctions

To arrive at an antisymmetric wavefunction we construct the Slater determinant :

Ψ =

¯¯¯ψ1s(1)α(1) ψ1s(1)β(1) · · · ψn(1)α(1) ψn(1)β(1)

ψ1s(2)α(2) ψ1s(2)β(2) · · · ψn(2)α(2) ψn(2)β(2)...

......

......

ψ1s(N)α(N) ψ1s(N)β(N) ψn(N)α(N) ψn(N)β(N)

¯¯¯ (7.8)

58

Page 71: 457454 Physical Chemistry Quantum Chemistry

The reason one can be sure that this wavefunction is the antisymmeterized is that

we know from linear algebra that the determinant is antisymmetric under exchange

of rows (corresponds to exchanging two electrons). It is also antisymmetric under

exchange of columns.

Another property of the determinant is that if two rows are the same (corresponds

to two electrons in the same state) the determinant is zero. This agrees with the

Puli exclusion principle.

As an example consider lithium:

• There are three electrons so we need three hydrogenic wavefunctions: ψ1sα,ψ1sβ, and ψ2sα (or ψ2sβ).

• We construct the Slater determinant as

Ψ1 =

¯¯ ψ1s(1)α(1) ψ1s(1)β(1) ψ2s(1)α(1)

ψ1s(2)α(2) ψ1s(2)β(2) ψ2s(2)α(2)

ψ1s(3)α(3) ψ1s(3)β(3) ψ2s(3)α(3)

¯¯ (7.9)

or

Ψ2 =

¯¯ ψ1s(1)α(1) ψ1s(1)β(1) ψ2s(1)β(1)

ψ1s(2)α(2) ψ1s(2)β(2) ψ2s(2)β(2)

ψ1s(3)α(3) ψ1s(3)β(3) ψ2s(3)β(3)

¯¯ (7.10)

• The short hand notation for these states is (1s)2(2s)1

7.3.1. The Total Hamiltonian

The total Hamiltonian for a many electron (ignoring spin-orbit coupling which

will be discussed next semester) atom is

H =NXi=1

"−~22me∇2i −

Ze2

4π 0ri+Xj>i

e2

4π 0rij

#(7.11)

59

Page 72: 457454 Physical Chemistry Quantum Chemistry

8. Diatomic Molecules and the Born

Oppenheimer Approximation

Now that we have applied quantum mechanics to atoms, we are able to begin the

discussion of molecules.

This chapter will be limited to diatomic molecules.

8.1. Molecular Energy

A diatomic molecule with n electrons requires that 3n+6 coordinates be specified.

Three of these describe the center of mass position.

3n of these describe the position of the n electrons.

This leaves three degrees of freedom (R, θ, φ) which describe the position of the

nuclei relative to the center of mass. R determines the internuclear separation

and θ and φ determine the orientation.

60

60

Page 73: 457454 Physical Chemistry Quantum Chemistry

8.1.1. The Hamiltonian

In the center of mass coordinates the Hamiltonian for a diatomic molecule is

H = TN + Te + VNN + VNe + Vee. (8.1)

TN is the nuclear kinetic energy operator and is given by

TN = −~2

2μ∇2N = −

~2

2μR2∂

∂RR2

∂R+~2

2μJ2, (8.2)

where J is angular momentum operator for molecular rotation and μ = m1m2

m1+m2is

the reduced mass of the diatomic molecule.

Te =P

i− ~22me∇2ei is the kinetic energy operator for the electrons.

VNN =ZAZBee

2

4π 0Ris the nuclear—nuclear potential energy operator.

VNe = −P

i

hZAe

2

4π 0rAi+ ZBe

2

4π 0rBi

iis the nuclear—electron potential energy operator.

Vee =P

i>je2

4π 0rjiis the electron—electron potential energy operator.

61

Page 74: 457454 Physical Chemistry Quantum Chemistry

8.1.2. The Born—Oppenheimer Approximation

The Born—Oppenheimer approximation: The nuclei move much slower thanthe electrons. (classical picture)

We put the Born—Oppenheimer approximation to work by first defining an effec-

tive Hamiltonian

Heff = Te + VNN + VNe + Vee. (8.3)

The approximation comes in by treating R as a parameter rather than an operator

(or variable). So one writes

Heffψe(R, ri) = Ee(R)ψe(R, ri). (8.4)

ψe is the so-called electronic wavefunction.

Now the Schrödinger equation for the diatomic molecule is³TN + Heff

´ψ(R, ri) = Eψ(R, ri). (8.5)

Since the Hamiltonian is a sum of two terms, one can write the wavefunction

ψ(R, ri) as a product wavefunction

ψ = ψNψe, (8.6)

where ψN is the so-called nuclear wavefunction.

Substituting the product wavefunction into the Schrödinger equation gives³TN + Heff

´ψNψe = EψNψe (8.7)³

TN +Ee(R)´ψNψe/ = EψNψe/³

TN +Ee(R)´ψN = EψN .

62

Page 75: 457454 Physical Chemistry Quantum Chemistry

The last equation is exactly like a Schrödinger equation with a potential equal to

Ee(R).

One now models Ee(R) or determines it experimentally.

8.2. Molecular Vibrations

As stated earlier R is the internuclear separation and θ and φ determine the

orientation. Consequently, R is the variable involved with vibration whereas θ

and φ are involved with rotation.

Considering only the R part of the Hamiltonian (under the Born—Oppenheimer

approximation), we have∙− ~

2

∂2

∂R2+Ee(R)

¸ψvib = Evibψvib. (8.8)

It is convenient at this point to expand Ee(R) in a Taylor series about the equi-

librium position, Req:

Ee(R) = E0 +

µ∂E

∂R

¶Req

(R−Req) +1

2!

µ∂2E

∂R2

¶Req

(R−Req)2 + · · · . (8.9)

Now E0 is just a constant which, by choice of the zero of energy, can be set to an

arbitrary value.

Since we are at a minimum,¡∂E∂R

¢Req

must be zero, so the linear term vanishes.

One defines³∂2E∂R2

´Req≡ ke as the force constant.

The remaining terms in the expansion can collective be defined as O[(R−Req)3] ≡Vanh, the anharmonic potential.

63

Page 76: 457454 Physical Chemistry Quantum Chemistry

As a first approximation we can neglect the anharmonicity. With this, the Schrödinger

equation becomes∙− ~

2

∂2

∂R2+1

2ke(R−Req)

2

¸ψvib = Evibψvib. (8.10)

If we let x = (R−Req) this becomes∙− ~

2

∂2

∂x2+1

2kex

2

¸ψvib = Evibψvib, (8.11)

which is exactly the harmonic oscillator equation. Hence

ψvib,n = AnHn(√αx)e−αx

2/2, (8.12)

where α ≡q

keμ~ .

And

Evib,n = hcωe(n+1

2), (8.13)

where ωe ≡ 12π

qkeμ.

8.2.1. The Morse Oscillator

Neglecting anharmonicity and using the harmonic oscillator approximation works

well for low energies. However, it is a poor model for high energies.

For high energies we need a more realistic potential–one that will allow of bond

dissociation.

The Morse potential

Ee(R) = De[1− e−β(R−Req )]2, (8.14)

64

Page 77: 457454 Physical Chemistry Quantum Chemistry

where De is the well depth and β = 2πcωe

qμ2De

is the Morse parameter. Note:

this expression for the Morse potential has the zero of energy at the bottom of

the well (i.e. R = Req, ;Ee(Req) = 0).

The Morse Potential can also be written as

Ee(R) = De[e−2β(R−Req ) − 2e−β(R−Req )]. (8.15)

Now the zero of energy is the dissociated state (i.e. R→∞, ;Ee(R→∞) = 0).

We approach this quantum mechanical problem exactly like all the other.

The Schrödinger equation is∙− ~

2

∂2

∂R2+De[1− e−β(R−Req )]2

¸ψvib = Evibψvib (8.16)

This is another differential equation that is difficult to solve.

As it turns out, this Schrödinger equation can be transformed into a one of a broad

class of known differential equations called confluent hypergeometric equations–

the solutions of which are the confluent hypergeometric functions, 1F1.

Doing this yields the wavefunctions of the form

ψvib,n(z) = zApne−z1F1(−n, 1 + 2Apn, 2z), (8.17)

z =

√2Deμ

βhe−βx,

A =

√2μ

βh,

pn =pDe +

−12− n

A

and energy levels of the form

Evib,n = −De + hcωe(n+1

2)− hcωexe(n+

1

2)2, (8.18)

65

Page 78: 457454 Physical Chemistry Quantum Chemistry

where ωexe together is the anharmonicity constant, with xe =hcωe4De

.

∗ ∗ ∗ See Handout ∗ ∗∗

8.2.2. Vibrational Spectroscopy

Infrared (IR) and Raman spectroscopy are the two most widely used techniques

to probe vibrational levels.

The spectral peaks appear at v = 4Ehc(in units of wavenumbers, cm−1).

The transition from the n = 0 to the n = 1 state is called the fundamental

transition.

Transitions from n = 0 to n = 2, 3, 4 · · · are called overtone transitions.

Transitions from n = 1 to 2, 3, 4 · · · , n = 2 to 3, 4, 5 · · · , etc. are called hottransitions (or hot bands)

Since the energy levels depend on mass, isotopes will have a different transition

energy and hence appear in a different place in the spectrum. Heavier isotopes

have lower transition energies.

66

Page 79: 457454 Physical Chemistry Quantum Chemistry

9. Molecular Orbital Theory and

Symmetry

9.1. Molecular Orbital Theory

One of the most important concepts in all of chemistry is the chemical bond.

In freshman chemistry we learn of one model for chemical bonding–VSEPR (va-

lence shell electron-pair repulsion) theory, where hybridized atomic orbitals deter-

mine the bonding geometry of a given molecule.

We are now prepared to discuss a bonding theory that is more rigorously based

in quantum mechanics.

Basically we will treat the molecules in the same way as all our other quantum

mechanical problems (e.g., particle in a box, harmonic oscillator, etc.)

As you might expect, it is not possible to obtain the exact wavefunctions and

energy levels so, we must settle for approximate solutions.

As a first example, let us consider the molecular hydrogen ion H+2 .

The Hamiltonianfor H+2 is

H = TN + Tel + VNel + VNN (9.1)

67

67

Page 80: 457454 Physical Chemistry Quantum Chemistry

We use the Born-Oppenheimer approximation and treat the nuclear coordinates

as a parameters rather than as variables. So we only worry about parts of the

Hamiltonian that deal with the electron.

The effective Hamiltonian becomes

H = Tel + VNel (9.2)

=−~22me∇2 − e2

4π 0rA− e2

4π 0rB.

The eigenfunctions of this Hamiltonian are called molecular orbitals.

The molecular orbitals are the analogues of the atomic orbitals.

• Atomic orbitals: Hydrogen is the prototype and all other atomic orbitalsare built from the hydrogen atomic orbitals.

• Molecular orbitals: The hydrogen molecular ion is the prototype and allother molecular orbitals are built from the hydrogen molecular ion molecular

orbitals.

There is one significant difference between the above, which is the hydrogen atomic

orbitals are exact whereas the hydrogen molecular ion molecular orbitals are not

exact.

In fact, we shall see that these molecular orbitals are constructed as linear com-

binations of atomic orbitals.

9.2. Symmetry

Let the atoms of the hydrogen molecular ion lie on the z-axis of the center of mass

coordinate system.

68

Page 81: 457454 Physical Chemistry Quantum Chemistry

Inversion symmetry

• The potential field of the hydrogen molecular ion is cylindrically symmetricabout the z-axis.

• Because of the symmetry the electron density at (x, y, z) must equal theelectron density at (−x,−y,−z).

• The above symmetry therefore requires that the molecular orbitals be eigen-functions of the inversion operator, ı. That is

ıψ(x, y, z) = ψ(−x,−y,−z) = aψ(x, y, z). (9.3)

• Moreover the eigenvalue a can be either +1 or −1.

• If a = +1 the molecular wavefunction is even with respect to inversion andis called gerade and labelled with a “g”: ıψg = ψg

• If a = −1 the molecular wavefunction is odd with respect to inversion andis called ungerade and labelled with a “u”: ıψu = −ψu

• The terms gerade and ungerade apply only to systems that posses inversionsymmetry.

Cylindrical symmetry

69

Page 82: 457454 Physical Chemistry Quantum Chemistry

• The cylindrical symmetry implies that the potential energy can not dependon the φ.

• The molecular wavefunction is described by an eigenvalue λ = 0,±1,±2, . . .

— We use λ to label the molecular orbitals as shown in the table

λ 0 ±1 ±2 · · ·label σ π δ · · ·

Mirror plane symmetry

70

Page 83: 457454 Physical Chemistry Quantum Chemistry

• There is also a symmetry about the x-y plane called horizontal mirror planesymmetry: operator σh.

• Thus the molecular wavefunction must be an eigenfunction of σh with eigen-value ±1.

— If the eigenvalue is +1 (even with respect to σh) the molecular orbitalis called a bonding orbital.

— If the eigenvalue is −1 (odd with respect to σh) the molecular orbitalis called an antibonding orbital.

• There are also vertical mirror plane symmetries, but we will put that dis-cussion off for the time being.

71

Page 84: 457454 Physical Chemistry Quantum Chemistry

10. Molecular Orbital Diagrams

10.1. LCAO–Linear Combinations of Atomic Orbitals

Now that we know what symmetry the molecular orbitals must posses, we need

to find some useful approximations for them.

Useful can mean qualitatively useful or quantitatively useful.

Unfortunately we can’t have both.

We will discuss the approximation which models the molecular orbitals as linear

combinations of atomic orbitals (LCAO).

LCAO is qualitatively very useful but it lacks quantitative precision.

Let us again consider the hydrogen molecular ion H+2 : let one H atom be labelled

A and the other labelled B.

Linear combination of the 1s atomic orbital from each H atom is used for the

molecular orbital of H+2 :

(1sA) = ke−rA/a0 (10.1)

and

(1sB) = ke−rB/a0 (10.2)

72

72

Page 85: 457454 Physical Chemistry Quantum Chemistry

We construct two molecular orbitals as

Φ+ = C+(1sA + 1sB) (10.3)

and

Φ− = C−(1sA − 1sB) (10.4)

The normalization condition is ZΦ±Φ±dΩ = 1 (10.5)

As can be seen from the above figure, Φ+ represents a situation in which the

electron density is concentrated between the nuclei and thus represents a bonding

orbital.

Conversely Φ− represents a situation in which the electron density is very low

between the nuclei and thus represents an antibonding orbital

10.1.1. Classification of Molecular Orbitals

With atoms we classified atomic orbitals according to angular momentum.

For molecular orbitals we shall also classify them according to angular momentum.

But we shall also classify them according to their inversion symmetry and wether

or not they are bonding or antibonding.

73

Page 86: 457454 Physical Chemistry Quantum Chemistry

The classification according to angular momentum is as follows.

λ 0 ±1 ±2 · · ·orbital symbol σ π δ · · ·

Atomic orbitals with m = 0 form σ type molecular orbitals, e.g., s⇒ σ, pz ⇒ σ.

Those with m = ±1 form π type molecular orbitals, e.g., px ⇒ π etc.

The classification according to inversion symmetry is simply a subscript “g” or

“u”. For example, σg or σu etc.

The classification according to bonding or antibonding is an asterisk is used to

denote antibonding. For example, σg is a bonding orbital and σ∗u is an antibonding

orbital.

10.2. The Hydrogen Molecule

Let us now consider the hydrogen molecule. This molecules is a homonuclear

diatomic with two electrons.

If the two atoms are infinitely far apart. The ground state of the system would

consist of two separate hydrogen molecules in their ground atomic states: (1s)1

74

Page 87: 457454 Physical Chemistry Quantum Chemistry

As the atom are brought closer together, their respective s orbitals begin to over-

lap.

It is now more appropriate to speak in terms of molecular orbitals, so one forms

linear combinations of the atomic orbitals.

There are two acceptable linear combinations. These are

σg = 1sA + 1sB (10.6)

and

σ∗u = 1sA − 1sB. (10.7)

75

Page 88: 457454 Physical Chemistry Quantum Chemistry

It can be shown mathematically that the energy level associated with σg is lower

than σ∗u.

We can intuit this qualitatively however since the σ∗u orbital must have a node

whereas the σg does not.

It is also to be expected since we know H2 is a stable molecule.

10.3. Molecular Orbital Diagrams

The energy levels associated with the molecular orbitals are drawn schematically

is what is called a molecular orbital diagram.

The molecular orbital diagram for H2 is shown below

Molecular orbital diagrams can be drawn for any molecule. Some get very compli-

cated. We will focus on the second row homonuclear diatomics and some simple

heteronuclear diatomics.

76

Page 89: 457454 Physical Chemistry Quantum Chemistry

The molecular orbital diagrams for the second row homonuclear diatomics are

rather simple.

∗ ∗ ∗ See Supplement ∗ ∗∗

The supplement that follows this section contains examples for each of the second

row diatomics.

Heteronuclear diatomics are some what more complicated since there is a disparity

in the energy levels of the atomic orbitals for the separated atoms. This disparity

is not present for homonuclear diatomics.

A consequence of this energy level disparity is that molecular orbitals may be

formed from nonidentical atomic orbitals. For example a high lying 1s orbital

may combine with a low lying 2s orbital to form a σ molecular orbital.

The supplement that follows this section contains some examples of heteronuclear

diatomics.

Bond order

• One important property that can be predicted from the molecular orbital

diagrams is bond order.

• Bond order is defined as

BO =1

2(# of bonding electrons−# of antibonding electrons) (10.8)

• Examples follow in the supplement.

77

Page 90: 457454 Physical Chemistry Quantum Chemistry

10.4. The Complete Molecular Hamiltonian and Wavefunc-

tion

We have discussed molecular vibrations which under the Born-Oppenheimer ap-

proximation are governed by the vibrational Hamiltonian and described by the

vibrational wavefunction.

Likewise we have discussed molecular orbitals which are the electronic wavefunc-

tions.

Next semester we will discuss molecular rotations and just like for vibrations

and electronic transitions they are governed by the rotational Hamiltonian and

described by the rotational wavefunction.

We can succinctly express the Schrödinger equation for a molecule as follows.

(Next semester will we look at the details of this for polyatomic molecules)

HmolΨmol = EmolΨmol (10.9)³Hele + Hvib + Hrot

´ψeleψvibψrot = (Eele +Evib +Erot)ψeleψvibψrot

78

Page 91: 457454 Physical Chemistry Quantum Chemistry

11. An Aside: Light Scattering–Why

the Sky is Blue

This chapter addresses the topic of light scattering from two different perspectives.

• Classical electrodynamics

• Classical statistical mechanics

Since this is not a course on electrodynamics, we have to take several key results

from that theory on faith.

11.1. The Classical Electrodynamics Treatment of Light Scat-

tering

As usual we work under the electric dipole approximation and only focus on the

interaction of the electric field part of light with a dipole.

When the light interacts with the molecule an electric dipole is induced according

to

μ = αE, (11.1)

where α is the polarizability of the molecule describing the “flexibility” of its

electron cloud.

79

79

Page 92: 457454 Physical Chemistry Quantum Chemistry

For light, the electric field part is

E(t) = E0 cosωt. (11.2)

The polarizability also depends on the positions of nuclei to some degree. That

is, there is a vibrational (and rotational) contribution to the polarizability:

α(t) = α0 + α1 cosωvt (11.3)

(here for simplicity we assume only one vibrational mode).

Thus the light—matter interaction is described as

μ(t) = α(t)E(t) = (α0 + α1 cosωvt)E0 cosωt (11.4)

= α0E0 cosωt+ α1E0 cosωvt cosωt

= α0E0 cosωt| z Rayleigh

+α1E02

⎡⎣cos(ω − ωv)t| z Stokes Raman

+ cos(ω + ωv)t| z AntiStokes Raman

⎤⎦where a trig identity was used in the last step.

According to classical electrodynamics an oscillating dipole emits an electromag-

netic field at the oscillation frequency.

In this case we see the dipole oscillates at three distinct frequencies: ω, ω − ωv

and ω + ωv as part of three terms in the above expression.

The first term corresponds to Rayleigh scattering where the scattered light is at

the same frequency as the incident light.

The second term corresponds to Stokes Raman scattering where the scattered

light is shifted to the red of the incident frequency.

80

Page 93: 457454 Physical Chemistry Quantum Chemistry

The third term corresponds to anti-Stokes Raman scattering where the scattered

light is shifted to the blue of the incident frequency.

Classical electrodynamics can describe exactly how the oscillating electric dipole

emits electromagnetic radiation. It can be shown that the emitted intensity is

I =ω4

3c3μ20, (11.5)

where μ0 = α0E0 for the case of Rayleigh scattering and μ0 = α1E0/2 for the case

of Raman scattering.

To explicitly derive this expression we would need a fair bit of electrodynamics

and so the derivation is not shown here.

The important point to note is that I ∝ ω4 or alternatively I ∝ 1/λ4. There is avery strong dependence on frequency (or wavelength).

This quartic scattering dependence is, in fact, the reason why the sky is blue (from

the point of view of classical electrodynamics) and is called the Rayleigh scattering

law.

11.2. The Blue Sky

The spectrum of visible light from the sun incident on the outer atmosphere is

essentially flat as shown below.

81

Page 94: 457454 Physical Chemistry Quantum Chemistry

We just learned that light scatters as it traverses the atmosphere according to

Rayleigh’s scattering law: I(λ) ∝ 1/λ4.

The following figures illustrate why Rayleigh scattering implies that the sky is

blue.

11.2.1. Sunsets

We have focused on a blue sky, but red sunsets occur for the same reason–

Rayleigh scattering.

82

Page 95: 457454 Physical Chemistry Quantum Chemistry

If we look directly at the sun during a sunset (or sunrise) it appears red because

most of the blue light has scattered in other directions.

This more pronounced at dawn or dusk since the light must traverse more of the

atmosphere at those times then at noonday at which time the sun appears yellow

in color.

11.2.2. White Clouds

Wemight expect that clouds should be highly colored since they consist of droplets

of water which scatter light very effectively.

83

Page 96: 457454 Physical Chemistry Quantum Chemistry

The key difference between light scattering by clouds versus by the atmosphere is

the size of the scatterer.

The water droplets are much larger than the wavelenght of the light–quite the

opposite case as above.

In this limit an entirely different analysis is made–one does not have Rayleigh

scattering but instead has a process called Mie scattering.

In some contexts, particularly in liquid suspensions, Mie scattering is referred to

as Tyndall scattering

84

Page 97: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 2

Listed here are some of the key equations for Exam 2. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• The wavefunctions for the hydrogenic system are

ψnlm(r, θ, φ) = Rnl(r)Ylm(θ, φ) (11.6)

• The radial part is.

Rnl(σ) = Anl

µ2σ

n

¶l

e−σ/nL2l+1n+1

µ2σ

n

¶, (11.7)

where the normalization constant, Anl, depends on the n and l quantum

numbers as

Anl = −

sµ2Z

na0

¶3(n− l − 1)!2n[(n+ l)!]3

(11.8)

85

85

Page 98: 457454 Physical Chemistry Quantum Chemistry

• The energy levels for the hydrogenic system are given by

En = −Z2Rn2

(11.9)

• The wavefunctions for the harmonic oscillator are

ψn(y) = AnHn(y)e− y2

2 , y =

µkm

~2

¶ 14

x, An =1p

2nn!√π, (11.10)

where An is the normalization constant for the nth eigenfunction and Hn(y)

are the Hermite polynomials.

• The energy levels are

En = (n+1

2)~ω, ω =

rk

m(11.11)

• The Morse potential is

Ee(R) = De[1− e−β(R−Req )]2, (11.12)

where De is the well depth and β = 2πcωe

qμ2De

is the Morse parameter.

Note: this expression for the Morse potential has the zero of energy at the

bottom of the well (i.e. R = Req, ;Ee(Req) = 0).

• The Morse Potential can also be written as

Ee(R) = De[e−2β(R−Req ) − 2e−β(R−Req )]. (11.13)

Now the zero of energy is the dissociated state (i.e. R→∞, ;Ee(R→∞) =0).

• The energy levels for the Morse oscillator are of the form

Evib,n = −De + hcωe(n+1

2)− hcωexe(n+

1

2)2, (11.14)

where ωexe together is the anharmonicity constant, with xe =hcωe4De

.

86

Page 99: 457454 Physical Chemistry Quantum Chemistry

• Bond order is defined as

BO =1

2(# of bonding electrons−# of antibonding electrons) (11.15)

• The Rayleigh scattering law is

I(λ) ∝ 1/λ4 ∝ ω4 (11.16)

87

Page 100: 457454 Physical Chemistry Quantum Chemistry

Part III

Statistical Mechanics and TheLaws of Thermodynamics

88

88

Page 101: 457454 Physical Chemistry Quantum Chemistry

12. Rudiments of Statistical

Mechanics

When we study simple systems like a single molecule, we use a very detailed

theory, quantum mechanics.

However, most of the time in the real world we are dealing with macroscopic

systems, say, at least 100 million molecules.

It is simply impossible, even with the fastest computers, to write down and solve

the Schrödinger equation for those 100 million molecules, but often Avogadro’s

number of molecules.

So we need a less detailed theory called statistical mechanics, which allows one to

handle macroscopic sized systems without losing to much of the rigor.

12.1. Statistics and Entropy

Probability and statistics is at the heart of statistical mechanics.

We will need some definitions

• Ensemble: A large collection of equivalent macroscopic systems. The sys-tems are the same except that each one is in a different so-called microstate.

89

89

Page 102: 457454 Physical Chemistry Quantum Chemistry

• Microstate: The single particular state of one member of the ensemble givenby listing the individual states of each of the microscopic systems in the

macroscopic state.

• Configuration: The collection of all equivalent microstates. The number ofpossible configurations is defined as W.

Boltzmann developed an equation to connect the microscopic properties of an

ensemble to the macroscopic properties. The Boltzmann equation is

S = k lnW (12.1)

Where S is entropy and k is Boltzmann’s constant.

12.1.1. Combinations and Permutations

Consider a random system that when measured can appear in one of two outcomes

(e.g., flipping coins).

One valuable piece of statistical information about system is knowing how many

different ways the system appears p times in, say, outcome 1 after N measure-

ments.

This is given by the mathematical formula for combinations

C(N, p) =N !

p!(N − p)!. (12.2)

The number C(N, p) is also called the binomial coefficient because it gives the

coefficient for the pth order term in the expansion

(1 + x)N =NXp=0

C(N, p)xp. (12.3)

90

Page 103: 457454 Physical Chemistry Quantum Chemistry

This formula will allow us to derive a normalization constant so that we can obtain

the probability of obtaining p measurements of state 1.

Set x = 1 in the above. This gives

(1 + 1)N =NXp=0

C(N, p)(1)p (12.4)

2N =NXp=0

C(N, p).

So the probability of any one outcome of N measurements is

P (N, p) =1

2NC(N, p) =

1

2NN !

p!(N − p)!(12.5)

For combinations we did not care what order the results of the measurements

occurred.

Sometimes the order is important.

So rather than a particular combination, we are interested in a particular permu-

tation. This is given by

W (N, Ni) =N !

N1!N2!N3! · · ·(12.6)

where N is the total number of measurements and Ni is the number of indistin-

guishable results of type i.

∗ ∗ ∗ See Examples on Handout ∗ ∗∗

For both combinations and permutations we need to evaluated factorials.

91

Page 104: 457454 Physical Chemistry Quantum Chemistry

This is no problem for small numbers, but when we consider macroscopic systems

(1020 or so molecules) no calculator can handle factorials of such large numbers.

Sterlings Approximation:

• In place of evaluating factorials of large number one can use Sterling’s ap-proximation to approximate the value of the factorial.

• Sterling’s approximation is

ln(N !) ' N lnN −N (12.7)

12.2. Fluctuations

When we list the macroscopic properties of a material such as a beaker of benzene

or the air of the atmosphere, we speak of the average value of the property.

Macroscopic equilibrium is a dynamic rather than static equilibrium. Conse-

quently, the value of a certain property fluctuates about the average value. Often

this fluctuation is not important, but sometimes it is important.

The fluctuation about an average value for any observable property O is described

by the variance which is defined as

σ2O ≡ O2 − O2. (12.8)

σO is consider the range of the observable property.

It can be shown thatσOO≈ 1√

N, (12.9)

where N is the number of particles. So for example if N = 1024 then 1√N= 10−12

92

Page 105: 457454 Physical Chemistry Quantum Chemistry

For ensembles having large numbers of particles measured values of a property are

extremely sharply peaked about the average value.

93

Page 106: 457454 Physical Chemistry Quantum Chemistry

13. The Boltzmann Distribution

Consider a isolated system of N molecules that has the set i energy levelsassociated with it.

Since the system is isolated the total energy, E, and the total number of particles

will be constant.

The total energy is given by

E =Xi

Ni i, (13.1)

where Ni is the number of particles in energy state i.

The total number of particles is, of course,

N =Xi

Ni (13.2)

The number of configurations for the system is then given by the number of

distinct permutations of the system

W =N !

N1!N2! · · ·. (13.3)

A system in equilibrium always tries to maximize entropy and minimize energy

and so the equilibrium configuration is a compromise between these two cases.

94

94

Page 107: 457454 Physical Chemistry Quantum Chemistry

For the moment let us relax the isolation constraint.

Maximizing entropy corresponds to maximizing W (via S = k lnW ). This would

be the situation in which every particle was in a different energy state. That is

all Ni = 1 or 0.

Minimizing energy would be the case where all the particles are in the ground

state (say 1).

These two situations are contradictory and some compromise must be obtained.

We start by considering our original system–that being one with constant energy,

E and number of particles N

To determine the equilibrium configuration we must find the maximumW subject

to the constraint of constant energy and constant number of particles.

This is done using the mathematical technique of Lagrange multipliers (page 951

of your calc book).

We will not discuss this method in detail and consequently we cannot derive the

equilibrium configuration.

The derivation using Lagrange multipliers arrives at the configuration in which

the

Ni = Ngie

−β iPj gje

−β j| z pi

, (13.4)

where β ≡ 1kTand gj denotes the degeneracy of states having energy j.

95

Page 108: 457454 Physical Chemistry Quantum Chemistry

The pi represents the probability of finding the a randomly chosen particle or

system which has energy i. This is the Boltzmann distribution

Pi =gie

−β iPj gje

−β j(13.5)

Since we started with a isolated system, β and hence T are constants. A given

energy E will correspond to a unique temperature T.

The analysis readily generalizes to variable energy i.e., nonisolated systems by

considering T as a variable.

13.1. Partition Functions

We have already come across both the partition functions that we will use in this

class.

The first is W–the number of configurations. This is called the microcanonical

partition function.

This partition function is not very useful to us so we will not discuss it further.

The second partition function is

Q =Xj

gje−βEj (13.6)

and is called the canonical partition function.

96

Page 109: 457454 Physical Chemistry Quantum Chemistry

This was first encountered as the denominator of the Boltzmann distribution and

it is extremely important in statistical mechanics. (Note: the symbol Z is also

often used for the canonical partition function.)

The partition function is to statistical mechanics as the wavefunction is to quan-

tum mechanics. That is, the partition function contains all that can be known

about the ensemble.

We shall see in the next chapter that the partition function will provide a link

between the microscopic (quantum mechanics or classical mechanics) and the

macroscopic (thermodynamics).

In fact we have already seen this in the S = k lnW. But this an inconvenient

connection because, for among other reasons, energy levels and temperature do

not explicitly appear.

There are other partition functions that are useful in different situations but we

will do nothing more than list two important ones here: i) the grand canonical

partition function and ii) the isothermal—isobaric partition function

13.1.1. Relation between the Q and W

When we get to connecting quantummechanics with thermodynamics it will prove

convenient to use Boltzmann’s equation (S = k lnW ) but as was stated earlier it

is not convenient to use the microcanonical partition function (W ).

In the following we give an argument which provides a relation between the par-

tition functions. It is not an exact relation as we derive it, but it is a very good

approximation for large numbers of particles.

97

Page 110: 457454 Physical Chemistry Quantum Chemistry

The microcanonical partition function describes a system at fixed energy E. In

fact W is the number of available states of the ensemble at the particular energy

E. This is essentially the same as the degeneracy of the ensemble gE.

Conversely the canonical partition function describes a system with variable en-

ergy.

However, based on our previous discussion of fluctuations, even though the energy

of the ensemble is allowed to vary, the number of states with energy equal to the

average energy E is overwhelmingly large. That is, almost every state available

to the ensemble has energy E.

We can express these ideas mathematically to come up with a relation between

W and Q.

The canonical partition function is

Q =Xj

gje−β j , (13.7)

but to a good approximation

Q ' gEe−βE. (13.8)

Now since the degeneracy is essentially the microcanonical partition function we

have

Q 'We−βE. (13.9)

So the canonical partition function is a Boltzmann weighted version of the micro-

canonical partition function.

We will soon make use of the Boltmann’s equation in terms of the canonical

98

Page 111: 457454 Physical Chemistry Quantum Chemistry

partition function:

lnQ ' ln(We−βE) = lnW + ln(e−βE) (13.10)

= lnW|zS/k

− E

kT.

so,

S = k lnQ+E

T(13.11)

13.2. The Molecular Partition Function

We ended the previous chapter by stating the total molecular energy (about the

center of mass) as

= ele + vib + rot. (13.12)

This is a consequence of the Born Oppenheimer approximation

If we include the center of mass translational motion this is

= ele + vib + rot + trans (13.13)

The ith total energy level is

i = ele,n + vib,v + rot,J + trans,m. (13.14)

Now if we have a collection of molecules in a macroscopic system. A given con-

figuration (say, configuration j) of that system has total energy Ej.

So the canonical partition function is

Q =Xj

gje−βEj (13.15)

99

Page 112: 457454 Physical Chemistry Quantum Chemistry

But, each Ej is made up of the contributions of all of the molecules:

Ej =al +

bm +

cn + · · · (13.16)

The partition function for the molecule is written as

Q =Xj

gje−βEj =

Xl,m,n···

(gal gbmg

cn · · · )e−β(

al +

bm+

cn+··· ) (13.17)

=Xl

gal e−β a

l| z qmol,a

Xm

game−β a

m| z qmol,b

Xn

gane−β a

n| z qmol,c

· · ·

where the qmol,i are the molecular partition functions.

The total canonical partition function is the product of the molecular partition

functions.

For the case where the molecules are the same then all the qmol,i are the same:

qmol,i = qmol thus

Q =qNmolN !

. (13.18)

This allows us to focus only on a single molecule:

qmol =Xi

gie−β i =

Xn,v,J,m

gele,ngvib,vgrot,Jgtrans,me−β( ele,n+ vib ,v+ rot,J+ trans,m)(13.19)X

n

gele,ne−β ele ,n

| z qele

Xv

gvib,ve−β vib,v

| z qvib

XJ

grot,Je−β rot,J

| z qrot

Xm

gtrans,me−β trans,m

| z qtrans

We now collect below the expression for each of these partition functions. You

will get the chance to derive each of these for your home work

100

Page 113: 457454 Physical Chemistry Quantum Chemistry

The Translational Partition Function

qtrans =V

Λ3(13.20)

where

Λ ≡ h√2πmkT

(13.21)

is the thermal de Broglie wavelength.

The Rotational Partition Function (linear molecules)

We will discuss rotations next semester.

However, the high temperature limit, which works for all gases (of linear molecules)

except H2 is

qrot ≈T

σθr(13.22)

where θr ≡ h2

8π2Ik(I is the moment of inertia) and σ is the so-called symmetry

number in which σ = 1 for unsymmetrical molecules and σ = 2 for symmetrical

molecules.

The Vibrational Partition Function

qvib =e−

12β~ω

1− e−β~ω=

1

2 sinh 12β~ω

(13.23)

Note this is for the harmonic oscillator. At temperatures well below the dissocia-

tion energy this is a very good approximation. (You will derive this as a homework

problem.)

The Electronic Partition Function

There is usually only a very few electronic states of interest. Only at exceedingly

high temperatures does any state other that the ground state(s) become important

101

Page 114: 457454 Physical Chemistry Quantum Chemistry

so

qele =Xi

gele,ie−β tele ,i ≈ gele,ground (13.24)

102

Page 115: 457454 Physical Chemistry Quantum Chemistry

14. Statistical Thermodynamics

The partition function allows one to calculate ensemble averages which correspond

to macroscopically measurable properties such as internal energy, free energy,

entropy etc.

In this chapter we will obtain expressions for internal energy, U, pressure, P,

entropy, S, and Helmholtz free energy, A. With these quantities in hand we will,

in the subsequent chapters, formally develop thermodynamics with no need to

refer back to the partition function.

Ensemble averages

The ensemble average of any property is given by

O =1

Q

Xi

Oigie−β i . (14.1)

Internal energy

One critical property of an ensemble is the average (internal) energy U.

U ≡ E =1

Q

Xi

igie−β i . (14.2)

Let us look closer at the above expression. Recall that

Q =Xi

gie−β i . (14.3)

103

103

Page 116: 457454 Physical Chemistry Quantum Chemistry

Now take the derivative of Q with respect to β givesµ∂Q

∂β

¶n,V

=

̶

∂β

"Xi

gie−β i

#!n,V

=Xi

gi

µ∂e−β i

∂β

¶n,V

(14.4)

= −Xi

gi ie−β i

By comparing this to the expression for U, we see

U = − 1Q

µ∂Q

∂β

¶n,V

= −µ∂ lnQ

∂β

¶n,V

, (14.5)

where we used the identity 1y∂y∂x= ∂ ln y

∂x.

Pressure

Another important property is pressure.

When the ensemble is in the particular state i, d i = −pidV . So at constanttemperature and number of particles

pi = −µ∂ i

∂V

¶n,β

(14.6)

Thus the ensemble average pressure is given by

P = p = − 1Q

Xi

gi

µ∂ i

∂V

¶n,β

e−β i . (14.7)

Multiplying by β/β we get

P = − 1

βQ

Xi

gi

µ∂ i

∂V

¶n,β

βe−β i . (14.8)

Using the chain rule in reverse, i.e.,

∂e−β i

∂V=

−βe−β iz | µ∂e−β i

∂ i

¶µ∂ i

∂V

¶= −

µ∂ i

∂V

¶βe−β i (14.9)

104

Page 117: 457454 Physical Chemistry Quantum Chemistry

we proceed as

P =1

βQ

Xi

gi

µ∂e−β i

∂V

¶n,β

=1

βQ

̶

∂V

Xi

gie−β i

!n,β

(14.10)

=1

βQ

µ∂Q

∂V

¶n,β

=1

β

µ∂ lnQ

∂V

¶n,β

.

Entropy

We have already obtained the expression for entropy. It is

S =U

T+ k lnQ (14.11)

= −kβµ∂ lnQ

∂β

¶n,V

+ k lnQ

105

Page 118: 457454 Physical Chemistry Quantum Chemistry

Helmholtz Free Energy

Free energy is the energy contained in the system which is available to do work.

That is, it is the energy of the system minus the energy that is “tied-up” in the

random (unusable) thermal motion of the particle in the system: A ≡ U − TS

Free energy is probably the key concept in thermodynamics and so we will discuss

it in much greater detail later. We will make the distinction between the Helmholtz

free energy and the more familiar Gibb’s free energy (G) later as well.

The Helmholtz free energy has the most direct relation to the partition function

as can be seen from

A ≡ U − TS = −µ∂ lnQ

∂β

¶n,V

+ kTβ

µ∂ lnQ

∂β

¶n,V

− kT lnQ (14.12)

= −kT lnQ

Any thermodynamic property can now be obtained from the above functions as

we shall see in the following chapters.

106

Page 119: 457454 Physical Chemistry Quantum Chemistry

15. Work

We now begin the study of thermodynamics.

Thermodynamics is a theory describing the most general properties of macroscopic

systems at equilibrium and the process of transferring between equilibrium states.

Thermodynamics is completely independent of the microscopic structure of the

system.

15.1. Properties of Partial Derivatives

Of critical importance in mastering thermodynamics is to become proficient with

partial derivatives.

∗ ∗ ∗ See Handout ∗ ∗∗

15.1.1. Summary of Relations

1. The total derivative of z(x, y):

dz =

µ∂z

∂x

¶y

dx+

µ∂z

∂y

¶x

dy (15.1)

2. The chain rule for partial derivatives:µ∂z

∂x

¶y

=

µ∂z

∂u

¶y

µ∂u

∂x

¶y

(15.2)

107

107

Page 120: 457454 Physical Chemistry Quantum Chemistry

3. The reciprocal rule: µ∂z

∂x

¶y

µ∂x

∂z

¶y

= 1 (15.3)

4. The cyclic rule: µ∂z

∂x

¶y

= −µ∂z

∂y

¶x

µ∂y

∂x

¶z

(15.4)

5. Finally µ∂z

∂x

¶u

=

µ∂z

∂x

¶y

+

µ∂z

∂y

¶x

µ∂y

∂x

¶u

(15.5)

15.2. Definitions

System: a collection of particles

Macroscopic systems: Systems containing a large number of particles.

Microscopic systems: Systems containing a small number of particles.

Environment : Everything not included in the system (or set of systems)

Note that the distinction between the system and the environment is arbitrary

and is chosen as a matter of convenience.

15.2.1. Types of Systems

Isolated system: A system that cannot exchange matter or energy with its envi-

ronment.

Closed system: A system that cannot exchange matter with its environment but

may exchange energy.

108

Page 121: 457454 Physical Chemistry Quantum Chemistry

Open system: A system that may exchange matter and energy with its environ-

ment.

Adiabatic system: A closed system that also can not exchange heat energy with

its environment.

15.2.2. System Parameters

Extensive parameters (or properties): properties that depend on the amount ofmatter.

• For example, volume, mass, heat capacity.

Intensive parameters (or properties): properties that are independent of theamount of matter.

• For example, temperature, pressure, density.

Extensive properties can be “converted” to intensive properties through ratios:

Extensive propertyExtensive property

→ Intensive property. (15.6)

For example massvolume = density,

volumemoles = molar volume,

heat capacitymass = specific heat.

15.3. Work and Heat

A system may exchange energy with its environment or another system in the

form of work or heat.

• Work is exchanged if external parameters are changed during the process.

• Heat is exchanged if only internal parameters are changed during the process.

109

Page 122: 457454 Physical Chemistry Quantum Chemistry

Convention

Work, w, is positive (w > 0) if work is done on the system.

Work is negative (w < 0) if work is done by the system.

Heat, q, is positive (q > 0) if heat is absorbed by the system.

Heat is negative (q < 0) if heat is released from the system.

15.3.1. Generalized Forces and Displacements

In physics you learned that an infinitesimal change in work is given by the product

of force, F , times and infinitesimal change in position, dx:

dw = Fdx. (15.7)

For thermodynamics, we need a more general definition if infinitesimal work.

Any given external parameter, A may be considered as a ‘generalized force’ which

is coupled to a particular internal parameter, a, which acts as ‘generalized dis-

placement.’

Note that the generalized force need not have units of force (e.g., Newtons) and

the generalized displacement need not have units of position (e.g., meters), but

the product of the two must have units of energy (e.g., Joules).

The infinitesimal amount of work done on the system is then given by

dw = Ada, (15.8)

or more generally as

dw =Xi

Aidai (15.9)

110

Page 123: 457454 Physical Chemistry Quantum Chemistry

if more than one set of parameters change.

The following table gives some examples of generalized forces and displacementsGeneralized Force, A Generalized Displacement, a Contribution to dw

Pressure, −P Volume, dV −PdVStress, σ Strain, dε σdε

Surface tension, γ Surface area, dA γdAVoltage, E Charge, dQ EdQ

Magnetic Field, H Magnetization, dM HdMChemical Potential, μ Moles, dn μdn

Gravity, mg Height, dh mgdh

15.3.2. PV work

In principle all work is interchangeable so that without loss of generality we will

develop the formal aspects of thermodynamics assuming all work is due to changes

in volume under a given pressure. That is

dw = −PdV, (15.10)

this is called PV work.

When we get to applications of thermodynamics we will then be concerned with

the various forms of work like those shown in the table above.

Expanding Gases

Consider the work done by a gas expanding in piston from volume V1 to V2 against

some constant external pressure P = Pex (see figure)

111

Page 124: 457454 Physical Chemistry Quantum Chemistry

The force exerted on a gas by a piston is equal to the external pressure times the

area of the piston: F = PexA⇒ Pex = F/A.

Recall from physics that work is the (path) integral over force: w = −R x2x1

Fdx.

This can be manipulated as

w = −Z x2

x1

Fdx = −Z x2

x1

F

A|zPex

Adx|zdV

= −Z V2

V1

PexdV (15.11)

If Pex is independent of V then

w = −Z V2

V1

PexdV = −PexZ V2

V1

dV = −Pex4V (15.12)

112

Page 125: 457454 Physical Chemistry Quantum Chemistry

16. Maximum Work and Reversible

changes

Now that we have learned about PV work we will consider the situation where

the system does the maximum amount of work possible.

16.1. MaximalWork: Reversible versus Irreversible changes

The value of w depends on Pex during the entire expansion.

In the figure

wA = −Z V2

V1

PatmdV = −Patm (V2 − V1) (16.1)

113

113

Page 126: 457454 Physical Chemistry Quantum Chemistry

and

wB = w1 + w2, (16.2)

where

w1 = −Z Vi

V1

Patm+2WdV = −Patm+2W (Vi − V1) (16.3)

and

w2 = −Z V2

Vi

PatmdV = −Patm (V2 − Vi) (16.4)

Hence it is clear that |wB| > |wA| .

Now consider case in the figure below

The expansion is reversible. That is, there is always an intermediate equilibriumthroughout the expansion. Namely Pgas = Pex. So,

wrev = −Z V2

V1

PgasdV (16.5)

This is the limiting case of path B in the previous figure. Thus wrev is the maxi-

mum possible work that can be done in an expansion. wrev = wmax.

114

Page 127: 457454 Physical Chemistry Quantum Chemistry

16.2. Heat Capacity

Temperature and heat are different.

Temperature is not the amount of heat.

Temperature is an intensive property and heat is an extensive property.

However, heat is related to temperature through the heat capacity

C(T ) =dq

dT(16.6)

n.b., heat capacity is a function of T ; it is not a constant.

From this equation

dq = C(t)dT, (16.7)

That is, when the temperature of a substance having a heat capacity C(t) is

changed by dT, dq amount of heat energy is transferred.

The heat capacity also depends on the conditions during the temperature change,

e.g., CV (T ) =¡dqdT

¢Vand CP (T ) =

¡dqdT

¢Pare not the same

Heat capacity is an extensive property. To make an intensive property

1. divide by the number of moles to get molar heat capacity

CVm(T ) =1

n

µdq

dT

¶V

(16.8)

2. divide by mass to get specific heat

cV =1

m

µdq

dT

¶V

(16.9)

We will discuss heat capacity more later.

115

Page 128: 457454 Physical Chemistry Quantum Chemistry

16.3. Equations of State

The macroscopic properties of matter are related to one another via a phenom-

enological equation of state.

The state of a pure, homogeneous material (in the absence of external fields) is

given by the values of any two intensive properties.

(More complicated systems require more than two independent variables, but

behave in the same way as the more simple pure system, so we will focus our

development of thermodynamics on simple systems.)

The functional dependence of any property on the two independent variables is

an equation of state. e.g., T , P independent then heat capacity is a function of T

and P , C(T, P ).

16.3.1. Example 1: The Ideal Gas Law

The equation of state for volume of an ideal gas is

PV = nRT, (16.10)

where R is the gas constant (8.315 J K−1 mol−1) and n is the number of moles.

The ideal gas equation of state can be expressed in terms of intensive variables

only

PVm = RT, (16.11)

where Vm = Vn.

The equation of state can also be expressed in terms of density ρ = mV(and molar

mass m/n)

ρ =mP

nRT=

MP

RT. (16.12)

116

Page 129: 457454 Physical Chemistry Quantum Chemistry

16.3.2. Example 2: The van der Waals Equation of State

A more realistic equation of state was presented by van der Waals:

P =nRT

V − nb− n2a

V 2. (16.13)

The parameter a attempts to account for the attractive forces among the particles

The parameter b attempts to account for the repulsive forces among the particles

b originates from hard sphere collisions (see figure):

117

Page 130: 457454 Physical Chemistry Quantum Chemistry

In term of intensive variables

P =RT

Vm − b− a

V 2m

. (16.14)

16.3.3. Other Equations of State

The van der Waals equation of state is not the only one that has been proposed.

Some other equations of state are

• BerthelotP =

nRT

V − nb− n2a

TV 2=

RT

Vm − b− a

TV 2m

(16.15)

• DietericiP =

nRTe−anRTV

V − nb=

RTe−a

RTVm

Vm − b(16.16)

• Redlich-Kwang

P =nRT

V − nb− n2a√

TV (V − nb)=

RT

Vm − b− a√

TVm (Vm − b)(16.17)

118

Page 131: 457454 Physical Chemistry Quantum Chemistry

17. The Zeroth and First Laws of

Thermodynamics

Over the course of the next two lectures we will discuss the four core laws of

thermodyanmics.

Today we will cover the zeroth and first laws, which deal with temperature and

total energy respectively.

Next time we will cover the second and third laws which both deal with entropy.

17.1. Temperature and the Zeroth Law of Thermodynamics

Temperature tells us the direction of thermal energy (heat) flow.

• Heat flows from high T to Low T.

Temperature scales

• Celsius: A relative scale based on water (T = 0C for melting ice and

T = 100C for boiling water)

• Kelvin: An absolute temperature scale based on the ideal gas law. Thetemperature at which (for fixed V and n) the pressure is zero is defined as

T = 0 K

• T (Kelvin) = T (Celsius) + 273.15

119

119

Page 132: 457454 Physical Chemistry Quantum Chemistry

Standard conditions

• standard temperature and pressure (STP): T = 273.15 K and P = 1 atm.

(Vm(STP) = 22.414 L/mol)

• standard ambient temperature and pressure (SATP): T = 298.15 K and

P = 1 bar. (Vm(SATP) = 24.789 L/mol)

Diathermic wall : A wall that allows heat to flow through it.

Adiabatic wall : A wall the does not allow heat to flow through it.

Thermal equilibrium: If two systems are in contact along a diathermic wall and

no heat flows across the wall, then the systems are in thermal equilibrium.

The zeroth law of thermodynamics

• Mathematical statement:

If TA = TB and TB = TC , then TA = TC (17.1)

This the mathematical statement of transitivity

• Verbal statement: If system A is in thermal equilibrium with system B

and system B is in thermal equilibrium system C then system A is also in

thermal equilibrium with system C.

The zeroth law implies that if an arbitrary system, C, is chosen as a thermometer

then it will read the same temperature when it is in thermal contact along a

diathermic wall with system A as when it is in thermal contact along a diathermic

wall with system B.

120

Page 133: 457454 Physical Chemistry Quantum Chemistry

17.2. The First Law of Thermodynamics

Definitions:

• State: the state of a system is defined by specifying a minimum number inintensive variables

• State Function: A function of the chosen independent variables that de-scribes a property of the state (e.g., V (T, P )). The value of the state func-

tion depends only on that given state and on no other possible state of the

system.

17.2.1. The internal energy state function

For characterizing the change in energy of a system, one is concerned with the

work done on the system (w) and the heat supplied to the system (q). The energy

of a system is called the internal energy (U) of the system.

The first law of thermodynamics:

• Mathematical statement:4U = q + w (17.2)

or in differential form

dU = dq + dw (17.3)

• Verbal statement: The change in internal energy of a system is equal to theamount of work done on the system plus the amount of heat provided to the

system.

So for a system where all the work is PV work the first law becomes

4U = q −Z V2

V1

PexdV (17.4)

121

Page 134: 457454 Physical Chemistry Quantum Chemistry

in differential form this is

dU = dq − PexdV (17.5)

Although U can be expressed as a function of any two state variables, the most

convenient at this time are V and T. U → U(T, V ).

The total differential of U(T, V ) is

dU =

µ∂U

∂T

¶V

dT +

µ∂U

∂V

¶T

dV (17.6)

Consider adding heat at a constant volume then

dU =

µ∂U

∂T

¶V

dT +

µ∂U

∂V

¶T

dV = dq − PexdV. (17.7)

So, µ∂U

∂T

¶V

dT = dq =⇒µ∂U

∂T

¶V

=dq

dT= CV (17.8)

Hence the slope¡∂U∂T

¢Vis the heat capacity.

The other slope,¡∂U∂V

¢T, is called the internal pressure (it has no standard symbol).

A useful relation (derivation to come) isµ∂U

∂V

¶T

= T

µ∂P

∂T

¶V

− P (17.9)

Example: A van der Waals gas

P =nRT

V − nb− n2a

V 2⇒µ∂P

∂T

¶V

=nR

V − nb(17.10)

122

Page 135: 457454 Physical Chemistry Quantum Chemistry

so the useful relation becomesµ∂U

∂V

¶T

= TnR

V − nb− P =

nRT

V − nb− nRT

V − nb+

n2a

V 2

= +n2a

V 2(17.11)

The equation of state for U : Express U in terms of T, V, and P.

Start with the total differential of U

dU =

µ∂U

∂T

¶V

dT +

µ∂U

∂V

¶T

dV (17.12)

but¡∂U∂T

¢V= CV and

¡∂U∂V

¢T= T

¡∂P∂T

¢V− P (useful relation). Hence

dU = CV dT +

∙T

µ∂P

∂T

¶V

− P

¸dV (17.13)

is the equation of state for U.

A useful approximation is 4U = CV4T which is valid for

i) heat capacity nearly constant over 4T and with no phase transitions.

ii) ideal gas or at constant volume.

123

Page 136: 457454 Physical Chemistry Quantum Chemistry

18. The Second and Third Laws of

Thermodynamics

18.1. Entropy and the Second Law of Thermodynamics

We learned from statistical mechanics that entropy, S, is a measure of the disorder

of the system and is expressed via Boltzmann’s equation S = k lnW (where W is

the micocanonical partition function)

We expressed Boltzmann’s law in terms of the more convenient canonical partition

function as

S =E

T+ k lnQ. (18.1)

Now, the average energy of the system E is in fact what we call internal energy:

U ≡ E.

Furthermore we derived the simple relation between the Helmholtz free energy

and the canonical partition function as A = −kT lnQ.

Hence,

S =U

T− A

T=1

T(U −A). (18.2)

Since U, A, and T are state functions, S is also a state function .

So we may write

dS =1

T(dU − dA) (18.3)

124

124

Page 137: 457454 Physical Chemistry Quantum Chemistry

for an isothermal process.

Recall the definition of Helmholtz free energy–the energy of the system available

to do work.

We learned previously that the maximum amount of work one can extract from

the system is the work done during a reversible process. Hence dA = dwrev.

For now let us limit the discussion to reversible processes. Then

dS =1

T(dU − dwrev) =

1

T(dqrev + dw/ rev − dw/ rev ) (18.4)

=dqrevT

. (Reversible process)

Note: An alternative approach to thermodynamics which makes no reference to

molecules or statistical mechanics is to simply begin by defining entropy as dS ≡dqrevT

The principle of Clausius

• “The entropy of an isolated system will always increase in a spontaneous

process”

• Mathematical statement: (dS)U,V ≥ 0

For a general process: dU = dq − PexdV

For a reversible process Pex = P and dq = TdS so dU = TdS − PdV

125

Page 138: 457454 Physical Chemistry Quantum Chemistry

Since U, S, T, P, and V are state functions, dU = TdS − PdV holds for any

process, but in general, TdS is not heat and −PdV is not work. (see figure)

TdS is heat and −PdV is work only for reversible processes.

For some dU,

dq − PexdV = TdS − PdV ⇒ TdS = dq − PexdV + PdV (18.5)

TdS = dq + (P − Pex) dV

• Case i) Pex > P then (spontaneous) dV is negative so (P−Pex)dV is positive.

• Case ii) P > Pex then (spontaneous) dV is positive so (P−Pex)dV is positive.

• Case iii) P = Pex then (spontaneous) dV is zero so (P − Pex)dV is zero.

Thus for any spontaneous process TdS ≥ dq.

This is a mathematical statement of the second law of thermodynamics

126

Page 139: 457454 Physical Chemistry Quantum Chemistry

18.1.1. Statements of the Second Law

Unlike the first law, the second law has a number of equivalent statements

1. A cyclic process must transfer heat from a hot to cold reservoir if it is to

convert heat into work.

2. Work must be done to transfer heat from a cold to a hot reservoir.

3. A useful perpetual motion machine does not exist.

4. The entropy of the universe is increasing

5. Spontaneous processes are irreversible in character.

6. The entropy of an isolated system will always increase in a spontaneous

process (the principle of Clausius)

18.2. The Third Law of Thermodynamics

Consider the first law for a reversible change at constant volume.

dU = dq + dw = dq − PexdV (18.6)

From our earlier discussion of heat capacity dq = CV dT (CV since constant vol-

ume). So,

dU = CV dT (18.7)

but also dU = TdS. So

dS =CV dT

T=⇒ 4S =

Z T2

T1

CV

TdT. (18.8)

127

Page 140: 457454 Physical Chemistry Quantum Chemistry

A very similar derivation can be done for a reversible change at constant pressure

(we can not do it quite yet) to yield

4S =

Z T2

T1

CP

TdT (18.9)

18.2.1. The Third Law

Verbal statement

The third law of thermodynamics permits the absolute measurement of entropy.

To derive the mathematical statement of the third laws we starting with

4S =

Z T2

T1

CP

TdT (18.10)

now let T1 → 0

4S = S2 − S0 =

Z T2

0

CP

TdT (18.11)

Hence the mathematical statement of the third law is

S(T2) =

Z T2

0

CP

TdT + S0 (18.12)

From a macroscopic point of view S0 is arbitrary. However, a microscopic point of

view suggests S0 = 0 for perfect crystals of atoms or of totally symmetric molecules

(e.g., Ar, O2 etc.). S0 6= 0 for imperfect crystals and crystals of asymmetric

molecules (e.g., CO).

Alternative statement of the third law: Absolute zero is unattainable.

Consider the heat capacity near T → 0.

For S0 to have significance CPTmust be finite (not infinite) as T → 0. Thus CP → 0.

128

Page 141: 457454 Physical Chemistry Quantum Chemistry

But CP =dqdT→ 0 implies dT

dq→∞.

In other words, an infinitesimal amount of heat causes an infinite change in tem-

perature.

In view of what we have learned about fluctuations, the ever present random

fluctuations in energy provide the infinitesimal amount of heat and so you can

never reach absolute zero corresponding to an average energy of zero.

18.2.2. Debye’s Law

Heat capacity data only goes down so far. So one needs a theoretical extrapolation

down to T = 0. (Debye)

Postulate: CPm = aT 3. That is at low temperatures heat capacity goes as the

cube of the temperature.

C∗Pm, T∗ are the lowest temperature data points. So, a = C∗Pm/T

∗3.

129

Page 142: 457454 Physical Chemistry Quantum Chemistry

The molar entropy is

Sm(T∗) =

Z T∗

0

CP

TdT

CPm=aT3

=C∗PmT ∗3

Z T∗

0

T 2dT (18.13)

=C∗PmT ∗3

T 3

3

¯T∗0

=C∗Pm3

.

18.3. Times Arrow

Entropy and the second law give a direction to time.

For example, if we see a picture of your PChem book in mint condition and we see

a picture of your PChem book all battered and beaten. We know which picture

was taken first.

The interesting thing is that each molecule in a macroscopic system obeys time

invariant dynamics. Both Newton’s laws and Quantum dynamics (next semester)

are the same if you replace t with −t.

Yet, the behavior of the macrosystem definitely changes if you replace t with −t.

Thus the simple fact that you have an enormous number of particles induces a

perceived asymmetry in time.

130

Page 143: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 3

Listed here are some of the key equations for Exam 3. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• The Boltzmann equation isS = k lnW. (18.14)

• The Boltzmann distribution :

gie−β iP

j gje−β j

(18.15)

• The canonical partition function is

Q =Xj

gje−βEj (18.16)

131

131

Page 144: 457454 Physical Chemistry Quantum Chemistry

• The relation between the partition function and the molecular partitionfunction is

Q =qNmolN !

. (18.17)

• The Translational Partition Function

qtrans =V

Λ3(18.18)

where

Λ ≡ h√2πmkT

(18.19)

is the thermal de Broglie wavelength.

• The Rotational Partition Function (linear molecules) is

qrot ≈T

σθr, (18.20)

where θr ≡ h2

8π2Ik(I is the moment of inertia) and σ is the so-called sym-

metry number in which σ = 1 for unsymmetrical molecules and σ = 2 for

symmetrical molecules

• The Vibrational Partition Function is

qvib =1

2 sinh 12β~ω

. (18.21)

• The ensemble average of any property is given by

O =1

Q

Xi

Oigie−β i . (18.22)

• The relations between the canonical partition function and the thermody-namics variables are

Helmholtz Free Energy A = −kT lnQInternal energy U = − 1

Q

³∂Q∂β

´n,V

= −³∂ lnQ∂β

´n,V

Entropy S = −kβ³∂ lnQ∂β

´n,V+ k lnQ

Pressure P = 1βQ

¡∂Q∂V

¢n,β= 1

β

¡∂ lnQ∂V

¢n,β

132

Page 145: 457454 Physical Chemistry Quantum Chemistry

• PV work is

dw = −PdV. (18.23)

• Heat capacity:dq = C(t)dT. (18.24)

• General forms of the first law:

4U = q + w, (18.25)

in differential form this is

dU = dq − PexdV. (18.26)

Also,

dU = TdS − PdV. (18.27)

• The second lawTdS ≥ dq. (18.28)

• The third lawS(T2) =

Z T2

0

CP

TdT + S0 (18.29)

• Debye’s law for entropy at very low temperatures

Sm(T∗) =

C∗Pm3

, (18.30)

where C∗Pm is the molar heat capacity at the lowest temperature for which

there is data.

133

Page 146: 457454 Physical Chemistry Quantum Chemistry

Part IV

Basics of Thermodynamics

134

134

Page 147: 457454 Physical Chemistry Quantum Chemistry

19. Auxillary Functions and Maxwell

Relations

We have stated that thermodynamics as we are studying it deals with states in

equilibrium or transitions between equilibrium states.

Consequently, the concept of equilibrium plays a key role in much of what we will

discuss for the remainder of the year.

The equilibrium constant for a thermodynamic process,K, (which you are familiar

with from general chemistry) serves are a common point which connects thermo-

dynamics, electrochemistry, and kinetics–topics we will encounter throughout

the year.

19.1. The Other Important State Functions of Thermody-

namics

As was the case in quantum mechanics, here too is energy the key property with

which to work.

So far we have encountered two state functions which characterize the energy of a

macroscopic system–the internal energy and, briefly the Helmholtz free energy.

135

135

Page 148: 457454 Physical Chemistry Quantum Chemistry

From the first law as stated as

dU = TdS − PdV (19.1)

we say that the natural (most convenient) variables for the equation of state for

U are S and V . This is U = U(S, V )

Unfortunately S can not be directly measured and most often P is a more conve-

nient variable than V

Because of this fact, it is handy to define state functions which have different pairs

of natural variables, so that no mater what situation arises we have convenient

equations of state to work with.

The other pairs of natural variables being (S and P ), (T and V ) and (T and P )

The table below lists these state functions

State function Symbol Natural variables Definition Units

Internal Energy U S and V energy

Enthalpy H S and P H ≡ U + PV energy

Helmholtz free energy A T and V A ≡ U − TS energy

Gibbs free energy G T and P G ≡ H − TS energy

We consider each of these functions in turn

19.2. Enthalpy

We want a state function whose natural variables are S and P

Let us try the definition H ≡ U + PV.

136

Page 149: 457454 Physical Chemistry Quantum Chemistry

Now formally

dH = dU + d(PV ) = dU + PdV + V dP, (19.2)

but dU = TdS − PdV, so

dH = TdS − PdV/ + PdV/ + V dP (19.3)

= TdS + V dP.

Hence Enthalpy does indeed have the desired natural variables.

19.2.1. Heuristic definition:

Enthalpy is the total energy of the system minus the pressure volume energy. So

a change in enthalpy is the change in internal energy adjusted for the PV work

done. If the process occurs at constant pressure then the enthalpy change is the

heat given off or taken in.

For example, consider an reversibly expanding gas under constant pressure (dP =

0) and adiabatic (dq = 0) conditions.

The system does work during the expansion; in doing so it must lose energy. Since

the process is adiabatic no heat energy can flow in to compensate for the work

done and the gas cools.

The total internal energy decreases. The enthalpy of the system on the other hand

does not change–it is the internal energy adjusted by an amount of energy equal

to the PV work done by the system. As Freshmen we learn this as 4H = qp.

19.3. Helmholtz Free Energy

Now we want a state function whose natural variables are T and V

137

Page 150: 457454 Physical Chemistry Quantum Chemistry

Let us try the definition A ≡ U − TS.

Formally

dA = dU − d(TS) = dU − TdS − SdT, (19.4)

but dU = TdS − PdV, so

dA = TdS/ − PdV − TdS/ − SdT (19.5)

= −PdV − SdT.

Hence Helmholtz free energy does indeed have the desired natural variables.

19.3.1. Heuristic definition:

As we have said before Helmholtz free energy is the energy of the system which is

available to do work–It is the internal energy minus that energy which is “used

up” by the random thermal motion of the molecules.

19.4. Gibbs Free Energy

Finally we want a state function whose natural variables are T and P

Let us try the definition G ≡ H − TS.

Now formally

dG = dH + d(TS) = dH − TdS − SdT, (19.6)

but from above dH = TdS + V dP, so

dG = TdS/ + V dP − TdS/ − SdT (19.7)

= V dP − SdT.

Hence Gibbs free energy does indeed have the desired natural variables.

138

Page 151: 457454 Physical Chemistry Quantum Chemistry

19.4.1. Heuristic definition:

Gibbs free energy is the energy of the system which is available to do non PV

work–It is the internal minus both that energy which is “used up” by the random

thermal motion of the molecules and used up in doing the PV work.

19.5. Heat Capacity of Gases

19.5.1. The Relationship Between CP and CV

To find how CP and CV are related we begin with

dH = TdS + V dP (19.8)

at constant pressure and reversible conditions

dH = TdS (19.9)

dH = dq

but

dq = CPdT (19.10)

The constant pressure heat capcity can then be expressed in terms of enthalpy as

CP =

µ∂H

∂T

¶P

. (19.11)

So,

CP =

µ∂ (U + PV )

∂T

¶P

=

µ∂U

∂T

¶P

+ P

µ∂V

∂T

¶P

(19.12)

note¡∂U∂T

¢Pis not CV we need

¡∂U∂T

¢V. Use an identity of partial derivativesµ

∂U

∂T

¶P

=

µ∂U

∂T

¶V

+

µ∂U

∂V

¶T

µ∂V

∂T

¶P

(19.13)

139

Page 152: 457454 Physical Chemistry Quantum Chemistry

thus

CP =

µ∂U

∂T

¶V

+

µ∂U

∂V

¶T

µ∂V

∂T

¶P

+ P

µ∂V

∂T

¶P

(19.14)

= CV +

µ∂V

∂T

¶P

∙µ∂U

∂V

¶T

+ P

¸.

Recall the expression for internal pressure¡∂U∂V

¢T= T

¡∂P∂T

¢V− P . Then

CP = CV +

µ∂V

∂T

¶P

∙T

µ∂P

∂T

¶V

− P/ + P/¸

(19.15)

Finally

CP = CV + T

µ∂V

∂T

¶P

µ∂P

∂T

¶V

(19.16)

Example: Ideal gases

1. Ideal gas (equation of state: PV = nRT ): This equation is easily made

explicit in either P or V so we don’t need any of the above replacements

CP = CV + T

µ∂V

∂T

¶P

µ∂P

∂T

¶V

(19.17)

= CV + TnR

P

nR

V=

nRT

PVnR

Thus CP = CV + nR or

CPm = CVm +R (19.18)

19.6. The Maxwell Relations

Summary of thermodynamic relations we’ve seen so far

Definitions and relations:

• H = U + PV

140

Page 153: 457454 Physical Chemistry Quantum Chemistry

• A = U − TS

• G = H − TS

• CV =¡∂U∂T

¢V, CP =

¡∂H∂T

¢P

basic equations Maxwell relations working equations

dU = TdS − PdV¡∂T∂V

¢S= −

¡∂P∂S

¢V

dU = CV dT +£T¡∂P∂T

¢V− P

¤dV

dH = TdS + V dP¡∂T∂P

¢S=¡∂V∂S

¢P

dH = CPdT −£T¡∂V∂T

¢P− V

¤dP

dA = −PdV − SdT¡∂S∂V

¢T= +

¡∂P∂T

¢V

dS = CVTdT +

¡∂P∂T

¢VdV

dG = V dP − SdT¡∂S∂P

¢T= −

¡∂V∂T

¢P

dS = CPTdT −

¡∂V∂T

¢PdP

We will get plenty of practice with derivations based on these equations and on

the properties of partial derivatives. (See handout and Homework)

141

Page 154: 457454 Physical Chemistry Quantum Chemistry

20. Chemical Potential

20.1. Spontaneity of processes

Two factors drive spontaneous processes

1. The tendency to minimize energy

2. The tendency to maximize entropy

Let us begin with Helmholtz free energy

The total differential of A is (A = U − TS)

dA = dU − TdS − SdT = dq − PexdV − TdS − SdT (20.1)

For constant T and V, (dA)T,V = dq − TdS

From the second law, TdS ≥ dq for a spontaneous process, (dA)T,V ≤ 0 for aspontaneous process.

Hence at equilibrium (dA)T,V = 0.

For chemistry it is most often more convenient to use Gibbs free energy

The total differential of G is

dG = dH − TdS − SdT = dq − PexdV + PdV + V dP − TdS − SdT (20.2)

142

142

Page 155: 457454 Physical Chemistry Quantum Chemistry

For constant T and P = Pex, (dG)T,P = dq − TdS

Again from the second law, TdS ≥ dq for a spontaneous process, (dG)T,P ≤ 0 fora spontaneous process.

Hence at equilibrium (dG)T,P = 0.

So free energy provides a measure of the thermodynamic driving force towards

equilibrium.

Note free energy provides no information about how fast a process proceeds toequilibrium.

The free energy functions are the workhorses of applied thermodynamics so we

want to get a feel for them.

Returning to the total differentials of free energy,

dA = dU − TdS − SdT (20.3)

and

dG = dH − TdS − SdT. (20.4)

Expressing dU and dH generally as dU = TdS − PdV and dH = TdS + V dP

(remember that in general TdS cannot be identified with dq and PdV cannot be

identified with −w).

Plugging these into the total differentials of free energy gives

dA = −SdT − PdV (20.5)

and

dG = −SdT + V dP (20.6)

143

Page 156: 457454 Physical Chemistry Quantum Chemistry

These expressions are quite general, but i) only PV work and ii) closed systems.

The total differential of A is also

dA = dq + dw − TdS − SdT. (20.7)

For a reversible process dq = TdS and work is maximal.

Hence (dA)T = dwmax =⇒ (4A)T = wmax. As we have stated in words a number

of times before.

The total differential of G is also

dG = dq + dw + PdV + V dP − TdS − SdT. (20.8)

In general dw = dw0 − PexdV where dw0 is the non-PV work.

The total differential of G becomes

dG = dq + dw0 − PexdV + PdV + V dP − TdS − SdT. (20.9)

For constant T and P = Pex, (dG)T,P = dq + w0 − TdS.

For reversible processes q = TdS and this becomes

(dG)T,P = dw0max =⇒ (4G)T,P = w0max (20.10)

So, as stated earlier, the Gibbs free energy is the energy of the system available

to do non-PV work.

20.2. Chemical potential

What if the amount of substance can change?

144

Page 157: 457454 Physical Chemistry Quantum Chemistry

Extensive properties depend on the amount of “stuff”

For example A(T, V ) now becomes A(T, V, n) and the total differential becomes

dA(T, V, n) =

µ∂A

∂T

¶V,n

dT +

µ∂A

∂V

¶T,n

dV +

µ∂A

∂n

¶V,T

dn (20.11)

Let’s focus on the slope¡∂A∂n

¢V,T.

• This is a measure of the change in Helmholtz free energy of a system (at

constant T and V ) with the change in the amount of material.

• Physically, this is a measure of the potential to change the amount of mate-rial.

• It defines the chemical potential μ ≡¡∂A∂n

¢V,T

.

So we can also write

dA = −SdT − PdV + μdn (20.12)

What about the relation of the chemical potential to Gibbs’ free energy?

G = H − TS = U − TS| z =A

+ PV = A+ PV so,

dG = dA+ PdV + V dP (20.13)

= −SdT − P/ dV/ + P/ dV/ + V dP + μdn

= −SdT + V dP + μdn,

but from

dG =

µ∂G

∂T

¶P,n

dT +

µ∂G

∂P

¶T,n

dP +

µ∂G

∂n

¶P,T

dn (20.14)

145

Page 158: 457454 Physical Chemistry Quantum Chemistry

we see that

μ =

µ∂G

∂n

¶P,T

. (20.15)

So, μ is also a measure of the change in Gibbs free energy of a system (at constant

T and P ) with the change in the amount of material and it still has the same

physical meaning.

The Gibbs free energy per mole (Gm) for a pure substance is equal to the chemical

potential. (Gm = μ)

20.3. Activity and the Activity coefficient

When, for example, a solute is dissolved in a solvent, there exist complicated

interactions which cause deviations from ideal behavior.

To account for this one must introduce the concept of activity and the activity

coefficient.

Activity is hard to define in words and indeed it has an awkward mathematical

definition as we will soon see.

The activity coefficient has a more convenient definition which is that it is the

measure of how a particular real system deviates from some reference system

which is usually taken to be ideal.

The mathematical definition of activity ai of some species i is implicitly stated as

limζ→ζª

aig(ζ)

= 1 (20.16)

where g(ζ) is any reference function (e.g., pressure, mole fraction, concentration

etc.), and ζª is the value of ζ at the reference state.

146

Page 159: 457454 Physical Chemistry Quantum Chemistry

This implicit definition is awkward so for convenience one defines the activity

coefficient as the argument of the above limit,

γi ≡aig(ζ)

(20.17)

which we can rearrange as

ai = γig(ζ). (20.18)

The definition of activity implies that γi = 1 at g(ζª) (the reference state)

20.3.1. Reference States

Thermodynamics is founded on the concept of energy which we know to have an

arbitrary scale. That is, we can define are zero of energy any where we want.

Because of this it is always necessary to specify a reference state to which our real

state can be compared.

The choice of this state is completely up to us, but it is often the case that the

reference state is chosen to be some ideal state.

For example, if we are talking about a gas we will mostly likely choose the ideal

gas law in terms of pressure (P = nRT/V ) as our reference function and the

reference state being when P = 0 since we know all gases behave ideally in the

limit of zero pressure.

Let us consider the activity of a real gas for the above reference function and

reference state. Note: the activity of gases as referenced to pressure has the

special name fugacity (fugacity is a special case of activity).

147

Page 160: 457454 Physical Chemistry Quantum Chemistry

Our reference function is very simple: g(ζ) = ζ = P , so

γ =a

P⇒ a = γP. (20.19)

Thus the activity of our real gas is given by the activity coefficient times the

pressure of an ideal gas under the same conditions.

Based on the condition that γ → 1 as we approach the reference state (P = 0

in this case) we see that the activity (or fugacity) of a real gas becomes equal to

pressure for low pressures

20.3.2. Activity and the Chemical Potential

One cannot measure absolute chemical potentials, only relative potentials can be

measured. By convention we chose a standard state and measure relative to that

state.

The deviation of the chemical potential at the state of interest versus at the

reference state is determined by the activity at the current state (the activity at

the reference state is unity by definition).

μi − μªi = RT ln ai. (20.20)

Rather than referencing to the standard state one can also reference to any con-

venient “ideal” state. This ideal state is in turn referenced to the standard state.

For the state of interest

μi = μªi +RT ln ai (20.21)

and for the ideal state

μidi = μªi +RT ln aidi ⇒ μªi = μidi −RT ln aidi . (20.22)

148

Page 161: 457454 Physical Chemistry Quantum Chemistry

Thus,

μi = μidi −RT ln aidi +RT ln ai (20.23)

μi − μidi = +RT ln ai −RT ln aidi

= RT lnaiaidi

Example: Real and ideal gases at constant temperature, but any pressure.Starting from the begining

dμid = dGm = −Sm=0z|dT + VmdP (20.24)

dμid = VmdP

dμid =RT

PdP.

Now we integrate from the reference state to the current state of interestZμª

dμid =

ZPª

RT

PdP. (20.25)

This gives

μid − μª = RT lnP

Pª. (20.26)

The usual standard state is the ideal gas at Pª = 1, so

μid = μª +RT lnP. (20.27)

(Note that as P → 1, μid → μª).

Lets say our gas is not ideal, then at a given pressure

μ = μª +RT ln a. (20.28)

For gases activity is usually called fugacity and given the symbol f , so a = f for

real gasses. Thus

μ = μª +RT ln f. (20.29)

149

Page 162: 457454 Physical Chemistry Quantum Chemistry

Lets say that instead of referencing to the ideal gas at P = 1, we want to reference

to the ideal gas at the current pressure P.

This is easily done by using μª = μid −RT lnP in the above equation for μ,

μ = μid −RT lnP +RT ln f

μ = μid +RT lnf

P.

Example: The barometric equation for an ideal gas.We have an ideal gas so,

μid = μª +RT lnP (20.30)

where we will take the reference state to be at sea level, i.e. Pª = 1 atm.

So at sea level

μid(0) = μª +

=0z | RT ln 1 = μª (20.31)

and at elevation h

μid(h) = μª +RT lnPh (20.32)

The gas fields the gravitational force which gives it a potential energy per mole

of Mgh at height h. We add this energy per mole term to the chemical potential

(which is free energy per mole) thus at equilibrium

μid(0) = μid(h) +Mgh (20.33)

Referencing to the reference state we get

μª/ = μª/+RT lnPh +Mgh (20.34)

RT lnPh = −Mgh

Ph = e−MghRT

The last line is the barometric equation and it shows that pressure is exponentially

decreasing function of altitude.

150

Page 163: 457454 Physical Chemistry Quantum Chemistry

21. Equilibrium

First let us consider the equilibrium A­ B.

Since A and B are in equilibrium their chemical potentials must be equal

μA = μB (21.1)

Now,

μA = μªA +RT ln aA (21.2)

and

μB = μªB +RT ln aB

So the equilibrium condition becomes

μªA +RT ln aA = μªB +RT ln aB (21.3)

−4μª = μªA − μªB = RT ln aB −RT ln aA

−4μª = RT lnaBaA

Since chemical potential is free energy per mole, if we multiply the above by n

moles we have

−4Gª = nRT lnaBaA

as a consequence of the equilibrium condition.

The quantityaBaAdefines the equilibrium constant, Ka, for this process.

151

151

Page 164: 457454 Physical Chemistry Quantum Chemistry

Say the system A→ B is not in equilibrium then we can not write

μA = μB

but we can write

μA +

4μz | μB − μA = μB (21.4)

Proceeding as above we get

μªA +RT ln aA +4μ = μªB +RT ln aB (21.5)

4μ = μªB − μªA +RT ln aB −+RT ln aA4μ = 4μª +RT ln

aBaA

.

Again multiplying by n gives

4G = 4Gª + nRT lnaBaA

.

If the 4G < 0 then the transition A→ B proceeds spontaneously as written.

Consider a more complicated equilibrium

aA+ bB ­ cC + dD. (21.6)

The equilibrium condition is

aμA + bμB = cμC + dμD. (21.7)

In a manner similar to the above

aμªA+aRT ln aA+ bμªB+ bRT ln aB = cμªC + cRT ln aC +dμªD+dRT ln aD (21.8)

Rearranging gives

≡−4rxnGªz | aμªA + bμªB − cμªC − dμªD = RT ln

acCadD

aaAabB

(21.9)

152

Page 165: 457454 Physical Chemistry Quantum Chemistry

the equilibrium constant is

Ka =acCa

dD

aaAabB

=⇒4Gª = −RT lnKa (21.10)

Note: n is absent in the above since the molar values are implied by the stoi-

chiometry.

21.0.3. Equilibrium constants in terms of KC

Equilibrium constant in terms of a condensed phase concentration:

K 0C =

[C]c [D]d

[A]a [B]b, (21.11)

which is related to Ka by

Ka = K 0C

µγcCγ

dD

γaAγbB

¶. (21.12)

If the reactants are solutes then as the solution is diluted all the activity coefficients

go to unity and K 0C → Ka.

21.0.4. The Partition Coefficient

Up to now we have only considered miscible solutions.

We now consider the problem of determining the equilibrium concentrations of a

solute A in both phases of an immiscible mixture.

153

Page 166: 457454 Physical Chemistry Quantum Chemistry

The equilibrium equation is

Aα ­ Aβ (21.13)

The equilibrium expression for this process is

4Gα→β = 0 = 4Gªα→β − nRT lnKa, (21.14)

where,4Gªα→β ≡ Gª

β −Gªα . The equilibrium constant for this process has a special

name; it is called the partition coefficient, P β/α ≡ Kβ/αpart, for species A in the α—β

mixture.

We can solve for the partition coefficient to yield

P β/α =aβAaαA= e−

4Gªα→β

nRT . (21.15)

For low concentrations

P β/α ' [A]β

[A]α. (21.16)

Knowledge of the partition function is important on the delivery of drugs because,

to enter the body, the drugs must transfer between an aqueous phase and a oil

phase.

154

Page 167: 457454 Physical Chemistry Quantum Chemistry

For most drugs

0 < Po/wpart < 4 (21.17)

Partition coefficient Delivery mechanism

low Po/wpart (likes water) injection

medium Po/wpart oral

high Po/wpart (likes oil) skin patch/ointment

Factors other than the partition coefficient influence the drug delivery choice. For

example, can the drug handle the acidic environment of the stomach?

155

Page 168: 457454 Physical Chemistry Quantum Chemistry

22. Chemical Reactions

Up to now we have only been considering systems in the absence of chemical

reactions. After chemical reactions take place the system is in a final “product”

thermodynamic state that is in general different from the initial “reactant” state.

For any extensive property

• 4rxn(Property) = property of products − property of reactants

• Example

— Reaction: aA+bB= cC+dD

— 4rxnS = cSm,C + dSm,D − aSm,A − bSm,B

22.1. Heats of Reactions

Exothermic reaction: heat is given off to the surroundingsEndothermic reaction: heat is given taken in from the surroundings

At constant pressure (Pex = P

q = 4rxnU − w = 4rxnU − P4rxnV = 4rxnH (22.1)

4rxnH < 0 for Exothermic reactions.

4rxnH > 0 for Endothermic reactions.

156

156

Page 169: 457454 Physical Chemistry Quantum Chemistry

22.1.1. Heats of Formation

Hess’s Law of heat summation: 4rxnH is independent of chemical pathway

Example: C2H2+H2 = C2H4.This direct reaction is not easy but it can be done in steps

C2H2 + 52O2 → 2CO2 +H2O(liq) 4rxnH

ª = −1299.63 kJ2CO2 +2H2O(liq)→C2H4 + 3O2 4rxnH

ª = +1410.97 kJ

H2 + 12O2 →H2O(liq) 4rxnH

ª = −285.83 kJC2H2+H2 = C2H4 4rxnH

ª = −174.49 kJ

The heat of formation 4fHª is the 4rxnH at STP in forming a compound from

its constituent atoms in their natural states.

O2, H2, C(graphite) are examples of atoms in their natural state.

Example: Formation of water

• H2 + 12O2 = H2O not 2H2+O2 = 2H2O

• 4rxnH =P

i νi4fH(i), where νi is the stoichiometric factor of the ith com-

ponent.

Example: H2O(liq)→H2O(gas) at SATPH2 + 1

2O2 = H2O(gas) 4fH

ª = −241.818 kJH2 + 1

2O2 = H2O(liq) 4fH

ª = −285.830 kJH2O(liq)→H2O(gas) 4rxnH

ª = −241.818− (−285.830) = 44.012 kJ

22.1.2. Temperature dependence of the heat of reaction

4rxnH(T2) = 4rxnH(T1) +

Z T2

T1

4rxnCPdT (22.2)

157

Page 170: 457454 Physical Chemistry Quantum Chemistry

22.2. Reversible reactions

Recall the requirement for a spontaneous change: 4G < 0 for constant T and P.

4rxnG = G(products)−G(reactants) =Xi

νiμi, (22.3)

(remember μi = Gm,i for pure substance i).

As we saw before μi can be defined in terms of activity

μi = μªi +RT ln ai. (22.4)

So,

4rxnG =

4rxnGªz | Xi

νiμªi +RT

Xi

νi ln ai. (22.5)

Using the property of logarithms: a lnx + b ln y = ln(xayb) the above expression

becomes

4rxnG = 4rxnGª +RT ln

Yi

aνii (22.6)

4rxnG = 4rxnGª +RT lnQ,

where Q ≡Q

i aνii is the activity quotient.

At equilibrium, 4rxnG = 0 and Q = Ka (Thermodynamic equilibrium constant).

Ka depends on T but is independent of P.

For the reaction aA + bB = cC + dD

Ka =acCa

dD

aaAabB

(22.7)

• Note that the activity of any pure solid or liquid is for all practical purposesequal to 1.

158

Page 171: 457454 Physical Chemistry Quantum Chemistry

• For ideal gases, ai = PiPª =

XiPPª (Pª = 1 bar) This leads to the sometimes

useful relation

KP =P cCP

dD

P aAP

bB

=(PªaC)

c (PªaD)d

(PªaA)a (PªaB)

b= Ka

¡Pª¢c+d−a−b , (22.8)

or more generally KP = Ka (Pª) i νi .

So at equilibrium, 4rxnG = 4rxnGª +RT lnQ becomes

0 = 4rxnGª +RT lnKa ⇒4rxnG

ª = −RT lnKa. (22.9)

22.3. Temperature Dependence of Ka

Starting with G = H − TS or G/T = H/T − S.

From this³∂(G/T )∂(1/T )

´P= H.

Applying this to4rxnG

ª

T=4rxnH

ª

T−4rxnS (22.10)

gives µ∂(4rxnG

ª/T )

∂(1/T )

¶P

= 4rxnHª (22.11)

Using 4rxnGª = −RT lnKa, we getµ

∂ lnKa

∂(1/T )

¶P

ind.=of P

d lnKa

d(1/T )= −4rxnH

ª

R(22.12)

or (using dd(1/T )

= dTd(1/T )

ddT= −T 2 d

dT)

d lnKa

dT=4rxnH

ª

RT 2(22.13)

Integration gives

lnKa(T2) = lnKa(T1) +1

R

Z T2

T1

4rxnHªm

T 2(22.14)

For a reasonably small range T2 − T1 this is well approximated by

lnKa(T2) = lnKa(T1)−4rxnH

ªm

R

µ1

T2− 1

T1

¶(22.15)

159

Page 172: 457454 Physical Chemistry Quantum Chemistry

22.4. Extent of Reaction

There are other equilibrium “constants” that are used in the literature.

• From Pj = XjP , KX = KPP−4υg

• From nj = PjVRT(ideal gas approximation), Kn = KP

¡RTV

¢−4υg

• From concentration Cj =njV=

PjRT

, KC = KP (RT )−4υg

Equilibrium “constants”“constants” expression relation to Ka situation used

Kaactivity(products)activity(reactants) – when an exact answer is needed

KPpartial pressure(products)partial pressure(reactants)

Ka

KγPª−4υg gas reactions

KXmole fraction(products)mole fraction(reactants)

µKa

KγPª−4υg

¶P−4υg when eq. P is known

Knmoles(products)moles(reactants)

µKa

KγPª−4υg

¶¡RTV

¢−4υg when V is known and constant

KCconcentration(products)concentration(reactants)

µKa

KγPª−4υg

¶(RT )−4υg when concentration known

160

Page 173: 457454 Physical Chemistry Quantum Chemistry

23. Ionics

Many chemical processes involve electrolytes and or acids and bases.

To understand these processes we must know something about how ions behave

in solution.

23.1. Ionic Activities

Consider a salt in solution

Mv+Xv− → v+Mz+(aq) + v−X

z−(aq), (23.1)

where v+ (v−) is the number of cations (anions) and z+ (z−) is the charge on the

cation (anion).

The chemical potential for the salt may be written in terms of the chemical po-

tential for each of the ions:

μsalt = v+μ+ + v−μ− (23.2)

To determine the activity we start with

ln aj =μj − μªjRT

, j = + or − (23.3)

and

ln asalt =μsalt − μªsalt

RT. (23.4)

161

161

Page 174: 457454 Physical Chemistry Quantum Chemistry

Substituting the expression for μsalt into this gives

ln asalt =v+μ+ − v−μ− + v+μ

ª+ − v−μ

ª−

RT(23.5)

=v+μ+ − v+μ

ª+

RT| z v+ ln a+

+v−μ− − v−μ

ª−

RT| z v− ln a−

So,

ln asalt = v+ ln a+ + v− ln a− (23.6)

or, alternatively,

asalt = av+av− (23.7)

It is the case that 1 mole of salt behaves like v = v++ v− moles of nonelectrolytes

in terms of the colligative properties. This suggests that the interesting quantity

is μsaltv:

μsaltv=

μªsaltv+RT ln a

1/vsalt. (23.8)

We see that

a1/vsalt = (a

v+av−)1/v ≡ a±. (23.9)

The quantity a± is the mean ionic activity.

23.1.1. Ionic activity coefficients

The activity coefficients for ionic solutions can also be defined via

a+ = γ+m+, a− = γ−m−, (23.10)

where m+ = v+m and m− = v−m.

The mean ionic activity coefficient is

γ± = (γv++ γ

v−− )

1/v. (23.11)

162

Page 175: 457454 Physical Chemistry Quantum Chemistry

The quantities a+, a−, γ+ and γ− cannot be measured individually.

One can use the colligative properties to measure the ionic activity coefficients.

It is convenient to redefine the osmotic coefficient as

φ =−1000 g/kg

vmM1ln a1, (23.12)

where the subscript 1 refers to the solvent.

Similarly freezing point depression is redefined as

θ = vφKfm. (23.13)

So, vφ corresponds to the empirical factor i discussed earlier.

Recall how γ was calculated from the Gibbs-Duhem equation:

ln γ± = −j −Z m

0

j

m0dm0, (23.14)

where j = 1− φ.

23.2. Theory of Electrolytic Solutions

Ionic strength is defined as

I =1

2

Xi

z2imi, (23.15)

where z is the charge of the ion and m its concentration.

Results from Debye—Hückel theory: point charge in a continuumThe Debye—Hückel equation:

ln γ± =−α |z+z−|

√I

1−Ba0√I

, (23.16)

163

Page 176: 457454 Physical Chemistry Quantum Chemistry

where

α =e3

(εkT )3/2

µ2πρ•L

1000

¶1/2, (23.17)

B =8πLe2ρ•

1000εkT, (23.18)

a0 is the radius of closest approach, e is the charge on the electron, ρ• is the

density of the pure solvent, ε is the dielectric constant for the pure solvent and L

is Avogadro’s number.

Notice that the parameters α and B depend only on the solvent.

One important approximation to this equation is to neglect the B term to get the

Debye—Hukel limiting Law (DHLL):

ln γ± = −α |z+z−|√I. (23.19)

This gives the dependence of ln γ± for dilute solutions (m → 0). It is seen that

the DHLL correctly predicts the√m dependence of ln γ±, which is observed ex-

perimentally (recall I = 12

Pi z2imi).

A useful empirical approximation is to set Ba0 = 1 and to add an empirical

correction to get the :

ln γ± =−α |z+z−|

√I

1−√I

+ 2βm

µv2+ + v2−v+ + v−

¶. (23.20)

This equation works well to ionic strengths of about I = 0.1

23.3. Ion Mobility

Current, I is given by the rate of change (in time) of charge, Q:

I =dQ

dt(23.21)

164

Page 177: 457454 Physical Chemistry Quantum Chemistry

(Electrical) work, w, is required to move a change through a potential (or voltage),

ε :

w = −εQ (23.22)

Power is given by the product of the voltage and the current:

p = −εI (23.23)

Resistance is given by the ratio of the voltage to current:

R =ε

I

Conductance is the inverse of the resistance (R−1).

Some relevant constants

• charge of an electron e = 1.602177× 10−19 C.

• Faraday’s constant F = Le = 96485 C/mol (Avogadro’s number of electrons)

23.3.1. Ion mobility

165

Page 178: 457454 Physical Chemistry Quantum Chemistry

The total current passing through an ionic solution is determined by the sum of

the current carried by the cations and by the anions

I = I+ + I− (23.24)

Now

Ii =dQi

dt= |zi| e

dNi

dt, (23.25)

where i = +,−.

For uniform ion velocity (vi) the number of ions arriving at the electrode during

any given time interval 4t is

4Ni =Ni

VAvi4t =⇒ dNi

dt=

Ni

VAvi (23.26)

so

Ii = |zi| eNi

VAvi (23.27)

Recall Coulomb’s law

Fi = zieE, (in vacuum) (23.28)

where E is the electric field, E = dεdx.

Also recall Newton’s law

Fi = mai = mdvidt= zieE. (23.29)

The moving ions experience a viscous drag f that is proportional to their velocities.

So the total force on the ions is a sum of the Coulomb force and the viscous drag

Fi = zieE − fvi (in solution). (23.30)

The ions quickly reach terminal velocity, i.e., the viscous drag equals the Coulomb

force. Hence Fi = 0.

zieE = fvi =⇒ vi =zieE

f. (23.31)

The drag f has three basic origins.

166

Page 179: 457454 Physical Chemistry Quantum Chemistry

1. Stoke’s Law type force

• “spherical” ion moving through a continuous medium

• this contribution is independent of the other ions

2. Electrophoretic effect.

• oppositely charged ions “pull” at each other

3. Relaxation effects

• solvation shell must re-adjust as ion moves. a “dressed” ion.

167

Page 180: 457454 Physical Chemistry Quantum Chemistry

A more fundamental quantity than ion velocity is the ion mobility, ui which is the

ion’s velocity per field,

ui =viE. (23.32)

For the case for parallel plate capacitors E = εl, where l is the separation of the

plates. So,

ui =vil

ε. (23.33)

Here the current carried by ion i is

Ii = |zi| eNi

VAuiε

l. (23.34)

Suppose a salt has a degree of dissociation α (α = 1 for strong electrolytes) to

produce ν+ cations and ν− anions, then each mole of salt gives: N+ = αν+Ln

and N− = αν−Ln.

The current then becomes

Ii = |zi| eανiLn

VAuiε

lF=Le= ανin |zi|uiAF

ε

V l(23.35)

It is of interest to determine the ratio of the current carried by the cation versus

the anion.

I+I−=

α/ ν+n/ |z+|u+A/ F/ εV l/

α/ ν−n/ |z−|u−A/ F/ εV l/=

=1z | ν+ |z+|ν− |z−|

u+u−

=u+u−

(23.36)

Thus the ratio of the currents is determined by simply the ratio of the mobilities.

168

Page 181: 457454 Physical Chemistry Quantum Chemistry

24. Thermodynamics of Solvation

An extremely important application of thermodynamics is to that of ion solvation.

Solvation describes how a solute dissolves in a solvent.

We will focus on ions in solution.

As a basic treatment of solvation we shall consider the solvent as a non-structural

continuum and the ion as a charged particle.

Of course this is an approximation and numerous statistical mechanical models

for solvents which incorporate a more realistic structure can be used, but we will

stick with this simple thermodynamic model.

The way to investigate the ion—solvent interaction upon solvation from a thermo-

dynamics point of view is to consider the change in the properties of the ion in a

vacuum versus the ion in solution.

Primarily we will determine 4Gv→s ≡ Gion in solv. −Gion in vac.

Since Gibbs free energy corresponds to non-PV work, 4Gv→s can be determined

by calculating the reversible work done in transferring an ion into the bulk of the

solvent.

169

169

Page 182: 457454 Physical Chemistry Quantum Chemistry

24.1. The Born Model

The Born model is a simple solvation model in which the ions are taken to be

charged spheres and the solvent is take to be a continuum with dielectric constant

εs

170

Page 183: 457454 Physical Chemistry Quantum Chemistry

4Gv→s for the Born model is obtained by considering the following contribution

to the work of ion transfer from the vacuum state to the solvated state (see figure)

• Begin with the state in which the charged sphere (the ion) is in a vacuum.

• Determine the work, wdis, done in discharging the sphere.

• Assume the uncharged sphere can pass from the (neutral) vacuum to the

neutral solvent without doing any work, wtr = 0. (This is an approximation).

• Determine the work, wch, done in charging the sphere which is now in thesolvent.

171

Page 184: 457454 Physical Chemistry Quantum Chemistry

So,

4Gv→s = wdis + wtr + wch = wdis + wch (24.1)

Work done in discharging the sphere:

The act of discharging a sphere involves bringing out to infinity from the surface

infinitesimal amounts of charge.

The work done is discharging is some what complicated since as one removes the

charge the work done in removing more charge changes according to the amount

of charge currently on the sphere.

172

Page 185: 457454 Physical Chemistry Quantum Chemistry

This is expressed mathematically as

wdis =

Z 0

ze

Z ∞

ri

σ

4π 0r2drdσ (24.2)

=

Z 0

ze

σ

4π 0ridσ

= − (ze)2

8π 0ri,

where z is the oxidation state of the ion, e is the charge of the electron, ri is the

radius of the sphere (ion) and 0 is the permittivity of free space.

Work done in charging the sphere:

The only difference in charging the sphere is that the sign of the work will be differ-

ent and that since we are charging in a solvent we must multiply the permittivity

of free space by the dielectric constant of the solvent.

So,

wch = +(ze)2

8π 0εsri(24.3)

24.1.1. Free Energy of Solvation for the Born Model

Combining the above two expression for work gives

4Gv→s = − (ze)2

8π 0ri+

(ze)2

8π 0εsri(24.4)

=(ze)2

8π 0ri

µ1

εs− 1¶

The above expression is 4Gv→s/ion. For n moles of ions (nL = N)

4Gv→s =N (ze)2

8π 0ri

µ1

εs− 1¶

(24.5)

173

Page 186: 457454 Physical Chemistry Quantum Chemistry

The dielectric constant of any solvent is always greater than unity so 1εs− 1 is

always negative hence 4Gv→s < 0. Thus ions always exist more stably in solution

than in a vacuum.

24.1.2. Ion Transfer Between Phases

We can quickly generalize the Born model to describe ion transfer between phases

in a solution of two immiscible phases

Consider an immiscible solution of two phases α and β having dielectric constants

εα and εβ.

Since Gibbs free energy is a state function we can write the change in free energy

for transfer of an ion form the β phase to the α phase as

4Gβ→α =

=−4Gv→βz | 4Gβ→v +4Gv→α (24.6)

= −N (ze)2

8π 0ri

µ1

εβ− 1¶+

N (ze)2

8π 0ri

µ1

εα− 1¶

=N (ze)2

8π 0ri

µ1

εα− 1

εβ

¶The Partition Coefficient

We can now write the partition coefficient for the Born model as

Pα/βi = e

−4Gªβ→α

nRT = e− L(ze)2

8πri 0RT1εα− 1εβ (24.7)

24.1.3. Enthalpy and Entropy of Solvation

We may employ the standard thermodynamic relations which we have derived

earlier to obtain the entropy and enthalpy for the Born model.

174

Page 187: 457454 Physical Chemistry Quantum Chemistry

From µ∂G

∂T

¶P

= −S ⇒µ∂4Gv→s

∂T

¶P

= −4Sv→s, (24.8)

we find entropy to be

4Sv→s = −∂

∂T

"N (ze)2

8π 0ri

µ1

εs− 1¶#

. (24.9)

The only variable in the above equation that has a temperature dependence is the

dielectric constant of the solvent so,

4Sv→s = −N (ze)2

8π 0ri

∂T

µ1

εs

¶=

N (ze)2

8π 0riε2s

∂εs∂T

. (24.10)

Enthalpy is obtained via the relation:

4Hv→s = 4Gv→s + T4Sv→s (24.11)

=N (ze)2

8π 0ri

µ1

εs− 1¶+

N (ze)2 T

8π 0riε2s

∂εs∂T

=N (ze)2

8π 0ri

µ1

εs+

T

ε2s

∂εs∂T− 1¶

24.2. Corrections to the Born Model

The Born model is very valuable because of its simplicity–qualitative statements

about solvation and ion transfer between phases can be made.

Unfortunately however, the Born model does not make quantitatively correct pre-

dictions in many cases.

We simply list here several phenomena that more sophisticated theories of solva-

tion must consider

175

Page 188: 457454 Physical Chemistry Quantum Chemistry

1. The solvophobic effect: a cavity must form in the solvent to accommodate

the ion.

2. Changes in solvent structure: the local environment of the ion has a different

arrangement of solvent molecules than that of the bulk solvent, so the initial

structure of the solvent must breakdown and the new structure must form.

3. Specific interactions: any interaction energy specific to the particular ion-

solvent pair: Hydrogen bonding being the prime example.

4. Annihilation of defects: A small ion may be captured in a micro-cavity

within the solvent releasing the energy of the micro-cavity defect.

176

Page 189: 457454 Physical Chemistry Quantum Chemistry

25. Key Equations for Exam 4

Listed here are some of the key equations for Exam 4. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• Some thermodynamic relations

H = U + PV dH = TdS + V dP

A = U − TS dA = −SdT − PdV

G = H − TS dG = −SdT + V dP

• The chemical potential equation

μi = μªi +RT ln ai (25.1)

• The 4G equation (this should be posted on your refrigerator)

4G = 4Gª +RT lnQ. (25.2)

177

177

Page 190: 457454 Physical Chemistry Quantum Chemistry

At equilibrium 4G = 0 and

4Gª = −RT lnKa (25.3)

• For an ideal gasCPm = Cvm +R (25.4)

• The Debye—Hukel limiting Law (DHLL):

ln γ± = −α |z+z−|√I. (25.5)

• The ratio of the current carried by the cation versus the anion in terms ofion mobility is

I+I−=

u+u−

(25.6)

• The chemical potential equation

μi = μªi +RT ln ai (25.7)

• The 4G equation (this should be posted on your refrigerator)

4G = 4Gª +RT lnQ. (25.8)

At equilibrium 4G = 0 and

4Gª = −RT lnKa (25.9)

• 4G for the Born model:

4Gv→s =N (ze)2

8π 0rs

µ1

εs− 1¶

(25.10)

• 4G for transfer of an ion form the β phase to the α phase,

4Gβ→α =N (ze)2

8π 0ri

µ1

εα− 1

εβ

¶(25.11)

178

Page 191: 457454 Physical Chemistry Quantum Chemistry

Chemistry 352: PhysicalChemistry II

179

179

Page 192: 457454 Physical Chemistry Quantum Chemistry

Part V

Quantum Mechanics andDynamics

180

180

Page 193: 457454 Physical Chemistry Quantum Chemistry

26. Particle in a 3D Box

We now return to quantum mechanics and investigate some of the important

models that we omitted from the first semester.

In particular we will look at the particle in a box in more than one dimension.

We will also solve models which deal with rotations.

26.1. Particle in a Box

Recall that the important ideas from the 1D particle in a box problem were

The potential, V (x), is given by

V (x) =

⎧⎪⎨⎪⎩∞ x ≤ 00 0 < x < a

∞ x ≥ a

. (26.1)

Because of the infinities at x = 0 and x = a, we need to partition the x-axis into

the three regions shown in the figure.

181

181

Page 194: 457454 Physical Chemistry Quantum Chemistry

Now, in region I and III, where the potential is infinite, the particle can never

exist so, ψ must equal zero in these regions.

The particle must be found only in region II.

The Schrödinger equation in region II is (V (x) = 0)

Hψ = Eψ =⇒ −~2

2m

d2ψ(x)

dx2= Eψ, (26.2)

The general solution of this differential equation is

ψ(x) = A sin kx+B cos kx, (26.3)

where k =q

2mE~2 .

Now ψ must be continuous for all x. Therefore it must satisfy the boundary

conditions (b.c.): ψ(0) = 0 and ψ(a) = 0.

From the ψ(0) = 0 b.c. we see that the constant B must be zero because

cos kx|x=0 = 1.So we are left with ψ(x) = A sin kx for our wavefunction.

182

Page 195: 457454 Physical Chemistry Quantum Chemistry

The second b.c., ψ(a) = 0, places certain restrictions on k.

In particular,

kn =nπ

a, n = 1, 2, 3, · · · . (26.4)

The values of k are quantized. So, now we have

ψn(x) = A sinnπx

a. (26.5)

The constant A is the normalization constant.

Solving for A gives

A =

r2

a. (26.6)

Thus our normalized wavefunctions for a particle in a box are (in region II)

ψn(x) =

r2

asin

nπx

a. (26.7)

We found the energy levels to be

En =n2π2~2

2ma2~= h

2π=n2h2

8ma2. (26.8)

26.2. The 3D Particle in a Box Problem

We now consider the three dimensional version of the problem.

The potential is now

V (x, y, z) =

(0, 0 < x < a, 0 < y < b, 0 < z < c

∞, else. (26.9)

183

Page 196: 457454 Physical Chemistry Quantum Chemistry

Now the Schrödinger equation is

Hψ = Eψ ⇒ −~22m∇2ψ = Eψ

⇒ −~22m

µ∂2ψ

∂x2+

∂2ψ

∂y2+

∂2ψ

∂z2

¶= Eψ. (26.10)

It is generally true that when the Hamiltonian is a sum of independent terms, we

can write the wavefunction as a product of wavefunctions

ψ(x, y, z) = ψx(x)ψy(y)ψz(z). (26.11)

This lets us perform a mathematical trick which is sometimes useful in solving

partial differential equations.

Subbing the product wavefunction into the Schrödinger equation we get

−~22m

µ∂2ψxψyψz

∂x2+

∂2ψxψyψz

∂y2+

∂2ψxψyψz

∂z2

¶= Eψxψyψz (26.12)

−~22m

µψyψz∂

2ψx

∂x2+

ψxψz∂2ψy

∂y2+

ψxψy∂2ψz

∂z2

¶= Eψxψyψz.

We now divide both sides by ψxψyψz to get

−~22m

µ1

ψx

∂2ψx

∂x2+1

ψy

∂2ψy

∂y2+1

ψz

∂2ψz

∂z2

¶= E. (26.13)

This equation is now of the form

f(x) + g(y) + h(z) = C, (26.14)

where C is a constant.

If we take the derivative with respect to x we get

d

dx

→f(x) + g(y) + h(z) = C,

df(x)

dx+

dg(y)

dx+

dh(z)

dx=

dC

dx,

df(x)

dx= 0, (26.15)

184

Page 197: 457454 Physical Chemistry Quantum Chemistry

So, f(x) is a constant. Similarly for g(y) and h(z)

Applying this to our Schrödinger equation means that we have converted our

partial differential equation into three independent ordinary differential equations,

−~22m

1

ψx

d2ψx

dx2= Ex =⇒

−~22m

d2ψx

dx2= Exψx (26.16)

−~22m

1

ψy

d2ψy

dy2= Ey =⇒

−~22m

d2ψy

dy2= Eyψy

−~22m

1

ψz

d2ψz

dz2= Ez =⇒

−~22m

d2ψz

dz2= Ezψz

which we recognize as the 1D particle in a box equations.

Hence we immediately have

ψx =

r2

asin

nxπx

a, (26.17)

ψy =

r2

bsin

nyπy

b,

ψz =

r2

csin

nzπz

c

and

Ex,nx =n2xh

2

8ma2, (26.18)

Ey,ny =n2yh

2

8mb2,

Ez,nz =n2zh

2

8mc2.

The total wavefunction is

ψ =2√2√

abcsin

nxπx

asin

nyπy

bsin

nzπz

c(26.19)

and the total energy is

E = Ex,nx +Ey,ny +Ez,nz . (26.20)

185

Page 198: 457454 Physical Chemistry Quantum Chemistry

DegeneracyThe 3D particle in a box model brings up the concept of degeneracy.

When n(> 1) states have the same total energy they are said to be n-fold degen-

erate.

Let the 3D box be a cube (a = b = c) then the states

(nx = 2, ny = 1, nz = 1), (26.21)

(nx = 1, ny = 2, nz = 1),

(nx = 1, ny = 1, nz = 2)

have the same total energy and thus are degenerate.

186

Page 199: 457454 Physical Chemistry Quantum Chemistry

27. Operators

27.1. Operator Algebra

We now take a mathematical excursion and discuss the algebra of operators.

Definitions

• Function: A function, say f , describes how a dependent variable, say y, is

related to an independent variable, say x: y = f(x)

— e.g., y = x2, y = sinx, etc.

• Operator: An operator, say O, transforms a function, say f , into another

function, say g: Of(x) = g(x).

• Algebra: An algebra is a specific collection of rules applied to a set of objectsand a particular operation

— Rules

∗ Transitivity∗ Associativity∗ Existence of an identity∗ Existence of an inverse

— e.g., Addition on the set of real numbers, Multiplication on the set ofreal numbers

187

187

Page 200: 457454 Physical Chemistry Quantum Chemistry

— Note: Commutivity is not a requirement of an algebra

∗ example 1: multiplication on the set of real number is commutive:ab = ba

∗ example 2: multiplication on the set of n× n matrices is not com-

mutive: ab 6= ba in general. e.g.,"1 0

2 1

#"3 1

1 1

#=

"3 1

7 3

#(27.1)

but "3 1

1 1

#"1 0

2 1

#=

"5 1

3 1

#6="3 1

7 3

#(27.2)

Algebraic rules for operators

1. Equality:

if α = β, then αf(x) = g(x) = βf(x) (27.3)

2. Addition:

if αf(x) = g(x) and βf(x) = h(x), (27.4)

then (α+ β)f(x) = αf(x) + βf(x) = g(x) + h(x)

3. Multiplication:

αβf(x) = α³βf(x)

´(27.5)

βαf(x) = β (αf(x)) ,

but in general αβf(x) 6= βαf(x).

4. Inverse:

if αf(x) = g(x) and βg(x) = f(x) (27.6)

then β = α−1 and is said to be α inverse

188

Page 201: 457454 Physical Chemistry Quantum Chemistry

Linear operators:

• A special and important class of operators

• They obey all of the above properties in addition to

— α (f(x) + g(x)) = αf(x) + αg(x), and

— α(λf(x)) = λαf(x), where λ is a complex number.

Hermitian operators:

• A special class of linear operators

• All observables in quantum mechanics are associated with Hermitian oper-

ators

• The eigenvalues of Hermitian operators are real

Some important operators

1. • x: xf(x) = xf(x)

• d: df(x) = ddxf(x)

• d2: d2f(x) = d³df(x)

´= d

¡ddxf(x)

¢= d

dx

¡ddxf(x)

¢= d2

dx2f(x)

• ı: ıf(x, y, z) = f(−x,−y,−z)

• ∇: ∇f(x, y, z) =³

∂∂xex +

∂∂yey +

∂∂zez´f(x, y, z)

• ∇2: ∇2f(x, y, z) =³

∂2

∂x2+ ∂2

∂y2+ ∂2

∂z2

´f(x, y, z)

Commutators:

We have seen that in general αβ 6= βα. This leads to the construction of the

commutator, [, ]: hα, β

i≡ αβ − βα. (27.7)

189

Page 202: 457454 Physical Chemistry Quantum Chemistry

If αβ = βα, thenhα, β

i= 0 and α and β are said to commute with one another.

The eigenvalue equation:

If αf(x) = g(x) and g(x) = af(x), then the operator equation, αf(x) = g(x)

becomes the eigenvalue equation

αf(x) = af(x). (27.8)

The eigenvalue equation is of fundamental importance in quantum theory. We

shall see that eigenvalues of certain operator can be identified as experimental

observables.

Commuting operators and simultaneous sets of eigenfunctions.

If αf(x) = af(x) and β and α commute, then βf(x) = bf(x).

The proof goes as follows: On the one hand,

β (αf) = β (af) = aβf (27.9)

because f is an eigenfunction of α.

On the other hand,

β (αf) = α³βf´

(27.10)

because β and α commute.

Thus

α³βf´= a

³βf´. (27.11)

which states that βf is an eigenfunction of α with eigenvalue a. The only way for

this to be true is if βf = bf.

190

Page 203: 457454 Physical Chemistry Quantum Chemistry

27.2. Orthogonality, Completeness, and the Superposition

Principle

Theorem 1: The eigenfunctions of a Hermitian operator corresponding to differ-ent eigenvalues are orthogonal:Z

spaceψ∗jψk = 0, j 6= k. (27.12)

Theorem 2: The eigenfunctions of a Hermitian operator form a complete set

Corollary (the superposition principle): Any arbitrary function ψ in the

space of eigenfunctions ϕi can be written as a superposition of these eigenfunc-

tions:

ψ =Xi

aiϕi (27.13)

191

Page 204: 457454 Physical Chemistry Quantum Chemistry

28. Angular Momentum

We will encounter several different types of angular momenta, but fortunately

they are all described by a single theory

Before starting with the quantum mechanical treatment of angular momentum,

we first review the classical treatment.

28.1. Classical Theory of Angular Momentum

The classical angular momentum, L, is given by

L = x× p (28.1)

The vector cross-product can be computed by finding the following determinant:

L =

¯¯ ex ey ez

x y z

px py pz

¯¯ = Lxz |

(ypz − zpy)ex +

Lyz | (zpx − xpz)ey +

Lzz | (xpy − ypx)ez (28.2)

Hence,

Lx = (ypz − zpy) , (28.3)

Ly = (zpx − xpz) , (28.4)

Lz = (xpy − ypx) . (28.5)

Another quantity that we will find useful is

L2 = L · L = L2x + L2y + L2z (28.6)

192

192

Page 205: 457454 Physical Chemistry Quantum Chemistry

28.2. Quantum theory of Angular Momentum

So, in accordance with postulate II, we replace the classical variables with their

operators. That is,

Lx = (ypz − zpy) =~i

µy∂

∂z− z

∂y

¶, (28.7)

Ly = (zpx − xpz) =~i

µz∂

∂x− x

∂z

¶, (28.8)

Lz = (xpy − ypx) =~i

µx∂

∂y− y

∂x

¶. (28.9)

Recall the basic commutators. ∙∂

∂u, u

¸= 1, (28.10)∙

∂u, v

¸= 0,

where u, v = x, y, or z and u 6= v.

From these basic commutators one can derivehLx, Ly

i= i~Lz,

hLy, Lz

i= i~Lx,

hLz, Lx

i= i~Ly (28.11)

and hL2, Lx

i=hL2, Ly

i=hL2, Lz

i= 0 (28.12)

It is often convenient to express the angular momentum operators in spherical

polar coordinates as follows.

Lx = i~µsinφ

∂θ+ cot θ cosφ

∂φ

¶, (28.13)

Ly = −i~µcosφ

∂θ− cot θ sinφ ∂

∂φ

¶, (28.14)

193

Page 206: 457454 Physical Chemistry Quantum Chemistry

Lz = −i~∂

∂φ(28.15)

L2 = −~2µ

∂2

∂θ2+ cot θ

∂θ+

1

sin2 θ

∂2

∂φ2

¶(28.16)

28.3. Particle on a Ring

Consider a particle of mass μ confined to move on a ring of radius R.

The moment of inertia is I = μR2

The Hamiltonian is given by

H =L2z2I=−~22I

d2

dφ2(28.17)

(note that we use d rather than ∂ since the problem is one-dimensional).

The Schrödinger equation becomes

−~22I

d2ψ

dφ2= Eψ (28.18)

Notice that this Schrödinger equation is exactly the same form as the particle in

a box. The only difference is the boundary conditions.

The boundary condition for the particle in a box were ψ was zero outside the box.

Now the boundary condition is that ψ(φ) = ψ(φ + 2π). The wavefunction must

by 2π periodic.

The allowable wavefunctions are

ψm(φ) =

⎧⎪⎨⎪⎩A cosmφ

A sinmφ

Aeimφ

, (28.19)

194

Page 207: 457454 Physical Chemistry Quantum Chemistry

m = 0, ±1, ±2, ±3, . . .These wavefunctions are really the “same.” It will be most convenient to use

ψm(φ) = Aeimφ as our wave functions.

Plugging ψm(φ) = Aeimφ into the Schrödinger equation gives

−~22I

d2Aeimφ

dφ2= EmAe

imφ (28.20)

~2m2

2IAeimφ = EmAe

imφ

Therefore the energy levels (the eigenvalues) for a particle in a ring are

Em =~2m2

2I=

m2h2

8π2I. (28.21)

Next we need to find the normalization constant, A.

1 =

Z 2π

0

ψ∗ψdφ (28.22)

1 =

Z 2π

0

A2e−imφeimφdφ

1 = A2Z 2π

0

dφ = 2πA2,

thus

A =

r1

2π. (28.23)

Hence the normalized wavefunctions for a particle on a ring are

ψ =1√2π

eimφ. (28.24)

28.4. General Theory of Angular Momentum

To discuss angular momentum in a more general way it is convenient to define

two so-called ‘ladder’ operators

L+ ≡ Lx + iLy (28.25)

195

Page 208: 457454 Physical Chemistry Quantum Chemistry

and

L− ≡ Lx − iLy (28.26)

We collect here the commutators of L+ and L−:hLz, L+

i= L+ ⇒ L+Lz = LzL+ − L+ (28.27)h

Lz, L−

i= −L− ⇒ L−Lz = LzL− + L− (28.28)

Now, since Lz and L2 commute there must exist a set of simultaneous eigenfunc-

tions ψiLzψi = mψi (28.29)

and

L2ψi = k2ψi (28.30)

Physically, k~ represents the length of the angular momentum vector and m~represents the projection onto the z-axis. (Note: for simplicity in writing we are‘hiding’ the ~ in the wavefunctions.)

On these physical grounds we conclude |m| ≤ k, i.e., k sets an upper and lower

limit on m.

Let’s define the maximum value of m to be a new quantum number l ≡ mmax.

(Thus l ≤ k).

And let’s define the minium value of m to be a new quantum number l0 ≡ mmin.

(Thus −l0 ≤ k)

Now, at least one of the eigenfunctions in the set ψi yields the eigenvalue mmax

(or l) when operated on by Lz. Let’s call that eigenfunction ψl;

Lzψl = lψl. (28.31)

Now we can operate on both sides of this equation with L−:

L−Lzψl = L−lψl (28.32)

196

Page 209: 457454 Physical Chemistry Quantum Chemistry

Using the commutator relation L−Lz = LzL− + L− we get³LzL− + L−

´ψl = lL−ψl (28.33)

LzL−ψl + L−ψl = lL−ψl

Bringing the second term on the left hand side over to the right hand side gives

LzL−ψl = lL−ψl − L−ψl (28.34)

LzL−ψl| z ψl−1

= (l − 1)L−ψl| z ψl−1

We see that L−ψl ≡ ψl−1 is in fact an eigenfunction of Lz (with associated eigen-

value (l − 1)) and is thus a member of ψi .

The eigenfunction ψl−1 has an associated eigenvalue that is one unit less then the

maximum value.

The above procedure can be repeated n times so that Ln−ψl = ψl−n provided n

does not exceed l − l0.

The eigenfunction ψl−n has an associated eigenvalue that is n units less then the

maximum value, i.e.,

Lzψl−n = (l − n)ψl−n. (28.35)

The largest value of n is l − l0. For that case,

Lzψl0 = (l − l + l0)ψl0 = l0ψl0 . (28.36)

Similar behavior is seen for the operator L+, except in the opposite direction–the

eigenvalue is increased by one unit for each action of L+. For example

L+Lzψl0 = L+l0ψl0 (28.37)³

LzL+ − L+´ψl0 = l0L+ψl0

LzL+ψl0 = (l0 + 1)L+ψl0 .

197

Page 210: 457454 Physical Chemistry Quantum Chemistry

The raising and lowering nature of L+ and L− is why they are called ladder

operators.

We can not act with L+ and L− indefinitely since we are limited by l–we reach

the ends of the ladder. This requires that

L−ψl0 = 0 (28.38)

(we can’t go lower than the lowest step) and

L+ψl = 0 (28.39)

(we can’t go higher than the highest step).

Often times the ladder operators appear in tandem either as L−L+ or L+L− so it

is useful list some identities for these products

L−L+ = L2 − L2z − Lz (28.40)

and

L+L− = L2 − L2z + Lz (28.41)

We can use these identities to derive a relation between the quantum numbers k

and l.

We begin with

L−L+ψl = L−³L+ψl

´= 0, (28.42)

but from the first of the above identities

L−L+ψl =³L2 − L2z − Lz

´ψl = (k

2 − l2 − l)ψl (28.43)

Therefore

k2 − l2 − l = 0⇒ k =pl(l + 1). (28.44)

198

Page 211: 457454 Physical Chemistry Quantum Chemistry

We we can also consider

L+L−ψl0 = L+³L−ψl0

´= 0 (28.45)

and

L+L−ψl0 =³L2 − L2z + Lz

´ψl0 = (k

2 − l02 + l0)ψl0 . (28.46)

substituting in the relation we just found for k gives

l(l + 1)− l02 + l0 = 0; (28.47)

simplifying gives

l = −l0 (28.48)

Thus mmax = l, mmin = −l and so m = l, l − 1, l − 2, . . . , −l + 1 ,−l.

This also implies that the number of ‘rungs’ is 2l+1 and that l must be either an

integer or a half-integer.

28.5. Quantum Properties of Angular Momentum

The eigenfunctions of angular momentum are entirely specified by two quantum

numbers l and m: ψlm.

L2ψlm = l(l + 1)ψlm Lzψlm = mψlm (28.49)

If we write out the first of these explicitly in spherical polar coordinates as a

partial differential equation we obtain

∂2ψlm

∂θ2+ cot θ

∂ψlm

∂θ+

1

sin2 θ

∂2ψlm

∂φ2+ l(l + 1)ψlm = 0 (28.50)

The solutions to this partial differential equation are known to be the spherical

harmonic functions

ψlm = Ylm(θ, φ). (28.51)

199

Page 212: 457454 Physical Chemistry Quantum Chemistry

The spherical harmonics are functions of two variables, but they are a product of

a function only of θ and a function only of φ,

ψlm = Ylm(θ, φ) = AP|m|l (θ)eimφ, (28.52)

where the P |m|l (θ) are the Legendra polynomials and A is normalization constant.

Both the spherical harmonics and the Legendra polynomials are tabulated. They

are also built-in functions of Mathematica.

The spherical harmonics (and hence the angular momentum wavefunctions) are

orthonormal; meaning,Z 2π

0

Z π

0

Y ∗l0m0(θ, φ)Ylm(θ, φ) sin θdθdφ =

(1 l0 = l and m0 = m

0 l0 6= l or m0 6= m(28.53)

28.5.1. The rigid rotor

Rotational energyFor general rotation in three dimensions the is

H =~2

2IL2, (28.54)

so the Schrödinger equation is

Hψlm = Elmψlm ⇒~2

2IL2ψlm = Elmψlm ⇒

~2

2Il(l + 1)ψlm = Elmψlm. (28.55)

Thus

Elm =l(l + 1)~2

2I=

l(l + 1)h2

8π2I= El. (28.56)

There is no m dependence for the energy. In other words, the energy levels are

determined only by the value of l.

We know that there are 2l+1 different m values for a particular l value. All 2l+1

of these wavefunctions correspond to the same energy. We say the there is a 2l+1

degeneracy of the energy levels.

200

Page 213: 457454 Physical Chemistry Quantum Chemistry

29. Addition of Angular Momentum

29.1. Spin Angular Momentum

We learned above that l may take on integer or half-integer values.

Systems in which l takes on half-integer values are peculiar.

These systems have no classical analogs.

One example of such a system is the spin of an electron, l = s = 1/2. The values

of m = ms are limited to +1/2 and −1/2.

One peculiarity of this system is that the wavefunctions are 4π periodic (and 2π

antiperiodic):

ψs(θ) = −ψs(θ + 2π) (29.1)

and

ψs(θ) = ψs(θ + 4π). (29.2)

That means that the system has to ‘rotate’ twice (in spin space not coordinate

space) to get back to its original state.

∗ ∗ ∗ See in-class demonstration: the belt trick ∗ ∗∗

201

201

Page 214: 457454 Physical Chemistry Quantum Chemistry

29.2. Addition of Angular Momentum

In atoms the are a number of sources of angular momentum: The l’s and s’s of

each of the electrons.

One measures, however, the total angular momentum, J.

The electrons in many electron atoms couple. The are two main coupling schemes

which account for the total angular momentum of the atom.

1. LS coupling (also called Russell-Saunders coupling)

• works well for low atomic weight atoms (first couple of rows of the

periodic table)

• find the total spin angular momentum S =Ms,max, (Ms =P

imsi)

• find the total orbital angular momentum L = Mmax, (M =P

imi)

• then J = L+ S

2. jj coupling

• applies to higher atomic weight atoms

• find subtotal angular momentum for each electron ji = li + si

• then find total angular momentum by J =P

i ji.

• we will not use this method.

29.2.1. The Addition of Angular Momentum: General Theory

Consider two sources of angular momentum for a system represented by the op-

erators J1 and J2 (J1 and J2 could be L or S angular momentum; we use J when

we speak generally.)

202

Page 215: 457454 Physical Chemistry Quantum Chemistry

The total angular momentum is JT = J1 + J2.

The total z-component of the angular momentum is JzT = Jz1 + Jz2

The last statement implies that the orientation quantum number of the total

system is simple the sum of that for the components

M = m1 +m2 (29.3)

We need to determine the allowed values of the total angular momentum quantum

number J.

The maximum value of J is determined by the maximum value of M by

Jmax =Mmax = m1max +m2max = j1 + j2 (29.4)

This corresponds to a situation in which component angular momentums add in

the most favorable manner

The minimum value of J is determined by the case when the components add in

the least favorable manner. That is,

Jmin = |j1 − j2| . (29.5)

The total angular momentum is quantized is exactly the same manner as any

other angular momentum. Thus the allowed values of J are

J = j1 + j2, j1 + j2 − 1, . . . , |j1 − j2|+ 1, |j1 − j2| . (29.6)

29.2.2. An Example: Two Electrons

The table below shows the total spin angular momentum S for a two electron

system

203

Page 216: 457454 Physical Chemistry Quantum Chemistry

spin state ms1 ms1 MS S

α(1)α(2) 12

12

1 1

β(1)β(2) −12−12−1 1

α(1)β(2) + β(1)α(2) 0 0 0 1

α(1)β(2)− β(1)α(2) 0 0 0 0

Counting states:

The spin degeneracy, gS, of the states is given by 2S + 1. In the above example

the degeneracy is gS = 3 for the S = 1 states and gS = 1 for the S = 0 states.

29.2.3. Term Symbols

We have already seen several term symbols, those being 1S and 3S during our

discussion of helium.

Term symbols are simply shorthand notion used to identify states. Term symbols

are useful for predict and understanding spectroscopic data. So, it is worthwhile

to briefly discuss them.

In general the term symbol is simply notates the total orbital angular momentum

and spin degeneracies of a particular set of states (or a state in the case of a singlet

state).

The orbital degeneracy is given by gL = 2L+ 1.

For historical reasons L values are associated with a letter like the l values of a

hydrogenic system are.

L 0 1 2 3 4 5

symbol S P D F G H.

204

Page 217: 457454 Physical Chemistry Quantum Chemistry

The term symbol for a particular states is constructed from the following general

templategSLJ .

Many electron atoms have term symbols associated with their states.

Rules:

1. All closed shells have zero spin and orbital angular momentums: L = 0,

S = 0. These states are all singlet S states, notated by 1S

2. An electron and a “hole” lead to equivalent term symbols.

• E.g., p1 and p5 have the same term symbol.

3. Hund’s Rule for the ground state only.

1. The ground state will have maximum multiplicity.

2. If several terms have the same multiplicity then ground state will be

that of the largest L.

3. Lowest J value (regular) “electron”, Highest J value (inverted) “hole”

29.2.4. Spin Orbit Coupling

A charge possessing angular momentum has a magnetic dipole associated with it.

An electron has orbital and spin magnetic dipoles.

These dipoles interact with a certain spin—orbit interaction energy ESO.

The spin—orbit Hamiltonian is

HSO = hcA[L · S (29.7)

HSO =hcA

2

³J2 − L2 − S2

´,

205

Page 218: 457454 Physical Chemistry Quantum Chemistry

where A is the spin—orbit coupling constant.

From the Hamiltonian the spin—orbit interaction energy is

ESO =hcA

2[J(J + 1)− L(L+ 1)− S(S + 1)] (29.8)

206

Page 219: 457454 Physical Chemistry Quantum Chemistry

30. Approximation Techniques

As we learned last semester, there are very few models for which we can obtain

an exact solution.

Consequently we must be satisfied with using approximation methods.

Last semester, we always took the simplest approximation to give the qualitative

properties of the unsolvable system.

Now we will consider two important quantitative approximation methods: (i)

perturbation theory and (ii) variational theory

30.1. Perturbation Theory

The basic procedure of perturbation theory

• Find a solvable system that is similar to the system at hand.

• Treat the difference between the two systems as a perturbation to the solv-able system

• Use the solvable system’s wavefunctions as a zeroth order approximation tothe wavefunctions for the unsolvable system.

• These wavefunctions are used to find a first order correction to the energy.

207

207

Page 220: 457454 Physical Chemistry Quantum Chemistry

• The first order energy is then used to make a first order approximation tothe wavefunction.

• The procedure is repeated to get higher and higher order approximations.

This process get algebraically intensive so we will only go as far as listing the first

order energy correction.

The nth state energy in perturbation theory:

En = E(0)n +E(1)

n + . . . , (30.1)

where E(0)n is the nth state energy for the unperturbed (solvable) system and E

(1)n

is the first order correction. This is given by

E(1)n =

Zallspace

ψ(0)∗n H(1)ψ(0)n dx, (30.2)

where H(1) is the first order correction to the Hamiltonian–the perturbation.

Example: the quartic oscillator

• Consider the quartic oscillator described by the potential V (x) = 12kx2+ax4

where a is very small and can be treated as a perturbation.

• The obvious solvable system is the harmonic oscillator:

H = − ~2

2m

d2

dx2+1

2kx2. (30.3)

This has energy levelsEn = ~ω(n+12) and wavefunctionsAnHn(

√αx)e−αx

2/2,

where α =q

km~

• The perturbative part of the Hamiltonian is

H(1) = ax4. (30.4)

208

Page 221: 457454 Physical Chemistry Quantum Chemistry

• For example, the ground state energy correction is then calculation from

E(1)0 =

Z ∞

−∞ψ(0)∗0 H(1)ψ

(0)0 dx (30.5)

=

Z ∞

−∞A0e

−αx2/2ax4A0e−αx2/2dx

= aA20

Z ∞

−∞x4e−αx

2

dx

=3√πaA20

4α52

,

so the first order ground state energy for a quartic oscillator is

E0 '~ω2+3√πaA20

4α52

.

30.2. Variational method

The basic idea behind the variational method is to use a trial wavefunction with

an adjustable parameter. The value of the parameter which minimizes the energy,

Etrial, gives a trial wavefunction which is closest to the real wavefunction.

The basis for this is the variation theorem which states

Etrial ≥ E.

We will not prove this theorem here.

The trial energy is calculated by

Etrial =

Rallspace

ψ∗trialHψtrialdxRallspace

ψ∗trialψtrialdx(30.6)

The trial energy is now a function of the adjustable parameter, p, that we use to

minimize the trial energy by setting

dEtrialdp

= 0 (30.7)

209

Page 222: 457454 Physical Chemistry Quantum Chemistry

and solving for p. (Strictly speaking we should check that we have a minimum

and not a maximum or inflection point, but with reasonably good trial functions

one is pretty safe in having a minimum.)

210

Page 223: 457454 Physical Chemistry Quantum Chemistry

31. The Two Level System and

Quantum Dynamics

Our entire discussion of quantum mechanics thus far had dealt only with time

independent quantum mechanics.

The time variable never appears in any expression.

Obviously there are cases where quantum objects move with time. For example,

firing an electron down a particle accelerator.

We shall finally get to quantum dynamics in this chapter, but first we will discuss

the very important model of the two level system.

31.1. The Two Level System

If the harmonic oscillator is the most important model in all a physics, the two

level system is a close second.

The spin system discussed above is an example of a two level system.

The two level system is inherently quantum mechanical in nature. Unlike the

harmonic oscillator it has no classical analogue.

211

211

Page 224: 457454 Physical Chemistry Quantum Chemistry

Consequently, we can not use our usual procedure of writing down the classical

Hamiltonian and then replacing the variables with their corresponding operators.

The two level system consists of two states ψ1 and ψ2 separated by energy 4 =

2 − 1 as shown below

The states ψ1 and ψ2 are orthonormal:ZTLS

ψ∗jψkdΩ =

(1 j = k

0 j 6= k, (31.1)

whereRTLS dΩ means integration over the two level space (which is really just the

sumP2

i=1).

The states ψ1 and ψ2 are eigenfunctions of the two level Hamiltonian,

H = 1δ1,° + 2δ2,°, (31.2)

where δj,° “projects out” the jth state of the wavefunction being acted on.

212

Page 225: 457454 Physical Chemistry Quantum Chemistry

For example let some arbitrary wavefunction ψ = aψ1 + bψ2, then

Hψ = ( 1δ1,° + 2δ2,°) (aψ1 + bψ2) (31.3)

= 1δ1,° (aψ1 + bψ2) + 2δ2,° (aψ1 + bψ2)

= a 1ψ1 + b 2ψ2

Another orthonormal set of wavefunctions are the so-called ‘left’

ψL =1√2ψ1 +

1√2ψ2 (31.4)

and ‘right’

ψR =1√2ψ1 −

1√2ψ2 (31.5)

states.

We can invert above equations and solve for ψ1 and ψ2 in terms of ψL and ψR

ψ1 =1√2ψL +

1√2ψR (31.6)

and

ψ2 =1√2ψL −

1√2ψR. (31.7)

213

Page 226: 457454 Physical Chemistry Quantum Chemistry

31.2. Quantum Dynamics

So far we have been concerned with the eigenfunctions and eigenvalues (energy

levels) of the various quantum systems that we have discussed.

What has been kept hidden up to now is the fact that the eigenfunctions are really

multiplied by a phase factor of the form .

Ψn(x, t) ≡ ψn(x)e− i~ Ent (31.8)

We can verify this by obtaining the time independent Schrödinger equation from

the more general time dependent

i~∂Ψn(x, t)

∂t= HΨn(x, t) (31.9)

i~∂ψn(x)e

− i~ Ent

∂t= Hψn(x)e

− i~ Ent

i~ψn(x)∂e−

i~ Ent

∂t= Hψn(x)e

− i~ Ent

i~ψn(x)

µ− i

~En

¶e−

i~ Ent = Hψn(x)e

− i~ Ent

Enψn(x)e− i~ Ent = e−

i~ EntHψn(x)

Enψn(x) = Hψn(x) (31.10)

Does this mean the eigenstates are not stationary states? To determine this we

need to calculate the probability of finding the particle in the same eigenstate at

some future time. This is given by

P (x, t) =

¯ZΨ∗n(x, 0)Ψn(x, t)dx

¯2(31.11)

=

¯Zψ∗n(x)ψn(x)e

− i~ Entdx

¯2=

¯e−

i~ Ent

Zψ∗n(x)ψn(x)dx

¯2=

¯e−

i~ Ent(1)

¯2= 1,

214

Page 227: 457454 Physical Chemistry Quantum Chemistry

so no matter what time t we check we will always find the system in the same

eigenstate. Thus the eigenstates are stationary states.

In general the state of the system need not be in one particular eigenstate; it may

be in a superposition of any number of eigenstates.

The “left” and “right” wavefunctions that we saw in the discussion of the two

level system are examples of superposition states.

The phase factor does become important for superposition states.

As an example consider the state

Φ(x, t) =1√2Ψ1(x, t) +

1√2Ψ2(x, t) (31.12)

exposing the phase factors we get

Φ(x, t) =1√2ψ1(x)e

− i~ E1t +

1√2ψ2(x)e

− i~ E2t (31.13)

Let’s now track the probability of finding the particle in the same superposition

state. Similar to before we calculate

P (x, t) =

¯ZΦ∗(x, 0)Φ(x, t)

¯2=

¯Z µ1√2ψ∗1(x) +

1√2ψ∗2(x)

¶µ1√2ψ1(x)e

− i~ E1t +

1√2ψ2(x)e

− i~ E2t

¶¯2=

¯¯12Z Ã

ψ∗1(x)ψ1(x)e− i~ E1t + ψ∗1(x)ψ2(x)e

− i~ E2t

+ψ∗2(x)ψ1(x)e− i~ E1t + ψ∗2(x)ψ2(x)e

− i~ E2t

!dx

¯¯2

. (31.14)

The “cross-terms” (those of the form ψ∗1(x)ψ2(x) and ψ∗2(x)ψ1(x)) are zero when

215

Page 228: 457454 Physical Chemistry Quantum Chemistry

integrated because the eigenfunctions are orthogonal. This leaves

P (x, t) =

¯ZΦ∗(x, 0)Φ(x, t)

¯2(31.15)

=

¯1

2

Z ³ψ∗1(x)ψ1(x)e

− i~ E1t + ψ∗2(x)ψ2(x)e

− i~ E2t

´dx

¯2=

¯1

2

µe−

i~ E1t

Zψ∗1(x)ψ1(x)dx+ e−

i~ E2t

Zψ∗2(x)ψ2(x)dx

¶¯2=

¯1

2

³e−

i~ E1t + e−

i~ E2t

´¯2=1

4

³e+

i~ E1t + e+

i~ E2t

´³e−

i~ E1t + e−

i~ E2t

´=

1

4

³1 + e+

i~ (E1−E2)t + e−

i~ (E1−E2)t + 1

´=1

2

µ1 + cos

(E1 −E2)

~t

¶.

The probability of find in the system in its original superposition states is not one

for all times t.

216

Page 229: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 1

Listed here are some of the key equations for Exam 1. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• The short cut for getting the normalization constant .

N =

sZspace

|ψunnorm(x, y, z)|2 dxdydz. (31.16)

• The normalized wavefunction:

ψnorm =1

Nψunnorm. (31.17)

• How to get the average value for some property,

hαi =Zspace

ψ∗αψdxdydz. (31.18)

217

217

Page 230: 457454 Physical Chemistry Quantum Chemistry

• The Laplacian∇2 =

µ∂2

∂x2+

∂2

∂y2+

∂2

∂z2

¶. (31.19)

• Normalized wavefunctions for the 3D particle in a box,

ψn(x) =2√2√

abcsin

nxπx

asin

nyπy

bsin

nzπz

c. (31.20)

• The energy levels for the 3D particle in a box,

Enx,ny,nz =n2xh

2

8ma2+

n2yh2

8mb2+

n2zh2

8mc2. (31.21)

• Orthonormality: Zspace

ψ∗jψk =

(1, j = k

0, j 6= k. (31.22)

• Superpostion:ψ =

Xi

aiϕi (31.23)

• Commonly used comutators of the angular momentum operators arehLx, Ly

i= i~Lz,

hLy, Lz

i= i~Lx,

hLz, Lx

i= i~Ly (31.24)

and hL2, Lx

i=hL2, Ly

i=hL2, Lz

i= 0. (31.25)

• The energy levels for a particle in a ring are

Em =~2m2

2I=

m2h2

8π2I. (31.26)

• The normalized wavefunctions for a particle on a ring are

ψ =1√2π

eimφ. (31.27)

218

Page 231: 457454 Physical Chemistry Quantum Chemistry

• The eigenfunctions of angular momentum are entirely specified by two quan-tum numbers l and m: ψlm.

L2ψlm = l(l + 1)ψlm Lzψlm = mψlm (31.28)

• The energy levels for the rigid rotor are

El =l(l + 1)~2

2I. (31.29)

• Degeneracy for general angular momentum is

gJ = 2J + 1. (31.30)

• The first order energy correction in pertubation theory is

E(1)n =

Zallspace

ψ(0)∗n H(1)ψ(0)n dx, (31.31)

• The trial energy in variation theory is calculated by

Etrial =

Rallspace

ψ∗trialHψtrialdxRallspace

ψ∗trialψtrialdx(31.32)

• In generalΨn(x, t) ≡ ψn(x)e

− i~ Ent (31.33)

• The left and right superposition states are

ψL =1√2ψ1 +

1√2ψ2 (31.34)

and

ψR =1√2ψ1 −

1√2ψ2 (31.35)

219

Page 232: 457454 Physical Chemistry Quantum Chemistry

Part VI

Symmetry and Spectroscopy

220

220

Page 233: 457454 Physical Chemistry Quantum Chemistry

32. Symmetry and Group Theory

We now take a short break from physical chemistry to discuss ideas from the

mathematical field of group theory.

Inherent to group theory is symmetry.

As far as we are concerned, we will

• determine the symmetry of a particular molecule.

• The types of symmetry it has will determine to which symmetry group itbelongs.

• The mathematical properties of all the possible groups have been workedout

• These mathematical properties translate into a wide variety of variety ofphysical properties including

— Bonding

— Properties of wavefunctions

— Vibrational modes

— Many more applications

221

221

Page 234: 457454 Physical Chemistry Quantum Chemistry

32.1. Symmetry Operators

Any operator that leaves |ψ|2 invariant are symmetry operators for that particularsystem:

O |ψ|2 = |ψ|2 . (32.1)

This implies

Oψ = ±ψ. (32.2)

That is, the eigenvalues for the particular symmetry operator are 1 or −1.

For molecules we will be dealing with point group symmetry operators. These

operators deal with symmetry about the center of mass.

We have seen two such operators in ı and σh.

An example of symmetry operator that is not a point group symmetry operator

would be an operator that performed some sort of translation in space. This type

of operator arrises in the treatment of extended crystal structures.

∗ ∗ ∗ See Handout on Symmetry Elements ∗ ∗∗

32.2. Mathematical Groups

In mathematics the term “group” has special meaning. It is a set of objects and

a single operation, which has the following properties.

1. The group is associative (but not necessarily communative) with respect to

the operation.

2. An identity element exits and is a member of the group

222

Page 235: 457454 Physical Chemistry Quantum Chemistry

3. The “product” of any two members of the group yield a member of the

group.

4. The inverse of every member of the group is also in the group. In other

words, for any member of the group one can find another member of the

group which, upon “multiplication,” yields the identity element.

∗ ∗ ∗ See Handout on Naming Point Groups ∗ ∗∗

∗ ∗ ∗ See Handout on Assigning Point Groups ∗ ∗∗

Associated with a given group is a “multiplication” table.

32.2.1. Example: The C2v Group

The C2v group consists of the symmetry elements E, C2, σv (in-plane) and σ0v

(transverse).

Water is an example of a molecule described by this point group.

The multiplication table for the C2v group is

C2v E C2 σv σ0v

E E C2 σv σ0v

C2 C2 E σ0v σv

σv σv σ0v E C2

σ0v σ0v σv C2 E

32.3. Symmetry of Functions

In the absence of degeneracy, the wavefunctions must be symmetric or antisym-

metric with respect to all elements of the group.

223

Page 236: 457454 Physical Chemistry Quantum Chemistry

Connecting with the C2v group example lets consider the wavefunctions for water.

In this case one can collect the eigenvalues (either +1 or −1) for each of the foursymmetry operators as a four component vector. As it turns out there is four

possible sets of eigenvalues–hence four different vectors:

A1 = (1, 1, 1, 1)

A2 = (1, 1,−1,−1)B1 = (1,−1, 1,−1)B2 = (1,−1,−1, 1).

To see where these four vectors come from, consider the following.

• The first value has to be +1 since the only eigenvalue of E is 1

• The eigenvalue of C2 can be +1 or −1

— When it is +1 the vectors are labelled A

— When it is −1 the vectors are labelled B

• The eigenvalue of σv can be either +1 or −1

— When it is +1 the vectors are labelled with a subscript 1

— When it is −1 the vectors are labelled with a subscript 2

• The eigenvalue of σ0v can be either +1 or −1

• Finally there is a restriction do to the fact that the eigenvalues must obeythe group multiplication table.

— This restriction forces the eigenvalues of σv and σ0v to be the same for

the A type vectors and opposite for the B type vectors.

224

Page 237: 457454 Physical Chemistry Quantum Chemistry

The above considerations leave four vectors.

In fact, there will always be the same number of vectors as symmetry elements.

Altogether, the vectors represent what is call an irreducible representation of the

group.

These vectors make up the :

C2v E C2 σv σ0v

A1 1 1 1 1

A2 1 1 −1 −1B1 1 −1 1 −1B2 1 −1 −1 1

∗ ∗ ∗ See Handout on Character Tables ∗ ∗∗

32.3.1. Direct Products

The direct product of a two vectors is defined as

(x1, x2, x3, . . .)⊗ (y1, y2, y3, . . .) = (x1y1, x2y2, x3y3, . . .) (32.3)

For the example of the C2v group consider

B1 ⊗B2 = (1,−1, 1,−1)⊗ (1,−1,−1, 1)= (1, 1,−1,−1) = A2 (32.4)

32.4. Symmetry Breaking and Crystal Field Splitting

We shall investigate how degeneracies of energy levels are broken as one reduces

the overall symmetry of the system.

225

Page 238: 457454 Physical Chemistry Quantum Chemistry

In doing this we will, for simplicity, consider only proper rotations (Cn). Mirror

symmetry will not be considered (although in real applications one must consider

all symmetry).

First consider a free atom. In this case there is complete rotational symmetry.

Thus the symmetry group is the spherical group (see character table handout.)

This is the group associated with the particle on a sphere model and the angular

part of the hydrogen atom. The vectors are the labeled according to the angular

momentum quantum numbers S, P, D, F, etc.

The degeneracies of these vectors are 1 for S, 3 for P, 5 for D and so on as is

familiar to us already.

Now consider the free atom being placed in a crystal lattice of octahedral sym-

metry. For example placed at the center of a cube which has other atoms at the

centers of each face of the cube.

When moving to octahedral symmetry we now must look at the character table for

such a case–the O group (remember we are considering only proper rotations).

The S vector has the symmetry of a sphere (x2 + y2 + z2) and hence is totally

symmetric. It is also nondegenerate so it will be, of course, nondegenerate in

the octahedral case. It remains totally symmetric so it is now represented by the

vector A1.

The P vector is triply degenerate and has the symmetry of x, y and z as we see

from the character table for the spherical group. In the octahedral crystal the

degeneracy remains in tact and these states are represented by the T1 group.

226

Page 239: 457454 Physical Chemistry Quantum Chemistry

The D vector has a degeneracy of five and the symmetry of 2z2− x2− y2, xz, yz,

xy, x2 − y2. Looking at the table for the O group we see the degeneracy splits:

two states become E type and the remaining three become T2 type.

The F states have a degeneracy of 7 and the symmetry of z3, xz2, yz2, xyz,

z(x2 − y2), x(x2 − 3y2) and y(3x2 − y2). In an octahedral environment the states

split with one becoming A2, three becoming T1 and three becoming T2. This is

not readily apparent from the character tables so one needs to inspect a little

harder to see it (see homework).

The octahedral group is still highly symmetric. Lets say that two atoms on oppo-

site sides of the cube are moved slightly inward. The remaining four atoms remain

in place.

This breaks the octahedral symmetry and the system now assumes D4 symmetry.

Now the A1 vector of the O group becomes the A1 vector of the D4 group. The

triply degenerate T1 vector splits into a A2 state and a doubly degenerate E state.

The E states from the O group become a A1 type state and a B1 type state.

The T1 states from the O group become a A2 type state and a E type state.

The T2 states from the O group become a B2 and a E type state.

227

Page 240: 457454 Physical Chemistry Quantum Chemistry

33. Molecules and Symmetry

From our chapter on diatomic molecules last semester we have learned a great

deal which caries over directly to polyatomic molecules.

So, in this chapter we simply investigate some of the specific details regarding

polyatomic molecules.

33.1. Molecular Vibrations

As for diatomic molecules, it is convenient to work with center of mass coordinates.

With polyatomic molecules one needs to specify the coordinates of N nuclei rather

than just two nuclei.

To do so we begin with the 3N nuclear degrees of freedom.

As for the diatomic case 3 degrees of freedom determine the center of mass motion.

That leaves us with 3N − 3 coordinates to specify.One must now consider two different types of polyatomic molecules: Linear and

Nonlinear.

• For linear molecules there are 2 rotational degrees of freedom

• For nonlinear molecules there are 3 rotational degrees of freedom

This now leaves one with 3N − 5 vibrational degrees of freedom for linear poly-

atomic molecules and 3N − 6 vibrational degrees of freedom for nonlinear mole-

cules.

228

228

Page 241: 457454 Physical Chemistry Quantum Chemistry

33.1.1. Normal Modes

Polyatomic molecules can undergo very complicated vibrational motion.

Regardless of what type of vibrational motion is taking place, however, that mo-

tion is some linear combination of fundamental vibrational motions called normal

modes.

This is analogous to writing an arbitrary wavefunction as a linear combination of

eigenfunctions. One example was the “left” and “right” states of the two level

system.

The number of normal modes equals the number of vibrational degrees of freedom.

At low energies the normal modes are well approximated as harmonic oscillators.

33.1.2. Normal Modes and Group Theory

The symmetry of the normal modes are associated with entries in the character

table of the point group of any particular polyatomic molecule.

Example: Water

The point group symmetry of the water molecule is C2v. The character table is

C2v E C2 σv σ0v

A1 1 1 1 1

A2 1 1 −1 −1B1 1 −1 1 −1B2 1 −1 −1 1

Water has three nuclei and it is nonlinear so it has 3(3)− 6 = 3 normal modes.The three modes are the bending vibration, the symmetric stretching vibration

and the asymmetric stretch.

229

Page 242: 457454 Physical Chemistry Quantum Chemistry

The normal modes are associated with a particular vector (row) of the character

table by considering the action of the each of the symmetry elements on the normal

mode.

For the bending mode, the vibration is complete unchanged by any of the sym-

metry elements. Consequently the bending mode is associated with A1

The same is true for the symmetric stretching mode. It too is associated with A1.

The asymmetric stretch, however, is associated with B1 since C2 and σ0v transform

the mode into its opposite and σv leaves it unchanged.

230

Page 243: 457454 Physical Chemistry Quantum Chemistry

34. Vibrational Spectroscopy and

Group Theory

We now investigate how group theory and, in particular, the character tables can

be used to determine IR and Raman spectra and selection rules for polyatomic

molecules

34.1. IR Spectroscopy

IR absorption is exactly the same as regular electronic absorption except the

frequency of the electromagnetic radiation is much less.

The typical “energies” for IR absorption are from 400 to 4000 cm−1. This is in

the Infrared region of the electromagnetic spectrum.

As for electronic absorption one typically employs the electric dipole approxima-

tion.

The electric dipole approximation

• Molecule is viewed as a collection of charges

• Multipole expansion

monopole+ dipole+ quadrapole+ · · · (34.1)

231

231

Page 244: 457454 Physical Chemistry Quantum Chemistry

• Light—matter interaction is dominated by the light—dipole coupling so theother interactions are ignored.

In order for absorption of the electromagnetic radiation to take place, it must be

able to couple to a changing (oscillating) electric dipole.

The electric dipole is

μ = μxex + μyey + μzez (34.2)

where μx = qx, μy = qy, μz = qz.

The upshot of all this is as far as group theory is concerned is the following

selection rule:

• The vibrational coordinates for an IR active transition must have the samesymmetry as either x, y, or z for the particular group.

Example: Water

Recall that the point group symmetry of the water molecule is C2v.

We now need a column of the character table which we have ignored up to this

point.

The character table is

C2v E C2 σv σ0v Functions

A1 1 1 1 1 z, x2, y2, z2

A2 1 1 −1 −1 xy

B1 1 −1 1 −1 x, xz

B2 1 −1 −1 1 y, yz

The last column describes the symmetry of several important functions for the

point group.

232

Page 245: 457454 Physical Chemistry Quantum Chemistry

Among these functions are x, y, and z.

So we can see immediately that the IR active modes of any molecule having this

point group will be A1, B1, and B2.

The A2 mode is IR forbidden and any vibrations having this symmetry will not

appear in the IR spectrum (or it may appear as a very weak line).

From before we know the modes of water have A1, and B1 symmetry and hence

are all IR active and appear in the IR spectrum

34.2. Raman Spectroscopy

Raman spectroscopy is somewhat different than IR spectroscopy in that vibra-

tional frequencies are measured by way of inelastic scattering of high frequency

(usually visible) light.

The light loses energy to the material in an amount equal to the vibrational energy

of the molecules is the sample.

This lose of energy shows up in the scattered light as a new down shifted frequency

from that of the original input light frequency.

Unlike IR absorption which is based on the electric dipole, Raman scattering is

based on the polarizability of the molecule

Roughly speaking the polarizability of a molecule determines how the electron

density is distorted through interaction with an electromagnetic field.

233

Page 246: 457454 Physical Chemistry Quantum Chemistry

The molecular quantity of interest is the polarizability tensor,↔α.

We will not get into tensors in this course except to say the polarizability tensor

elements are proportional to the quadratic functions, x2, y2, z2, xy, xz, yz, (or

any combinations thereof).

One can now inspect the character table to determine which modes will be Raman

active.

For the example of water, all modes are Raman active

Rule of Mutual exclusion

• Vibrational mode can be both IR and Raman active or inactive

• If, however, the molecule has inversion symmetry (contains ı as a symmetryelement) then no modes will be both IR and Raman active.

234

Page 247: 457454 Physical Chemistry Quantum Chemistry

35. Molecular Rotations

Recall that the three degrees of freedom that described the position of the nuclei

about the center of mass were (R, θ, φ). The R was involved in vibrations. We

now turn our attention to the angular components to describe rotations.

Recall also the Kinetic energy operator for the nuclei in the center of mass coor-

dinates

TN = −~2

2μ∇2N = −

~2

2μR2∂

∂RR2

∂R+~2

2μJ2. (35.1)

We will now be concerned only with the angular part,

−~2

2IJ2. (35.2)

Now, under the Born-Oppenheimer approximation, R is a parameter. For constant

R the rotational energy is given by

Erot =J(J + 1)~2

2μR2=

J(J + 1)h2

8π2I. (35.3)

This is the so-called rigid rotor energy.

It is common to define

Be ≡h

8π2I(35.4)

as the rotational constant. Then

Erot = J(J + 1)hBe (35.5)

with a degeneracy of

gJ = 2J + 1 (35.6)

235

235

Page 248: 457454 Physical Chemistry Quantum Chemistry

35.1. Relaxing the rigid rotor

Of course the rigid rotor is not a perfectly correct model for a diatomic molecule.

There are two corrections we will now make

1. Vibrational state dependence:

• The R value is dependent on the particular vibrational level.

• One defines a rotational interaction constant that depends on the vi-brational level, n.

Bn ≡ Be −µn+

1

2

¶αe, (35.7)

where αe is an empirical rotational—vibrational interaction constant.

2. Centrifugal stretching:

• Rotation tends to stretch the diatomic distance R.

• This is corrected for by the term

−J2(J + 1)2Dc, (35.8)

where

Dc ≡4B3

e

ω2e(35.9)

is the centrifugal stretching constant.

35.2. Rotational Spectroscopy

A rotational transition can occur in the same vibrational level n. This is called a

pure rotational transition. Alternatively, a rotational transition can accompany a

vibrational transition.

In either case the selection rule for the transition is 4J = ±1.

236

Page 249: 457454 Physical Chemistry Quantum Chemistry

It turns out that typical rotational energy gaps are on the order of a few wavenum-

bers or less.

Thermal energy, kT, at room temperature is about 200 cm−1. This means that

at room temperature the many excited rotational states are populated.

∗ ∗ ∗ See Handout ∗ ∗∗

The selection rules and the thermalized states combine to yield a multi-peaked

ro-vibrational spectrum.

∗ ∗ ∗ See Handout ∗ ∗∗

35.3. Rotation of Polyatomic Molecules

There are a few additional details regarding rotations for polyatomic molecules as

compared to diatomics

Of course one could set-up an arbitrary center of mass coordinate system. But

one system is special–the principle axes coordinate system.

The principle axes coordinate system is the one in which the z-axis is taken to be

along the principle symmetry axis.

The total moment of inertia, I = Ixx + Iyy + Izz

The Hamiltonian in the principle axes system is

H =~2

2

"J2xIxx

+J2yIyy

+J2zIzz

#(35.10)

237

Page 250: 457454 Physical Chemistry Quantum Chemistry

There are four classes of polyatomic molecules regarding rotations

1. Linear (e.g., carbon dioxide)

• Izz = 0, Ixx = Iyy

• J2 = J2x + J2y

• The Hamiltonian isH =

~2

2IxxJ2 (35.11)

• The rotational energy is

Erot = hBJ(J + 1), (35.12)

where

B =h

8π2Ixx(35.13)

2. Symmetric tops (e.g., benzene)

• Ixx = Iyy

• J2 = J2x + J2y + J2z

• The Hamiltonian is

H =~2

2

"J2x + J2yIxx

+J2zIzz

#(35.14)

• The rotational energy is

Erot = hBJ(J + 1) + h(A−B)K2, (35.15)

where

A =h

8π2Izz, (35.16)

B =h

8π2Ixx(35.17)

and K is the quantum number describing the projection of the angular

momentum onto the z-axis

238

Page 251: 457454 Physical Chemistry Quantum Chemistry

3. Spherical tops (e.g., methane)

• Ixx = Iyy = Izz

• J2 = J2x + J2y + J2z

• The Hamiltonian isH =

~2

2IxxJ2 (35.18)

• The rotational energy is

Erot = hBJ(J + 1), (35.19)

where

B =h

8π2Ixx(35.20)

4. Asymmetric tops

• Ixx 6= Iyy 6= Izz

• These are more complicated and we will not discuss them in detail

239

Page 252: 457454 Physical Chemistry Quantum Chemistry

36. Electronic Spectroscopy of

Molecules

The electronic spectra of molecules are quite different than that of atoms.

Atomic spectra consist of single sharp lines due to transitions between energy

levels.

Molecular spectra, on the other hand, have numerous lines (bands) due to the

fact that electronic transitions are accompanied by vibrational and rotational

transitions.

36.1. The Structure of the Electronic State

Last semester we saw that under the Born—Oppenheimer approximation we were

able to write the molecular wavefunction as a product of an electronic part and a

nuclear part.

We found that in doing so the electronic energy level, Ee, was parameterized by

the internuclear distance, R.

Ee as a function of R describe the effective potential for the nuclei.

It had a qualitative shape similar to the Morse potential.

240

240

Page 253: 457454 Physical Chemistry Quantum Chemistry

In the figure below the ground and first excited electronic levels (as a function of

R) are shown.

Note: The potential minima are not at the same value of R for each of the

electronic states.

36.1.1. Absorption Spectra

In absorption spectroscopy, light promotes an electron from the ground electronic

state (and usually from the ground vibrational state too) to the excited electronic

state and any of the excited vibrational states of the excited electronic state.

∗ ∗ ∗ See Spectroscopy Supplement p1 ∗ ∗∗

36.1.2. Emission Spectra

In emission spectroscopy, light demotes an electron from the ground vibrational

state of the excited electronic state to any one of a number of excited vibrational

levels in the ground electronic state.

241

Page 254: 457454 Physical Chemistry Quantum Chemistry

∗ ∗ ∗ See Spectroscopy Supplement p2 ∗ ∗∗

36.1.3. Fluorescence Spectra

All during the process of absorption, the process of is taking place.

∗ ∗ ∗ See Spectroscopy Supplement p3 ∗ ∗∗

As seen in the supplement the fluorescence spectrum is shifted to lower energies

(red shifted) from the absorption spectrum.

This is known as the Stokes shift.

The main stream explanation for the stokes shift is as follows

• Light promotes the system from the ground vibrational and ground elec-

tronic state to excited vibrational levels in the excited electronic state.

• The system then very rapidly (on the order of tens to hundreds of fem-

toseconds) relaxes to the ground vibrational state of the excited electronic

state.

• This process is called

• The molecule than emits a photon to drop back down into an excited vibra-tional state of the ground electronic state.

• This requires a lower energy (or “more red”) photon. Hence the Stokes shift.

242

Page 255: 457454 Physical Chemistry Quantum Chemistry

36.2. Franck—Condon activity

We have seen than an electronic tranistion involves not only a change in the

electronic state but also in the vibrational state in general (and in the rotaitonal

state as well, but we will ingore this).

Assuming the electronic transition is allowed one must calculate the probability of

the vibrational transistion as well. This is down by evaulating the Franck—Condon

integral.

36.2.1. The Franck—Condon principle

When the Born—Oppenheimer approximation is applied to spectroscopic transi-

tions, one obtains the Franck—Condon principle.

The Franck—Condon principle states that the nuclei do not move during an elec-

tronic transition.

Physically this means that for a particular transition to be Franck—Condon ac-

tive there must be good overlap of the vibrational wavefunctions involved in the

transition.

Mathematically this means that the strength of a transition fromΨi = ψel,iψvib,i →Ψf = ψel,fψvib,f is given by¯

¯Z

allspace

Ψ∗f μelΨi

¯¯2

=

¯¯Z

elspace

Zvibspace

ψ∗el,fψ∗vib,f μelψel,iψvib,i

¯¯2

, (36.1)

243

Page 256: 457454 Physical Chemistry Quantum Chemistry

where μel is the electronic transition dipole. We can separate the integrals as¯¯Z

elspace

ψ∗el,f μelψel,i

¯¯2

| z if 6=0, allowed

¯¯Z

vibspace

ψ∗vib,fψvib,i

¯¯2

| z Franck—Condon

, (36.2)

244

Page 257: 457454 Physical Chemistry Quantum Chemistry

37. Fourier Transforms

As a spectroscopist it is imperative to have a deep understanding of the relation-

ship between time and frequency.

Spectroscopic data is obtained either in the time domain or in the frequency

domain and one should readily be able to look at data in one domain and know

what is happening in the other domain.

One should be familiar with qualitative aspects of this time—frequency relation,

such as if a signal oscillates in time it will have a peak in it frequency spectrum

at the frequency with which it is oscillating.

Furthermore, if the signal decays rapidly it will have a broad spectrum and, con-

versely, if the signal decays slowly it will have a narrow spectrum. The mathemat-

ics which governs these qualitative statements is Fourier transform theory which

we now review.

37.1. The Fourier transformation

The Fourier transformation, =, of a function f(t) will, in this work, by denoted

by a tilde, f(ω), and is given by

= [f(t)] = f(ω) =

Z ∞

−∞f(t)eiωtdt. (37.1)

245

245

Page 258: 457454 Physical Chemistry Quantum Chemistry

The Fourier transformation is unique and it has a unique inverse, =−1, which isgiven by

=−1hf(ω)

i= f(t) =

1

Z ∞

−∞f(ω)e−iωtdω. (37.2)

The above two relations form the convention used throughout this work.

Other authors use different conventions, so one must take care to know exactly

which convention is being used.

For simplicity the symbol = will be used to represent the Fourier transformationoperation, i.e., = [f(t)] = f(ω).Whereas the symbol =−1 will represent the inverseFourier transformation, i.e., =−1

hf(ω)

i= f(t).

246

Page 259: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 2

Listed here are some of the key equations for Exam 2. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• Vibrational degrees of freedom

— linear: 3N − 5

— not linear: 3N − 6

• The so-called rigid rotor energy is

Erot = J(J + 1)hBe. (37.3)

where

Be ≡h

8π2I(37.4)

is the rotational constant.

247

247

Page 260: 457454 Physical Chemistry Quantum Chemistry

• The degeneracy of the rigid rotor is

gJ = 2J + 1 (37.5)

• Franck—Condon Factor: ¯¯Z

vibspace

ψ∗vib,fψvib,i

¯¯2

(37.6)

• The Fourier transformation is

= [f(t)] = f(ω) =

Z ∞

−∞f(t)eiωtdt. (37.7)

• The inverse Fourier transformation is

=−1hf(ω)

i= f(t) =

1

Z ∞

−∞f(ω)e−iωtdω. (37.8)

248

Page 261: 457454 Physical Chemistry Quantum Chemistry

Part VII

Kinetics and Gases

249

249

Page 262: 457454 Physical Chemistry Quantum Chemistry

38. Physical Kinetics

We now turn our attention to the molecular level and in particular to molecular

motion.

38.1. kinetic theory of gases

A microscopic view of gases

Consider a gas of point mass (m), m is the molecular (or atomic) mass

• Each particle of mass m has velocity v, hence a momentum of p = mv and

a kinetic energy of KE = 12mv · v = 1

2mv2.

• A sample of N molecules is characterized by its number density n∗ = NV.

• From the ideal gas law PV = nRT = NLRT (L is Avogadro’s number):

NV= PL

RT= n∗

Consider the ith particle at position xi = (x, y, z) in coordinate (position) space.

Its velocity is vi = dxidt=¡dxidt, dyidt, dzidt

¢. This can represented in velocity space by

the vector vi = (vxi , vyi , vzi).

The velocities of the particles are characterized by a probability distribution func-

tion for velocities F (vx, vy, vz, t) which is in general a function of time, t.

250

250

Page 263: 457454 Physical Chemistry Quantum Chemistry

The number of particles, NVv , having velocities in a macroscopic volume, Vv, in

velocity space is

NVv = N

ZVv

F (v, t)dv = N

Z Z ZVv

F (vx, vy, vz, t)dvxdvydvz (38.1)

It is more convenient to switch to spherical polar coordinates in velocity space

(v, θ, φ); n.b., v is simply a magnitude (not a vector)–it is the speed.

The probability distribution function then becomes F (v, θ, φ, t)

If we choose the origin of our coordinate system to be at the center of mass of the

gas, then for many cases the velocity distribution will be isotropic–independent

of θ and φ.

F (v, θ, φ, t) = F (v, t). (38.2)

Furthermore, stationary distributions–those independent of time–are often en-

countered.

F (v, θ, φ, t) = F (v, θ, φ). (38.3)

251

Page 264: 457454 Physical Chemistry Quantum Chemistry

We shall consider stationary isotropic distributions F (v). So F (v) represents a

distribution of speeds.

It can be shown from first principles that

F (v) = 4π

µm

2πkbT

¶32

e−mv2

2kbT v2 (38.4)

where kb = 1.380658 × 10−23 is Boltzmann’s constant. This is the Maxwell’sdistribution (of speeds).

38.2. Molecular Collisions

The average speed of a particle can calculated from Maxwell’s distribution:

hvi = v =

Z ∞

0

vF (v)dv =

Z ∞

0

v34π(m

2πkT)32 e−

mv2

2kT dv (38.5)

=

s8kT

πm

µL

L

¶Lk=R=

Lm=M

r8RT

πM

It will be convenient to define number density as n∗ ≡ NVwhere N is the number

of particles. For an ideal gas (V = nRTP), n∗ =

N

n|z=L

PRT= LP

RT.

A simple model for molecular collisions:

252

Page 265: 457454 Physical Chemistry Quantum Chemistry

• Particles are hard spheres of radius σ.

• A Particle moving at v sweeps out a cylinder of radius σ and length v4t =⇒V = πσ2v4t.

∗ ∗ ∗ See Handout ∗ ∗∗

• The number of collisions equals the number of particles with their centersin V :

number of collisions = n∗πσ2v4t (38.6)

• The collision frequency = n∗πσ2v

For the above model we need to find the average collision frequency. Since the

molecules are moving relative to one another we must find the average relative

velocity, v12 = h|v1 − v2|i

It can be shown that

v12 =

r16RT

πM=√2v =⇒ collision

frequency=√2n∗πσ2v. (38.7)

From the above expression one defines the mean free path λ to be

λ =v/√

2n∗πσ2v/n∗=LP

RT=RT√2PLπσ2

(38.8)

Example: Ar at SATP (T = 298 K, P = 1 bar):

v = 380.48ms,

collisionfrequency

= 5.25× 109 s−1,

λ = 72.5 nm

253

Page 266: 457454 Physical Chemistry Quantum Chemistry

39. The Rate Laws of Chemical

Kinetics

Thermodynamics described chemical systems in equilibrium. For the study of

chemical reactions it is important understand systems that can be very far from

equilibrium. For this we turn to the field of chemical kinetics.

We can, from thermodynamics, address the question; Will the reaction occur?

We need kinetics, however, answer the question: How fast will the reaction occur?

39.1. Rate Laws

Consider a general four component reaction

aA + bB = cC + dD (39.1)

The time dependence of this reaction can be observed by following the disap-

pearance of either of the reactants or appearance of either of the products. That

is,

−d[A]dt

or − d[B]dtor

d[C]dtor

d[D]dt

(39.2)

BUT this is ambiguous because a moles of A reacts with b moles of B and a does

not, in general, equal b. We must account for the stoichiometry.

254

254

Page 267: 457454 Physical Chemistry Quantum Chemistry

We define the reaction velocity as

v =1

vi

d[I]dt

(39.3)

where vi = −a,−b, c or d and I = A, B, C, or D.

This definition is useful but must be used with caution since for complicated

reactions all the v’s may not be equal. An example of this is

aA +½

bB → cC + dDb0B0 → c0C0 + d0D0

(39.4)

A rate law is the mathematical statement of how the reaction velocity depends

on concentration.

v = f(conc.) (39.5)

For the most part, rate laws are empirical.

Many, but certainly not all, rate laws are of the form

v = k[A1]xA1 [A2]xA2 · · · [An]xAn . (39.6)

The reaction is said to be of order xAi in species Ai and it is of overall orderPi xAi .

In general an overall reaction is made up of so called elementary reactions

Reactant = Product overall rxn (39.7)

Reactant → Intermediates → Product

Note that we shall use an equal sign when talking about the overall reaction and

arrows when talking about the elementary reactions

Example

255

Page 268: 457454 Physical Chemistry Quantum Chemistry

Let

2A + 2B = C + D (39.8)

be the overall reaction. One possible set of elementary steps could be

elementary rxn molecularity

A + A → A0 Bimolecular

A0 → A00 Unimolecular

A00 + 2B→ C + D Trimolecular

.

The rate laws for elementary reactions can be determined from the stoichiometry

molecularity elementary rxn rate law

Unimolecular A→ Product v = k[A]

Bimolecular A + A → Product v = k[A]2

Bimolecular A + B → Product v = k[A][B]

Trimolecular A + A + A → Product v = k[A]3

Trimolecular A + A + B → Product v = k[A]2[B]

Trimolecular A + B + C → Product v = k[A][B][C]

.

Conversely, rate laws for overall reactions can not be determined by stoichiometry.

Connection to thermodynamicsConsider the overall or elementary reaction

aA + bBkf­kr

cC + dD (39.9)

where kf is the rate constant for the reaction to proceed in the forward direction

and kr is the rate constant for the reaction to proceed in the reverse direction.

Now, at equilibrium vf = vb which implies

kf [A]a[B]b = kr[C]c[D]d (39.10)

256

Page 269: 457454 Physical Chemistry Quantum Chemistry

bringing kr to the LHS and [A][B] to the RHS we get

kfkr=[C]c[D]d

[A]a[B]b= K 0

c (39.11)

where K 0c is the thermodynamic equilibrium “constant.”

So, we have succeeded in connecting thermodynamics to kinetics BUT we have

done so through the ratio of rate constants. The velocity of a reaction is lost

in this ratio and hence we still can not determine the speed of a reaction from

thermodynamics.

Examples of rate lawsConsider the (overall) reaction between molecular hydrogen and molecular iodine,

H2 + I2 = 2HI. (39.12)

The observed rate laws are vf = kf [H2][I2] and vr = kr[HI]2. This suggests that the

reaction is elementary. In fact, the reaction is not elementary. Moral: Kineticsis very much an empirical science.

Next consider the reaction between molecular hydrogen and molecular bromine,

H2 + Br2 = 2HBr. (39.13)

The observed rate law for this reaction is very complicated,

v =k[H2][Br2]1/2

1 + k0[HBr][Br2]

,

this does not obey any common form.

The above two example are seemingly very similar but they have very different

observed rate laws. Moral: Kinetics is very much an empirical science.

Objectives of chemical kinetics

257

Page 270: 457454 Physical Chemistry Quantum Chemistry

• To establish empirical rate laws

• To determine mechanisms of overall reactions

• To empirically study elementary reactions

• To establish theoretical links to statistical mechanics and quantum mechan-ics

— This involve nonequilibrium thermodynamics–more difficult

• To study chemical reaction dynamics

— the dynamics of molecular collisions that result in reactions

39.2. Determination of Rate Laws

Concentrations c(t) are measured not rates. To obtain the rate from the concen-

tration we must take its time derivative dc(t)dt

. That is we must measure c(t) as a

function of time and find the rate of change of this concentration curve.

The rates of chemical reactions vary enormously from sub-seconds to years. Con-

sequently no one experimental technique can be used.

• For slow reactions (hrs/days) almost any technique for measuring the con-centration can be used.

• For medium reactions (min) either a continuous monitoring technique or a

stopping technique can be used

— A stopping technique used rapid cooling or destruction of the catalyststo stop a reaction at a given point.

• Very fast (sec/subsec) reactions cause problems because the reaction goesfaster than one can mix the reactants.

258

Page 271: 457454 Physical Chemistry Quantum Chemistry

39.2.1. Differential methods based on the rate law

Methods based directly on the rate law rely on the determination of the time

derivative of the concentration.

The main problem with such a method is that randomness in the concentration

measurements gets amplified when taking the derivative.

1. Method of initial velocities

• for v = k[A]x[B]y rate laws.

• initially v0 = kaxby where a and b are the initial concentrations of A

and B respectively

• taking the log of both sides gives lnv0 = ln[kaxby] = ln k+x ln a+y ln b

• a and b can be varied independently so both x and y can be determined.

• problems

1. if the concentration drops very sharply

2. if there is an induction period

2. Method of isolation

• for v = k[A]x[B]y rate laws

• start with initial concentrations a and b equal to the stoichometry; thisgives the overall order of x+ y

• flood with, say, A so v ≈ kax[B]y

39.2.2. Integrated rate laws

The above differential methods look directly at the rate law which is a differential

equation. The differential equation is not solved.

259

Page 272: 457454 Physical Chemistry Quantum Chemistry

We now solve the differential equations to yield what are called the integrated

rate law.

The differential equations (rate law) and their solutions (integrated rate law) are

simply listed here for a few rate laws.type rate lawa) integrated rate lawa)

1st order 1vi

d[I]dt= k[I] [I] = [I0]e

vikt

2nd order 1vi

d[I]dt= k[I]2 1

[I]= 1

[I0]− vikt

nth orderb) 1vi

d[I]dt= k[I]n 1

[I]n−1 =1

[I0]n−1− (n− 1)vikt

enyzme 1vi

d[I]dt= k[I]

km+[I]km ln

[I0][I]+ ([I0]− [I]) = −vikt

a)[I] is the concentration of one of the reactants in an elementary reaction and

vi is the stoichiometric factor for [I] (n.b., vi is a negative number).b)The order need not be an integer. For example n = 3/2 is a three-halves

order rate law.

260

Page 273: 457454 Physical Chemistry Quantum Chemistry

40. Temperature and Chemical

Kinetics

40.1. Temperature Effects on Rate Constants

An empirical rate constant was proposed by Arrhenious:

d ln k

dT=

Ea

RT 2or (40.1)

d ln k

d(1/T )=

Ea

R, (40.2)

where Ea is the Arrhenious activation energy.

Integration of the above yields

ln k = lnA− Ea

RT=⇒ k = Ae−

EaRT (40.3)

(A is the constant of integration). This is the Arrhenious equation

Recall the equilibrium constant can also be obtained from kinetics

K 0c =

kfkr' Ka. (40.4)

Now, take the log of this:

lnKa = ln

∙kfkr

¸= ln kf − ln kr. (40.5)

261

261

Page 274: 457454 Physical Chemistry Quantum Chemistry

Substituting the Arrhenious equation for the rate constants gives

lnKa = ln

∙Afe

−EafRT

¸− ln

hAre

−EarRT

i(40.6)

= ln

∙Af

Ar

¸+

Ear −Eaf

RT

40.1.1. Temperature corrections to the Arrhenious parameters

The Arrhenious parameters A and Ea are constants.

Theoretical approaches to reaction rates predict rate constants of the form

k = aT je−E0/RT . (40.7)

Forcing this to coincide with the Arrhenious implies

Ea = E0 + jRT (40.8)

and

A = aT jej (40.9)

We can verify this by starting with the Arrhenious equation and substituting the

above expressions,

k = Ae−EaRT = aT jeje−

E0+jRTRT = aT jej/ e−j/ e

−E0RT = aT je

−E0RT√

(40.10)

40.2. Theory of Reaction Rates

Simple collision theory (SCT)

• Bimolecular reactions (A,B)

• Reaction rate determined by molecular collisions

262

Page 275: 457454 Physical Chemistry Quantum Chemistry

— Collision frequency for A–B collisions

zAB = πσABL2

s8RT

πLμ[A][B] (40.11)

where μ ≡ mAmBmA+mB

is the reduced mass and σAB is the collision diameter.

• The maximum reaction velocity is vmax = zABL, but intuitively the actual

reaction velocity will be less because

— the ability to react depends on orientation =⇒ a steric factor p

— a minimum amount of collisional energy is required=⇒ e−Emin/RT

• The actual reaction velocity is

v =pzABe

−Em inRT

L(40.12)

• The rate constant for a bimolecular reaction is

k =v

[A][B](40.13)

so SCT predicts

k =pzAB e

−Em inRT

L

[A][B]= pπσABL

s8RT

πLμe−

Em inRT (40.14)

263

Page 276: 457454 Physical Chemistry Quantum Chemistry

• Comparison to the (temperature corrected) Arrhenious equation suggests

A = pπσABL

s8RT

πLμe12 (40.15)

and

Ea = Emin +1

2RT (40.16)

Activated complex theory (ACT)

• An intermediate active complex is formed during the reaction, e.g.,

A + B → (AB)‡ → products. (40.17)

ACT is not limited to bimolecular reactions.

• The active complex is a state in the thermodynamic sense, thus we can applythermodynamics to it.

• For the above example, the equilibrium constant is defined as

K‡a =

a‡

aAaB

low'conc.

[‡][A][B]

(40.18)

264

Page 277: 457454 Physical Chemistry Quantum Chemistry

• Definition: transmission factor, f

— accounts for the fraction of activated complex that becomes product.

— From statistical mechanics, it can be shown that f = kbT/h where kbis Boltzmann’s constant and h is Planck’s constant.

• The reaction rate constant for reactants going to products for ACT is

k = fK‡a =

kbT

hK‡

a (40.19)

• Thermodynamics tells us that

4G‡ = −RT lnK‡a (40.20)

which can be written as

K‡a = e−

4G‡RT = e−

4H‡RT e

4S‡R (40.21)

where 4G‡ = 4H‡ − T4S‡.

• The ACT reaction rate constant now becomes

k =kbT

he−

4H‡RT e

4S‡R . (40.22)

This is Eyring’s equation

40.3. Multistep Reactions

Up to now, the reactions we have studied have been single step reactions.

In general, there is many steps from initial reactants to final products.

Reactions may occur in series or in parallel or both, in what is called a reaction

network.

Parallel reactions:

265

Page 278: 457454 Physical Chemistry Quantum Chemistry

• Parallel reactions are of the form, for example,

A + B1k1→ C (40.23)

A + B2k2→ D

• The rate constant for the disappearance of [A] is simply the sum of the tworate constants: k = k1 + k2

Series reactions:

• Series reactions necessarily include and intermediate product. They are ofthe form

A k1→ B k2→ C (40.24)

• The concentrations of A, B and C are determined by the system of differen-tial equations:

−d[A]dt

= k1[A]

d[B]dt

= k1[A]− k2[B]

d[C]dt

= k2[B],

which, when solved yields

[A] = [A0]e−k1t

[B] =k1[A0]k2 − k1

¡e−k1t − ek2t

¢[C] = [A0]− [A]− [B] = [A0]

µ1− k2e

−k1t − k1ek2t

k2 − k1

¶• See in class animation

266

Page 279: 457454 Physical Chemistry Quantum Chemistry

40.4. Chain Reactions

Chain reactions are reactions which have at least one step that is repeated indef-

initely. The simplest chain reactions have three distinct steps (discussed below)

Chain reactions are extremely important in polymer chemistry

Steps of a chain reaction

1. Initiation: Typically a molecule M reacts to form some highly reactive rad-

ical

M→ R·.

267

Page 280: 457454 Physical Chemistry Quantum Chemistry

2. Propagation: The radical formed in the initiation step reacts with some so

molecule M0 to form another molecule M00 and another radical R0·. This steprepeats an indefinite number of times.

R·+M0 →M00 +R0·.

3. Termination: The radicals interact with each other or with the walls of the

container to forma stable molecule

R0·+R0·→M000

or

R0·+wall→ removed

268

Page 281: 457454 Physical Chemistry Quantum Chemistry

41. Gases and the Virial Series

Unlike liquids and solids, a particular particle has much less significant interactions

with the other particles.

This simplifies the theoretical treatment of gases.

We will now look in detail at the gases.

41.1. Equations of State

Recall from last semester several of the equations of states for gases.

• The ideal gas equation of state

PV = nRT. (41.1)

The equation of state can also be expressed in term of density ρ = mV

ρ =mP

nRT. (41.2)

• The van der Waals gas equation of state

P =nRT

V − nb− n2a

V 2(41.3)

or

P =RT

Vm − b− a

V 2m

, (41.4)

where the parameter a accounts for the attractive forces among the particles

and parameter b accounts for the repulsive forces among the particles

269

269

Page 282: 457454 Physical Chemistry Quantum Chemistry

• BerthelotP =

nRT

V − nb− n2a

TV 2=

RT

Vm − b− a

TV 2m

(41.5)

• DietericiP =

nRTe−anRTV

V − nb=

RTe−a

RTVm

Vm − b(41.6)

• Redlich-Kwang

P =nRT

V − nb− n2a√

TV (V − nb)=

RT

Vm − b− a√

TVm (Vm − b)(41.7)

41.2. The Virial Series

Definition: Compressibility Factor: z = PVnRT

= PVmRT

.

• z is unity for an ideal gas because for such a gas PV = nRT.

• For a real gas z must approach unity upon dilution ( nV→ 0).

• z can be expended in a power series called the virial series.

The virial series in powers of nVis

z = 1 +B(T )³ nV

´+ C(T )

³ nV

´2+D(T )

³ nV

´3+ · · · , (41.8)

or

z = 1 +B(T )

µ1

Vm

¶+ C(T )

µ1

Vm

¶2+D(T )

µ1

Vm

¶3+ · · · . (41.9)

B(T ), C(T ), etc. are called the virial coefficients.

Conceptually B(T ) represents pair-wise interaction of the particles, C(T ) repre-

sents triplet interactions, etc.

270

Page 283: 457454 Physical Chemistry Quantum Chemistry

41.2.1. Relation to the van der Waals Equation of State

Recall the van der Waals equation

P =RT

Vm − b− a

V 2m

(41.10)

multiply both sides by VmRTto get

PVmRT

=Vm

R/ T/R/ T/Vm − b

−V/

m

RT

a

V2/m

=Vm

Vm − b− a

RTVm

=1

1− bVm

− a

RTVm(41.11)

but PVmRT

= z so

z =1

1− bVm

− a

RTVm. (41.12)

The first term is of the form 11−x which has the power series expansion

1

1− x= 1 + x+ x2 + · · · . (41.13)

Therefore

z = − a

RTVm+ 1 +

b

Vm+

µb

Vm

¶2+ · · · . (41.14)

the first term is proportional to 1Vmand so it can be combined with the 1

Vmterm

in the series expansion, hence

z = 1 +³b− a

RT

´ 1

Vm+

µb

Vm

¶2+ · · · . (41.15)

This series can now be compared term by term to the virial series to give expression

for the virial coefficients:

B(T ) =³b− a

RT

´, C(t) = b2, D(T ) = b3, etc. (41.16)

271

Page 284: 457454 Physical Chemistry Quantum Chemistry

41.2.2. The Boyle Temperature

The temperature at which B(T ) = 0 is called the Boyle temperature, Tb.

The virial series at Tb becomes

z(T = Tb) = 1 + 0

µ1

Vm

¶+ C(T )

µ1

Vm

¶2+D(T )

µ1

Vm

¶3+ · · ·

= 1 + C(T )

µ1

Vm

¶2+D(T )

µ1

Vm

¶3+ · · · . (41.17)

The lowest order correction are now³

1Vm

´2. The gas behaves more like an ideal

gas at Tb then for other temperatures.

41.2.3. The Virial Series in Pressure

One can also expand the compressibility factor in pressure

z = 1 +B0(T )P + C 0(T )P 2 +D0(T )P 3 + · · · . (41.18)

The relation of this expansion to the one in 1Vmcan be obtained. One finds (see

homework)

B0(T ) =B(T )

RT, (41.19)

272

Page 285: 457454 Physical Chemistry Quantum Chemistry

C 0(T ) =C(T )−B(T )2

(RT )2(41.20)

and

D0(T ) =D(T )− 3B(T )C(T )− 2B(T )3

(RT )3(41.21)

41.2.4. Estimation of Virial Coefficients

The virial coefficients can be estimated using empirical equations and tabulated

parameters.

• Estimates based on Beattie-Bridgeman constants:

B(T ) = B0 −A0RT− c

T 3, (41.22)

C(T ) =A0a

RT−B0b−

B0c

T 3, (41.23)

D(T ) =B0bc

T 3. (41.24)

where A0, B0, a, b, c are tabulated constants

• Estimates based on critical values (we will discuss critical values shortly, fornow treat them as empirical parameters):

B(T ) =9RTc128Pc

µ1− 6T

2c

T 2

¶. (41.25)

273

Page 286: 457454 Physical Chemistry Quantum Chemistry

42. Behavior of Gases

42.1. P, V and T behavior

We shall briefly consider the P, V and T behavior of dense fluids (e.g., liquids).

Taking volume as a function of P and T, we consider the total derivative

dV (T, P ) =

µ∂V

∂T

¶P

dT +

µ∂V

∂P

¶T

dP. (42.1)

We can change this from a extensive property equation to an intensive property

equation by dividing by V :

dV

V=1

V

µ∂V

∂T

¶P| z

α

dT +1

V

µ∂V

∂P

¶T| z

−κT

dP.

α is the coefficient of thermal expansion.

• At a given pressure, α describes the change in volume with temperature.

• Positive α means the volume of the fluid increases with increasing temper-ature.

κT is the isothermal compressibility

• At a given temperature, κT describes the change in volume with pressure.

• Positive κT means the volume of the fluid decreases with increasing pressure.

• κT is different from z, the compressibility factor.

274

274

Page 287: 457454 Physical Chemistry Quantum Chemistry

42.1.1. α and κT for an ideal gas

As an exercise we shall calculate α and κT using the ideal gas equation of state

(n.b., it is, of course, absurd to treat a liquid as an ideal gas). Starting with the

ideal gas law: V = nRTP

.

κT =−1V

µ∂V

∂P

¶T

=−1V

Ã∂¡nRTP

¢∂P

!T

=−1V

µ−nRT

P 2

¶(42.2)

=1

(PV )| z =nRT

nRT

P=

1

n/ R/ T/n/ R/ T/P

=1

P

and

α =1

V

µ∂V

∂T

¶P

=1

V

Ã∂¡nRTP

¢∂T

!P

=1

V P|z=n/R/T

n/ R/ =1

T(42.3)

42.1.2. α and κT for liquids and solids

In general, the compressibility and expansion of liquids (and solids) are very small.

So one can expand the volume in a Taylor series about a known pressure, P0.

At constant T

V (P ) = V0 +

µ∂V

∂P

¶T| z

−V0κT

(P − P0) +

µ∂V

∂P

¶2T

(P − P0)2 + · · · (42.4)

so,

V (P ) ≈ V0 [1− κT (P − P0)] . (42.5)

This approximation is quite good even over a rather large pressure range (P−P0 =100 atm or so).

275

Page 288: 457454 Physical Chemistry Quantum Chemistry

Likewise at constant P

V (T ) = V0 +

µ∂V

∂T

¶P| z

V0α

(T − T0) +

µ∂V

∂T

¶2T

(T − T0)2 + · · · (42.6)

so,

V (T ) ≈ V0 [1 + α(T − T0)] . (42.7)

As one final point, we can apply the cyclic rule for partial derivatives to determine

the ratio ακT:

α

κT=

¡∂V∂T

¢P

−¡∂V∂P

¢T

cyclic=rule

µ∂P

∂T

¶V

(42.8)

42.2. Heat Capacity of Gases Revisited

This section is a review from the first semester with an additional example beyond

the ideal gas.

42.2.1. The Relationship Between CP and CV

To find how CP and CV are related we begin with

CP =

µ∂H

∂T

¶P

,H = U + PV (42.9)

so

CP =

µ∂ (U + PV )

∂T

¶P

=

µ∂U

∂T

¶P

+ P

µ∂V

∂T

¶P

(42.10)

note¡∂U∂T

¢Pis not CV we need

¡∂U∂T

¢V. Use an identity of partial derivativesµ

∂U

∂T

¶P

=

µ∂U

∂T

¶V

+

µ∂U

∂V

¶T

µ∂V

∂T

¶P

(42.11)

276

Page 289: 457454 Physical Chemistry Quantum Chemistry

thus

CP =

µ∂U

∂T

¶V

+

µ∂U

∂V

¶T

µ∂V

∂T

¶P

+ P

µ∂V

∂T

¶P

(42.12)

= CV +

µ∂V

∂T

¶P

∙µ∂U

∂V

¶T

+ P

¸.

Recall the expression for internal pressure¡∂U∂V

¢T= T

¡∂P∂T

¢V− P . Then

CP = CV +

µ∂V

∂T

¶P

∙T

µ∂P

∂T

¶V

− P/ + P/¸

(42.13)

Finally

CP = CV + T

µ∂V

∂T

¶P

µ∂P

∂T

¶V

(42.14)

For solids and liquids: µ∂V

∂T

¶P

= V α,

µ∂P

∂T

¶V

κT(42.15)

so

CP = CV +α2TV

κT(42.16)

For gases we need the equation of state which often is conveniently explicit in P

or V but not both

1. Explicit in P : Replace µ∂V

∂T

¶P

with −¡∂P∂T

¢V¡

∂P∂V

¢T

(42.17)

2. Explicit in V : Replace µ∂P

∂T

¶V

with −¡∂V∂T

¢P¡

∂V∂P

¢T

(42.18)

277

Page 290: 457454 Physical Chemistry Quantum Chemistry

Examples

1. Ideal gas (equation of state: PV = nRT ): This equation is easily made

explicit in either P or V so we don’t need any of the above replacements

CP = CV + T

µ∂V

∂T

¶P

µ∂P

∂T

¶V

(42.19)

= CV + TnR

P

nR

V=

nRT

PV= nR

Thus CP = CV + nR or CPm = CVm +R

2. One term viral equation (equation of state: V = nRTP+nB). This is explicit

in V so use case 2 above

CP = CV + T

µ∂V

∂T

¶P

µ∂P

∂T

¶V

= CV − T

µ∂V

∂T

¶P

¡∂V∂T

¢P¡

∂V∂P

¢T

(42.20)

The partial derivatives areµ∂V

∂T

¶P

=nR

P+ nB0,

µ∂V

∂P

¶T

= −nRTP 2

, (42.21)

so

−¡∂V∂T

¢P¡

∂V∂P

¢T

= −nRP+ nB0

−nRTP 2

=n/ P (R+ PB0)

n/ RT. (42.22)

Thus

CP = CV + T/µnR

P+ nB0

¶ÃP (R+ PB0)

RT/

!(42.23)

= CV + nR

µ1 +

PB0

R

¶2or

CPm = CVm +R

µ1 +

PB0

R

¶2(42.24)

278

Page 291: 457454 Physical Chemistry Quantum Chemistry

42.3. Expansion of Gases

Expanding gases do work:

−w =Z V2

V1

PexdV (42.25)

As we learned last semester the value of w depends on Pex during the expansion.

Recall that if the expansion is reversible, there is always an intermediate equi-librium throughout the expansion. Namely Pgas = Pex. So,

−wrev =Z V2

V1

PgasdV (42.26)

For an ideal gas (P = nRTV) this becomes

−wrev =Z V2

V1

nRT

VdV = nRT ln

µV2V1

¶(42.27)

Also recall that −wrev is the maximum possible work that can be done in an

expansion. −wrev = −wmax.

42.3.1. Isothermal and Adiabatic expansions

We shall consider two limits for the expansion of gases

1. Isothermal expansion T is constant

2. Adiabatic expansion q = 0.

Isothermal expansion

• For the case of a ideal gas, U(T, V ) = U(T ) (independent of V ). So for

isothermal expansion 4U = 0 = q + w =⇒ q = −w.

279

Page 292: 457454 Physical Chemistry Quantum Chemistry

Adiabatic expansion

• Since q = 0, dU = dw = −PexdV = −PdV (reversible).

• For an ideal gasdU = −PdV = −nRT

VdV (42.28)

42.3.2. Heat capacity CV for adiabatic expansions

Considering an ideal gas going adiabatically from (T1, V1) to (T2, V2).

Recall

CV =

µ∂U

∂T

¶V

=⇒ dU = CV dT (42.29)

So from above

CV dT =−nRTV

dV =⇒ CV dT

T=−nRdV

V(42.30)

Going from (T1, V1) to (T2, V2):Z T2

T1

CV

TdT =

Z V2

V1

−nRV

dV. (42.31)

If CV (T ) is reasonably constant over the internal T1 to T2 then this is approxi-

mately

CV ln

µT2T1

¶= −nR ln

µV2V1

¶(42.32)

where CV =12(CV (T1) + CV (T2)) . Or, in terms of molar heat capacity

CVm ln

µT2T1

¶= −R ln

µV2V1

¶(42.33)

280

Page 293: 457454 Physical Chemistry Quantum Chemistry

42.3.3. When P is the more convenient variable

What if P is the more convenient variable? Then use H instead of U

Let us still consider an adiabatic expansion

H = U + PV, dH = dU + PdV + V dP (because both P and V can, in general,

change)

dH = dq + dw + P/ dV/ + V dP (42.34)

dH = V dP.

Now,

CP =

µ∂H

∂T

¶P

=⇒ dH = CpdT = V dP (42.35)

For an ideal gas this becomes

CpdT =nRT

PdP (42.36)

Going from (T1, P1) to (T2, P2):Z T2

T1

CP

TdT =

Z P2

P1

nR

PdP. (42.37)

If CP (T ) is reasonably constant over the internal T1 to T2 then this is approxi-

mately

CP ln

µT2T1

¶= nR ln

µP2P1

¶(42.38)

where CP =12(CP (T1) + CP (T2)) . Or, in terms of molar heat capacity

CPm ln

µT2T1

¶= R ln

µP2P1

¶(42.39)

From the above two cases

ln

µT2T1

¶=

R

CPm

ln

µP2P1

¶=−RCVm

ln

µV2V1

¶(42.40)

281

Page 294: 457454 Physical Chemistry Quantum Chemistry

So

ln

µP2P1

¶= −CPm

CVm| z ≡γ

ln

µV2V1

¶(42.41)

hence

ln

µP2P1

¶= −γ ln

µV2V1

¶= γ ln

µV1V2

¶= ln

µV1V2

¶γ

(42.42)

Thus µP2P1

¶=

µV1V2

¶γ

⇒ P2Vγ2 = P1V

γ1 , (42.43)

but PiVγi are arbitrary so this implies PV γ = constant (** NOTE: The axes

should be reversed **)

42.3.4. Joule expansion

Consider a gas expanding adiabatically against a vacuum (Pex = 0). In this case

q = 0 (adiabatic) and w = 0 (since −dw = PexdV ).

282

Page 295: 457454 Physical Chemistry Quantum Chemistry

This implies

4U = q + w = 0. (42.44)

Internal energy is constant.

We want to find¡∂T∂V

¢U.

Identity: µ∂T

∂V

¶U

= −µ∂T

∂U

¶V| z

1/CV

µ∂U

∂V

¶T

=1

CV

µ∂U

∂V

¶T

(42.45)

For an ideal gas¡∂U∂V

¢T= 0 (since U(T, V ) = U(T )). Thus in as much as the

gas can be considered ideal¡∂T∂V

¢U= 0. That is, for Joule type expansion the

temperature of the gas does not change. For real gases this is not strictly equal

to zero.

42.3.5. Joule-Thomson expansion

Consider the adiabatic expansion as illustrated by the figure below

283

Page 296: 457454 Physical Chemistry Quantum Chemistry

The work done on the left is

wL = −P14V = −P1(0− V1) = P1V1. (42.46)

The work done on the right is

wR = −P24V = −P2(V2 − 0) = −P2V2. (42.47)

Now,

4U = U2 − U1 = wL + wR = P1V1 − P2V2 (42.48)

Thus

U2 + P2V2 = U1 + P1V1 ⇒ H2 = H1 (42.49)

For Joule-Thomson expansion the enthalpy is constant.

We want to find¡∂T∂V

¢H≡ μ. (the Joule-Thomson coefficient).

Identity: µ∂T

∂P

¶H

= −µ∂T

∂H

¶P| z

1/CP

µ∂H

∂P

¶T

=1

CP

µ∂H

∂P

¶T

= μ (42.50)

284

Page 297: 457454 Physical Chemistry Quantum Chemistry

Recall the useful identity µ∂H

∂P

¶T

= V − T

µ∂V

∂T

¶P

(42.51)

Thus

μ =−V + T

¡∂V∂T

¢P

CP(42.52)

Example: The one term virial equation: (equation of state PV = nRT + nB)

μ =1

CP

µ−nRTP

− nB +nRT

P+ nTB0

¶(42.53)

μ =−B + TB0

CPm.

Limts:

• Low T : B0 is positive and B is negative, so μ is positive–the gas cools upon

expansion

• High T : B0 is nearly zero and B is positive, so μ is negative–the gas warms

upon expansion

• The Joule-Thomson inversion temperature is the temperature where μ = 0.

285

Page 298: 457454 Physical Chemistry Quantum Chemistry

43. Entropy of Gases

43.1. Calculation of Entropy

Entropy must be calculated along reversible paths. This is not a problem though

since entropy is a state function.

Entropy change for changes in temperature.

• At constant V :

— dU = dq + dwdq=CV dT=⇒ dU = CV dT, but also dU = TdS. So

dS =CV

TdT =⇒4S =

Z T2

T1

CV

TdT. (43.1)

At constant P : (use H = U + PV instead of U)

— dH = dU+PdV +V dP = dq−PdV +PdV +V dP . So dH = dqdq=CP dT=⇒

dH = CPdT, but also dHdq=TdS= TdS. So

dS =CP

TdT =⇒4S =

Z T2

T1

CP

TdT. (43.2)

Isothermal expansion of an ideal gas (PV = nRT ):

• Recall that for isothermal expansion of an ideal gas dU = 0 = TdS − PdV

⇒ dS = PdVT

.

286

286

Page 299: 457454 Physical Chemistry Quantum Chemistry

• Using the equation of state

dS =nRdV

V=⇒ 4S =

Z V2

V1

nR

VdV = nR ln

V2V1

. (43.3)

• Using the equation of state to express V1 and V2 in terms of P1 and P2.

dS = nR lnV2V1= nR ln

n/ R/ T/P2

n/ R/ T/P1

= −nR ln P2P1

. (43.4)

If two variables change in going from the initial to final states break the path into

two paths in which only one variable changes at a time.

Entropy of Mixing of an ideal gas

• Since the gas is ideal, there are simply two separate equations:

4SA = nAR lnVA + VB

VA, 4SB = nBR ln

VB + VAVB

(43.5)

and

4Smix = 4SA +4SB (43.6)

287

Page 300: 457454 Physical Chemistry Quantum Chemistry

• Recall Avogadro’s principle: n ∝ V for an ideal gas. So.

4Smix = R

⎛⎜⎜⎜⎝nA lnnA + nB

nA| z 1/XA

+ nB lnnB + nA

nB| z 1/XB

⎞⎟⎟⎟⎠ = −R (nA lnXA + nB lnXB)

(43.7)

43.1.1. Entropy of Real Gases

Consider the question: How does S → Sideal as P → 0 ?

Use Maxwell relation¡∂S∂P

¢T= −

¡∂V∂T

¢Pand single term viral equation, V =

nRTP+ nB.

So µ∂S

∂P

¶T

= −µ∂V

∂T

¶P

= −nRP− nB0 (43.8)

Hence

dS =

µ−nR

P− nB0

¶dP

→→=⇒ S2 − S1 = −nR ln

P2P1− nB0(P2 − P1) (43.9)

For an ideal gas B0 = 0, so

Sideal2 − Sideal1 = −nR ln P2P1

(43.10)

Thus

S2 − S1 = Sideal2 − Sideal1 − nB0(P2 − P1) (43.11)

Letting P1 → 0 and P2 → P θ (Standard pressure 1 bar), this becomes

S2 − S/ ideal

1= Sideal2 − S/ ideal

1− nB0(P2 − P1) (43.12)

Defining Sideal2 , P2 → P θ as Sθ. So,

S(P θ) = Sθ − nB0P θ (43.13)

288

Page 301: 457454 Physical Chemistry Quantum Chemistry

The entropy at any P and T can be obtained expresses as

S(T, P ) = Sideal(T, P )− nB0P (43.14)

Thus

S(T, P ) = Sθ(T )− nR lnP

P θ− nB0P (43.15)

289

Page 302: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 3

Listed here are some of the key equations for Exam 3. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• The Maxwell’s distribution of speeds is

F (v) = 4π

µm

2πkbT

¶ 32

e−mv2

2kbT v2. (43.16)

• The average speed of a particle is

hvi =r8RT

πM(43.17)

• The mean free path isλ =

RT√2PLπσ2

(43.18)

290

290

Page 303: 457454 Physical Chemistry Quantum Chemistry

• The reaction velocity isv =

1

vi

d[I]dt

(43.19)

• The relation between the rate constant and the thermodynamic equilibriumconstant is

Kc =kfkr

(43.20)

• The Arrhenious equationk = Ae−

EaRT (43.21)

• Important thermodynamic relation:

4G = 4H − T4S (43.22)

• Eyring’s equation is

k =kbT

he−

4G‡RT =

kbT

he−

4H‡RT e

4S‡R (43.23)

• The van der Waals gas equation of state:

P =RT

Vm − b− a

V 2m

. (43.24)

• Compressibility Factor:z =

PV

nRT=

PVmRT

. (43.25)

• The virial series is

z = 1 +B(T )

µ1

Vm

¶+ C(T )

µ1

Vm

¶2+D(T )

µ1

Vm

¶3+ · · · . (43.26)

• Relation between heat capacities for an ideal gas:

CPm = CVm +R (43.27)

291

Page 304: 457454 Physical Chemistry Quantum Chemistry

Part VIII

More Thermodyanmics

292

292

Page 305: 457454 Physical Chemistry Quantum Chemistry

44. Critical Phenomena

44.1. Critical Behavior of fluids

The point on the top of the coexistence curve is called the critical point. It is

characterized by a critical temperature, Tc, and a critical density ρc.

Law of rectilinear diameters: The average density [ρave =12(ρliq + ρvap)] is

linear in temperature.

293

293

Page 306: 457454 Physical Chemistry Quantum Chemistry

44.1.1. Gas Laws in the Critical Region

The vapor pressure of a substance is taken from the gas laws as the pressures

where A1 = A2 in the above figure.

Simple gas laws do not work well near critical points.

294

Page 307: 457454 Physical Chemistry Quantum Chemistry

44.1.2. Gas Constants from Critical Data

Consider the van der Waals equation at the critical point (Pc, Tc, Vmc)

Pc =RTc

Vmc − b− a

V 2mc

. (44.1)

There is an inflection point ( dPdVm

= 0, d2P

dV 2m= 0) at the critical point. So, setting

the first and second derivatives at the critical point equal to zero we get

dP

dVm

¯c

=−RTc

(Vmc − b)2+2a

V 3mc

= 0 (44.2)

andd2P

dV 2m

¯c

=2RTc

(Vmc − b)3− 6a

V 4mc

= 0 (44.3)

solving these three equations for Pc, Tc and Vmc gives

Vmc = 3b, (44.4)

Tc =8a

27bR, (44.5)

Pc =a

27b2. (44.6)

These values can be used to find the compressibility factor, z, at the critical point

zc =PcVmc

RTc=3

8= 0.375. (44.7)

Notice that both a and b whose values depend on the particular gas have dropped

out. That is (for the van der Waals Equation) zc = 0.375 for all gases.

The other equations of state give similar resultsvan der Waals Berthelot Dieterici Redlich-Kwong

zc 3/8 = 0.375 3/8 = 0.375 2/e2 ' 0.27 0.33

295

Page 308: 457454 Physical Chemistry Quantum Chemistry

44.2. The Law of Corresponding States

We have found that zc is predicted by the equations of state to be independent of

the particular gas. This is actually not too far from the truth experimentally.

One can define unitless “reduced” variables Tr = T/Tc, Pr = P/Pc, and Vr = V/Vc.

Then zr =PrVrRTr

.

zr is a “universal” function–it is nearly the same for all gasses.

∗ ∗ See Fig. 1.18 Laidler&Meiser ∗ ∗

44.3. Phase Equilibrium

Consider a homogeneous substance consisting of two phases α and β at a constant

T and V.

Suppose some amount of material, dn, goes from α→ β

• (dAα)T = −PdVα − μαdn

• (dAβ)T = −PdVβ + μβdn

• (dA)T,V = −P= 0 since V is constantz | (dVα + dVβ) +

¡μβ − μα

¢dn

For a spontaneous process A deceases (dA < 0)

At equilibrium dA = 0. This implies μβ = μα is the condition for equilibrium.

When α, β denote liquid (or solid) and vapor phases, then for a given T , the

pressure of the system when μβ = μα is the called the vapor pressure of the

material at temperature T.

296

Page 309: 457454 Physical Chemistry Quantum Chemistry

For phase changes at constant T and P then (dG)T,P =¡μβ − μα

¢dn. So again

μβ = μα is the condition for equilibrium.

44.3.1. The chemical potential and T and P

How does μ vary with T and P?

Generally for homogeneous substances,

dG = −SdT + V dP + μdn (44.8)

Now,

S = −µ∂G

∂T

¶P,n

(44.9)

So, µ∂S

∂n

¶P,T

= − ∂

∂n

∂G

∂T= − ∂

∂T

∂G

∂n= −

µ∂μ

∂T

¶P

. (44.10)

But S = nSm(T, P ) so, µ∂μ

∂T

¶P

= −Sm. (44.11)

Similarly, µ∂μ

∂P

¶T

= Vm. (44.12)

Now the total differential of μ is

dμ(T, P ) =

−Smz | µ∂μ

∂T

¶P

dT +

Vmz | µ∂μ

∂P

¶T

dP (44.13)

dμ(T, P ) = −SmdT + VmdP

297

Page 310: 457454 Physical Chemistry Quantum Chemistry

44.3.2. The Clapeyron Equation

At equilibrium μβ = μα so,

−SmαdT + VmαdP = −SmβdT + VmβdP (44.14)

NowdP

dT=

Smα − Smβ

Vmα − Vmβ=−4φSm−4φVm

4S=4HT=4φHm

T4φVm(44.15)

This is the Clapeyron Equation

dP

dT=4φHm

T4φVm(44.16)

44.3.3. Vapor Equilibrium and the Clausius-Clapeyron Equation

The above Clapeyron equation applies to any phase transition; consider the liquid-

vapor phase transition.

Now

4vV = Vm,vap − Vm,liq ' Vm,vap (44.17)

Assuming the vapor phase obeys the ideal gas equation of state,

4vV =RT

P(44.18)

Substituting this into the Clapeyron equation gives

dP

dT=4vHm

T RTP

=4vHmP

RT 2(44.19)

Collecting the T ’s on one side of the equation and the P ’s on the other we get

dP

P=4vHm

R

dT

T 2(44.20)

Now we identify dPP= d(lnP ) and dT

T 2= −d(1/T ) so this becomes

d(lnP ) = −4vHm

Rd(1/T ) (44.21)

298

Page 311: 457454 Physical Chemistry Quantum Chemistry

Rearranging again leads to

d(lnP )

d(1/T )= −4vHm

R(44.22)

This is the Clausius-Clapeyron equation.

44.4. Equilibria of condensed phases

Examples

• Solid—liquid

— ice—water, most other common liquids

• Solid—solid

— rhombic sulfur—monoclinic sulfur

— grey tin—white tin

— graphite—diamond

For example a diamond at STP is metastable with respect to graphite.

“A diamond is not forever!”

At equilibrium μα = μβ this implies (for incompressible liquids and solids)

μªα + Vmα(P − Pª) = μªβ + Vmβ(P − Pª) (44.23)

This can be rearranged so that terms independent of pressure (the standard chem-

ical potentials) are one side and the terms that depend of pressure are on the other

side

μªα − μªβ = (Vmβ − Vmα) (P − Pª) (44.24)

299

Page 312: 457454 Physical Chemistry Quantum Chemistry

Thus for any given T only one P allows for equilibrium.

Recall the Clapeyron equation

dP

dT=4fHm

T4fVm=

Hmβ −Hmα

T (Vmβ − Vmα)(44.25)

We make the good approximation that 4fHm is independent of T and solve the

Clapeyron equationZ →

dT

T=4fVmdP

4fHm⇒ ln

TfTªf

=4fVm(P − Pª)

4fHm(44.26)

where Tªf is the freezing temperature at standard pressure (1 bar).

44.5. Triple Point and Phase Diagrams

Definitions

• Phase Diagram: A graph of P vs. T for a system which shows the lines

of equal chemical potential

• Critical Point: The terminal point of the liquid-vapor line. At temper-atures above the critical point there is no distinction between vapor and

liquid.

• Triple Point: The point where all three phases coexist in equilibrium:

μsolid = μliq = μvap (44.27)

300

Page 313: 457454 Physical Chemistry Quantum Chemistry

45. Transport Properties of Fluids

Transport properties of matter deal with the flow (or flux) of some property along

a gradient of some other property.

Flux: movement of something through a unit area.

We now consider three transport properties of fluids:

1. Diffusion: The flux of material down a concentration gradient

2. Viscosity: The flux of momentum down a velocity gradient

3. Thermal Conductivity: The flux of energy down a temperature gradient

∗ ∗ See Transport Phenomena handout ∗ ∗

45.1. Diffusion

At equilibrium concentration on a bulk solution will be uniform.

So if there exists a concentration gradient there will be a net flux, J, of material

from high concentration to low concentration so as to establish an equilibrium.

J =1

A

dn

dt(45.1)

301

301

Page 314: 457454 Physical Chemistry Quantum Chemistry

The flux of material through a plane depends on the concentration difference

J = −DdC

dx=⇒ 1

A

dn

dt= −DdC

dx

where D is the diffusion constant

1

A

dn

dt= −DdC

dx(45.2)

This is Fick’s first law of diffusion (in one dimension).

The change in concentration in a lamina between x and dx with time is given by

the flux in minus the flux out of the lamina:

∂C

∂t=

J(x)− J(x+ dx)

dx= −∂J

∂x(45.3)

Using Fick’s first law for J∂C

∂t=

∂xD∂C

∂x. (45.4)

If D is truly constant we get Fick’s second law of diffusion:

∂C

∂t= D

∂2C

∂x2. (45.5)

302

Page 315: 457454 Physical Chemistry Quantum Chemistry

The solution of this partial differential equation depends on the boundary condi-

tions. Numerous methods of solution exist for this equation but they are beyond

the scope of the course.

The solution for two special boundary conditions are of interest and will simply

be presented here without derivation

1. Point source solution

C(x, t) =C0

2√πDt

e−x2

4Dt (45.6)

2. Step function solution

C(x, t) = C0

"1

2− 1√

π

Z x√4Dt

0

e−y2

dy

#(45.7)

=1

2C0

∙1− erf

µx√4Dt

¶¸=

1

2C0 erfc

µx√4Dt

¶where erf and erfc are tabulated functions respectively called the error func-

tion and complementary error function.

45.2. Viscosity

Viscosity, η, is the resistance to differential fluid flow, i.e., The tendency of a

liquid to flow at the same velocity throughout.

303

Page 316: 457454 Physical Chemistry Quantum Chemistry

The frictional (viscous) force is F = ηAdvdx. (The units of η are mass

lenght·time . 1 poise

= gcm·s .)

Poiseuille’s Formula

• Applies to Laminar (nonturbulent) flow

• For a liquid flowing trough a tube (radius r, length l), the volume of flow

4V in time 4t is4V

4t= −πr

44P

8ηl(45.8)

where 4P is the driving pressure, i.e., the difference in pressure on either

side of the tube.

• For a gas4V

4t=

πr2

16ηl

µP 2i − P 2

f

P0

¶(45.9)

where Pi is the inlet pressure, Pf is the outlet pressure and P0 is the pressure

at which the volume is read.

Stoke’s law: spheres falling through fluids

304

Page 317: 457454 Physical Chemistry Quantum Chemistry

• The frictional force (exerted upwards) is proportional to velocity: Ff = −fv.Stokes showed f = 6πηr

• Gravitational force (exerted downwards): Fg = 4πr3

3(ρ−ρ0)g, where g is the

gravitational acceleration (9.8 m/s2).

• Terminal velocity is reached when Ff + Fg = 0 giving

−fvterm +4πr3

3(ρ− ρ0)g = 0 (45.10)

vterm =4πr3(ρ− ρ0)g

3f

using f = 6πηr

vterm =4π/ r3/(ρ− ρ0)g

3¡6π/ ηr/

¢ =2r2(ρ− ρ0)g

9η(45.11)

• Related to diffusion constant:

D =kT

f

f=6πηr=

kT

6πηr(45.12)

45.3. Thermal conductivity

(This section closely follows parts of chapter 8 in Transport Phenomena by R.B.

Bird, W.E. Stewart and E. N. Lightfoot Wiley New York 1960)

The thermal conductivity, κ, of a material is a measure of the tendency of energy

in the form of heat to flow through the material.

Consider a slab of solid material of area A between two large parallel plates a

distance D apart. The plates are held at constant but different temperatures T1and T2 (T1 > T2) for a sufficiently long time that a steady state exists.

305

Page 318: 457454 Physical Chemistry Quantum Chemistry

Under such conditions, a linear steady state temperature distribution across the

material is established. And a constant rate of heat flow dqdtis needed to maintain

the temperature difference 4T = (T1 − T2)

1

A

dq

dt= −κ4T

D. (45.13)

If we take the limit where D becomes infinitesimally small (D→ dx) we obtain a

differential form of this equation:1

A

dq

dt= Qf = −κ

dT

dx, (45.14)

where Qf is the heat flux. This is called Fourier’s law of heat conduction(one-dimensional version).

Thermal conductivities are positive quantities so Fourier’s law says that heat flow

down a temperature gradient, i.e., from hot to cold.

45.3.1. Thermal Conductivity of Gases and Liquids

∗ ∗ See Reduced thermal conductivity handout ∗ ∗

From this handout we see that typically the thermal conductivity of gases at low

densities increases with increasing temperature, whereas the thermal conductivity

of most liquids decrease with increasing temperature.

306

Page 319: 457454 Physical Chemistry Quantum Chemistry

45.3.2. Thermal Conductivity of Solids

For the most part, the thermal conductivity of solids have to be determined ex-

perimentally because many factors contributing to the thermal conductivity are

difficult to predict.

In general metals are better heat conductors than nonmetals and crystals are

better heat conductors than amorphous materials.

Dry porous materials are poor heat conductors

Rule of Thumb: Thermal conductivity and electrical conductivity go hand inhand.

The Wiedemann, Frantz and Lorenz equation relates the thermal conductivity to

electrical conductivity, κel for pure metals:

κ

κelT= L = const. (45.15)

where L is the Lorenz number (typically 22 to 29 × 10−9 V2/K2).

The Lorenz number is taken as constant because it is only a very weak function

of temperature with a change of 10 to 20% per 1000 degrees being typical.

The Wiedemann, Frantz and Lorenz equation breaks down at low temperature

because metals become superconductive. There is no analog to superconductivity

for thermal conductivity.

307

Page 320: 457454 Physical Chemistry Quantum Chemistry

46. Solutions

Solutions are mixtures of two or more pure substances. So, in addition to the

parameters needed to characterize a pure substance, one also needs to keep track

of the amount of individual species in solution

46.1. Measures of Composition

There are several measures of composition of solutions

• mole ratio r = n1n2

• mole fraction X2 =n2

n1+n2, X1 = 1−X2

• molality m = 1000X2

M1X1, where M1 is the molecular weight of species 1

• Molarity c2 = n2L solution

46.2. Partial Molar Quantities

Thermodynamic properties, in general change upon mixing

4mix = properties of soln −X

properties of pure. (46.1)

For example,

4mixV = Vsoln − Vsolute − Vsolvent (46.2)

Consider a thermodynamic quantity, say, volume.

308

308

Page 321: 457454 Physical Chemistry Quantum Chemistry

In general, it is a function of T, P, n1 and n2: V (T, P, n1, n2). So, the total

derivative is

dV =

µ∂V

∂T

¶P,n1,n2

dT +

µ∂V

∂P

¶T,n1,n2

dP +

µ∂V

∂n1

¶T,P,n2

dn1 +

µ∂V

∂n2

¶T,P,n1

dn2,

(46.3)³∂V∂ni

´T,P,nj

≡ Vi, the partial molar volume.

Similarly

dG =

µ∂G

∂T

¶P,n1,n2

dT +

µ∂G

∂P

¶T,n1,n2

dP +

µ∂G

∂n1

¶T,P,n2

dn1 +

µ∂G

∂n2

¶T,P,n1

dn2,

(46.4)³∂G∂ni

´T,P,nj

≡ μi.

So now for the more general case of mixtures the chemical potential of a species

of the partial molar free energy for that species, rather than simply the molar free

energy as it was earlier.

46.2.1. Notation

The study of solutions brings with it a large number of symbols which we collect

here for future reference.Material

Pure liquid i V •i H•

i S•i G•i

Pure liquid i per mole V •mi H•

mi S•mi μ•i

Whole solution V H S G

Solution/(total moles) Vm Hm Sm Gm

Partial molar of i in solution Vi Hi Si μi

Apparent molar (of solute) φV φH

Reference state V ªi Hª

i Sªi μªi

309

Page 322: 457454 Physical Chemistry Quantum Chemistry

46.2.2. Partial Molar Volumes

Consider the partial molar volume

For constant T and P

dV = V1dn1 + V2dn2 (46.5)

Now, Vi depends on concentration, so change each amount of substance propor-

tional to the amount substance present,

dn1 = n1dλ, dn2 = n2dλ. (46.6)

So,

dV =¡V1n1 + V2n2

¢dλ

dλ=⇒ V = V1n1 + V2n2 (46.7)

That is, the total volume of the solution is equal to the sum of the partial molar

volumes each weighted by their respective number of moles.

The total volume, however, is not necessarily the mole weighted sum of the vol-

umes of each component in its pure (unmixed) state. More specifically

4mixV = V − (V •m1n1 + V •

m2n2) (46.8)

=¡V1n1 + V2n2

¢− (V •

m1n1 + V •m2n2)

=¡V1 − V •

m1

¢n1 +

¡V2 − V •

m2

¢n2

4mixV can be positive, negative or zero.

For example,

1. one unit of baseballs are mixed with one unit of basketballs. 4mixV < 0.

2. one unit of baseballs are mixed with one unit of books. 4mixV > 0.

310

Page 323: 457454 Physical Chemistry Quantum Chemistry

46.3. Reference states for liquids

For liquids there are two more convenient ideal states

1. neat (pure) solvent limit

1. all neighboring molecules are same as the given molecule

2. the ideal state for Raoult’s law

2. infinite dilution limit

1. all neighboring molecules are different than the given molecule

2. the ideal state for Henry’s law

Raoult’s law limit Henry’s law limit

46.3.1. Activity (a brief review)

Recall that activity gives a measure of the deviation of the real state from some

reference state

311

Page 324: 457454 Physical Chemistry Quantum Chemistry

Also recall that the mathematical definition of activity ai of some species i is

implicitly stated as

limζ→ζª

aig(ζ)

= 1 (46.9)

where g(ζ) is any reference function (e.g., pressure, mole fraction, concentration

etc.), and ζª is the value of ζ at the reference state.

This implicit definition is awkward so for convenience one defines the activity

coefficient as the argument of the above limit,

γi ≡aig(ζ)

(46.10)

which we can rearrange as

ai = γig(ζ). (46.11)

The definition of activity implies that γi = 1 at g(ζª) (the reference state)

That is γi → 1 as the real system approaches the reference state.

Connecting with the chemical potential we saw last semester that the deviation

of the chemical potential at the state of interest versus at the reference state is

determined by the activity at the current state (the activity at the reference state

is unity by definition).

μi − μªi = RT ln ai. (46.12)

46.3.2. Raoult’s Law

In discussing both Raoult’s law and Henry’s law, we are describing the behavior

of a liquid solution by measuring the vapor (partial) pressures of the components

312

Page 325: 457454 Physical Chemistry Quantum Chemistry

For simplicity we consider here only a two component solution.

dG = μ1dn1 + μ2dn2. (46.13)

Take differential change along a line of constant concentration, so

dG = (μ1n1 + μ2n2) dλ (46.14)

then

G = μ1n1 + μ2n2. (46.15)

Recall that

4mixG = G(soln)−G(pure components) (46.16)

Hence,

4mixG = μ1n1 + μ2n2 − μ•1n1 − μ•2n2 (46.17)

= (μ1 − μ•1)n1 + (μ2 − μ•2)n2.

Now,

μ1 − μ•1 = RT lnaia•i

low P' RT lnPi

P •i

, (46.18)

where Pi is the vapor pressure of the ith component above the solution.

313

Page 326: 457454 Physical Chemistry Quantum Chemistry

Thus

4mixG = RT

µn1 ln

a1a•1+ n2 ln

a2a•2

¶(46.19)

or at low P

4mixG = RT

µn1 ln

P1P •1

+ n2 lnP2P •2

¶(46.20)

46.3.3. Ideal Solutions (RL)

Raoult’s Law:

Pi = XiP•i (46.21)

That is, the vapor partial pressure of a component of a mixture is equal to the

mole fraction of the component times the vapor pressure that the component

would have if it were pure.

The change in free energy upon mixing for solutions ideally obeying Raoult’s law

is

4id(RL)mix G = RT

Ãn1 ln

X1P•1/

P •1/+ n2 ln

X2P•2/

P •2/

!(46.22)

4id(RL)mix G = RT (n1 lnX1 + n2 lnX2) (46.23)

Again, this is for an ideal solution in the Raoult’s Law sense.

From

S = −µ∂G

∂T

¶P

and H = −µ∂ (G/T )

∂ (1/T )

¶P

, (46.24)

the entropy of mixing for an ideal Raoult solution is

4id(RL)mix S = −R (n1 lnX1 + n2 lnX2) (46.25)

and the enthalpy of mixing is

4id(RL)mix H = 0 (46.26)

314

Page 327: 457454 Physical Chemistry Quantum Chemistry

(since G/T is independent of 1/T ).

The Reference State (RL)

Let us apply the definition of activity for the Raoult’s law reference state.

The reference function is g(ζ) = ζ = Xi. and the reference state is Xi = 1

So,

limXi→1

a(RL)i

Xi= 1 (46.27)

implies

a(RL)i = γ

(RL)i Xi, (46.28)

and γ(RL)i → 1 as Xi → 1

Deviations from Raoult’s Law

Raoult’s law is a purely statistical law. It does not require any kind of interaction

among the constituent particle making up the solution.

Since, in reality, there are specific interactions between particles, real solutions

generally deviate from Raoult’s law.

The physical interpretation of deviation from Raoult’s law is

• positive deviation: the molecules prefer to be around themselves rather thanother types of molecules.

• negative deviation: the molecules prefer to be around other types of mole-cules than themselves.

• no deviation: the molecules have no preference.

315

Page 328: 457454 Physical Chemistry Quantum Chemistry

It is very important to note that this deviation from Raoult’s law is a property of

the solution and NOT any given component.

For example, for a given component, mixing with one substance may lead to

a positive deviation but mixing with another substance may lead to a negative

deviation.

Positive deviation from Raoult’s lawNegative deviation from Raoult’s law

46.3.4. Henry’s Law

Henry’s Law:Pi = kXiXi, (46.29)

where kXi is the Henry’s law constant,

kXi= lim

Xi→0

µPi

Xi

¶(46.30)

Henry’s law applies to the solute not to the solvent and becomes more correct for

real solution as the concentration of solute goes to zero (Xi → 0), i.e., at infinite

dilution.

316

Page 329: 457454 Physical Chemistry Quantum Chemistry

The Reference State (HL)

Referring to the definition of activity again we see that the reference function is

g(ζ) = ζ = Xi. and the reference state is now Xi = 0

So,

limXi→0

a(HL)i

Xi= 1 (46.31)

implies

a(HL)i = γ

(HL)i Xi, (46.32)

and γ(HL)i → 1 as Xi → 0

If instead of mole fraction, molality or molarity is used then

a(HL)i = γ(HL)mi

mi (46.33)

and

a(HL)i = γ

(HL)Mi

Mi (46.34)

respectively.

Comparison of Raoult’s Law and Henry’s Law

Both Raoult’s law and Henry’s law become better approximations for real solu-

tions as the solution becomes pure. But, they apply to opposite species in the

solution. Raoult’s law applies to the dominant species, X1 → 1, whereas Henry’s

law applies to the subdominant species X2 → 0. So, in summary

• Raoult’s law: γ1 → 1 as X1 → 1

• Henry’s law: γ2 → 1 as X2 → 0

317

Page 330: 457454 Physical Chemistry Quantum Chemistry

46.4. Colligative Properties

Colligative properties: Properties of dilute solutions that are independent ofthe chemical nature of the solute

Examples

• Freezing point depression

• Boiling point elevation

• Vapor pressure lowering

• Osmotic pressure

We will consider the examples of freezing point depression and osmotic pressure

46.4.1. Freezing Point Depression

At Tf (freezing point), μ1(solid)| z μs1

= μ1(soln).

318

Page 331: 457454 Physical Chemistry Quantum Chemistry

Using the Raoult’s law reference state (since we are interested in the behavior of

the dominant species), μ1(soln) = μ•1 +RT ln a1:

μs1 = μ•1 +RT ln a1 (46.35)

Rearranging this and taking the derivative with respect to T yields

∂T

→ln a1 =

1

RT(μs1 − μ•1) =⇒

∂ ln a1∂T

=−1RT 2

µ∂μs1∂T− ∂μ•1

∂T

¶(46.36)

Now, using ∂μ∂T= H and integrating we getZ →

→d ln a1 =

µ−1RT 2

(Hs1 −H•

1)

¶dT =

4fH

RT 2dT (46.37)

ln a1 =

Z Tf

T•f

4fH

RT 2dT

For small changes in the freezing point we may approximate T by T •f in the

integrand. So,

ln a1 'Z Tf

T•f

4fH

RT •2fdT =

−4fH

RT •2fΘ, (46.38)

where Θ ≡ T •f − Tf . The freezing point depression is

Θ = −RT •2f ln a1

4fH

46.4.2. Osmotic Pressure

We consider the osmotic pressure at a constant temperature, T. (so, dG = V dP ).

319

Page 332: 457454 Physical Chemistry Quantum Chemistry

In the above figure μ1(left) = μ1(right), hence

μ•1 = μ•1 +RT ln a1 + V1Π, (46.39)

where V1 is the partial molar volume of the solvent in solution (difficult to measure)

and Π is the hydrostatic (osmotic) pressure.

From the above equation

ln a1 =V1Π

RT(46.40)

Now we make the approximations V1 = V •m1, a1 = X1 = 1−X2:

ln(1−X2) =V •m1Π

RT(46.41)

For dilute solutions X2 is small so ln(1−X2) may be expanded as

ln(1−X2) = −X2 +X22

2− X3

2

3− · · · ' −X2, (46.42)

but X2 =n2

n1+n2' n2

n1for dilute solutions. Thus

n2n1' V •

m1Π

RT=⇒ n2 '

V •1z |

n1V•m1Π

RT, (46.43)

320

Page 333: 457454 Physical Chemistry Quantum Chemistry

or,

Π =n2V •1|z'c

RT = cRT, (46.44)

where c is the concentration of the solute.

Note the similarity of this equation with the ideal gas equation: P = cRT. Thus

the solute in a very dilute solution behaves as if it were an ideal gas.

321

Page 334: 457454 Physical Chemistry Quantum Chemistry

47. Entropy Production and

Irreverisble Thermodynamics

We have seen that thermodynamics tells us if a process will occur and kinetics

tells us how fast a process will occur.

These two areas of physical chemistry appear to be rather disjoint.

We now we consider thermodynamics of nonequilibrium states and investigate

how (and how fast) these state move towards equilibrium.

This allows us to make a stronger connection between thermodynamics and ki-

netics.

The main concept of this approach is the idea of entropy production and, ulti-

mately, entropy production per unit time–how fast we are producing entropy.

47.1. Fundamentals

We know the difference between reversible and irreversible processes from before.

However, we will state their respective definitions here in a manner best suited

for this chapter.

322

322

Page 335: 457454 Physical Chemistry Quantum Chemistry

Reversible process: dynamical equations are invariant under time inversion(t→−t).

• e.g., the one dimensional wave equation,

1

c

∂2u

∂t2=

∂2u

∂x2t→−t=⇒ 1

c

∂2u

∂(−t)2 =∂2u

∂x2=⇒ 1

c

∂2u

∂t2=

∂2u

∂x2, (47.1)

is invariant under time reversal

Irreversible process: dynamical equations are not invariant under time inver-sion (t→−t).

• e.g., the one dimensional heat equation,

1

κ

∂T

∂t=

∂2T

∂x2t→−t=⇒ 1

κ

∂T

∂(−t) =∂2T

∂x2=⇒ −1

κ

∂T

∂t=

∂2T

∂x2, (47.2)

is not invariant under time reversal.

We will be concerned with the change in entropy, dS, which can be split into two

components dS = deS + diS.

Definitions

• deS is the change in entropy due to interactions with the exterior environ-

ment.

• diS is the change in entropy due to internal changes of the system

The quantity diS is called the entropy production.

323

Page 336: 457454 Physical Chemistry Quantum Chemistry

Splitting up dS into these two parts permits an easy discussion of both open and

isolated systems–the difference between the two appearing only in deS.

General criteria for irreversibility:

• diS = 0 (reversible change)

• diS > 0 (irreversible change)

For isolated systems have diS = dS and the principle of Clausius, diS = dS ≥ 0,holds.

47.2. The Second Law

As you might expect, the second law underlies all the concepts of this chapter.

We need a “local” formulation of the second law:

• Absorption of entropy in one part of the system, compensated by a sufficientproduction in another part is prohibited

— i.e., in every macroscopic region of the system the entropy production

due to irreversible processes is positive.

This is simply another in our long list of alternative statements of the second law.

324

Page 337: 457454 Physical Chemistry Quantum Chemistry

I

II

Considering the above figure of an isolated system, we write the principle of Clau-

sius as

dS = dSI + dSII ≥ 0. (47.3)

The local formulation statement implies

diSI ≥ 0 and diS

II ≥ 0 (47.4)

and the possibility of, for example, diSI < 0 and diSII > 0 such that di¡SI + SII

¢>

0 is excluded.

47.3. Examples

The idea of entropy production can be applied to any of the processes we have

talked about; mixing, phase changes, heat flow, chemical reactions, etc. As exam-

ple we now consider the last two of these: heat flow and chemical reactions.

325

Page 338: 457454 Physical Chemistry Quantum Chemistry

47.3.1. Entropy Production due to Heat Flow

Recall from the lecture on transport phenomena that the heat flux Qf is given by

Qf = −κ4T

D(47.5)

We are now interested in exposing the time dependence, so, using Qf =q4twe get

q

4t= −κA4T

D(47.6)

in differential form this isdq

dt= −κAdT

dx. (47.7)

Example: Find the entropy production in a system consisting of two identical

connected blocks of metal (I and II), one of which is held at temperature T1 and

the other at T2 (take T1 > T > T2) where T is the temperature at the interface.

326

Page 339: 457454 Physical Chemistry Quantum Chemistry

Considering the whole system

dS =dqIT1+

dqIIT2

=

deSz | deqIT1

+deqIIT2

+

diSz | diqIT1

+diqIIT2

. (47.8)

The quantity deqj is the amount of heat supplied by the environment to hold block

j at its fixed temperature.

Furthermore the heat going out of I through the connecting wall is equal to the

heat coming into II through the connecting wall:

diqI = −diqII . (47.9)

Using this we see that the entropy production is

diS = diqI

µ1

T1− 1

T2

¶, (47.10)

which we see is positive because diqI < 0 when T1 > T2.

We have still not made a connection to kinetics.

To do so we must consider the entropy production per unit time diSdt.

327

Page 340: 457454 Physical Chemistry Quantum Chemistry

For this examplediS

dt=

diqIdt

µ1

T1− 1

T2

¶From chapter 24 we know

diqIdt

=−Aκ4T

D. (47.11)

So,diS

dt=−Aκ4T

D

µ1

T1− 1

T2

¶(47.12)

To determine T we use the fact that the heat flow out of I is equal to the heat

flow into II:diqIdt

=−diqIIdt

. (47.13)

Using the above expression for heat flow gives us T since,

−κ/ A/D/

¡T1 − T

¢= −κ/ A/

D/¡T − T2

¢⇒ T =

T1 + T22

; (47.14)

a result we might have guessed.

47.3.2. Entropy Production due to Chemical Reactions

Definitions:

1. Chemical affinity: a ≡ − (4rxnG)T,P = −P

i viμi and a ≡ − (4rxnA)T,V =

−P

i viμi

2. Extent of reaction: ξ is defined by dξ = dnivi, where ni is the number of moles

of the ith component and vi the stoichiometric factor of the ith component.

328

Page 341: 457454 Physical Chemistry Quantum Chemistry

• e.g., for the reaction N2 + 3H2 → 2NH3

dξ =dnN2(−1) =

dnH2(−3) =

dnNH3(2)

(47.15)

and

a = 2μNH3 − μN2 − 3μH2 (47.16)

The connection to kinetics: reaction rate v = dξdt

The connection to thermodynamics:

(dA)T,V =Xi

μidni =Xi

viμi| z −a

µ1

vi

¶dni| z

= −adξ (47.17)

but

(dA)T,V =

dqz | (dU)T,V − TdS ⇒ dS =

dq

T−

−adξz | (dA)T,V

T(47.18)

so

dS =

deSz|dq

T+

diSz|adξ

T(47.19)

The entropy production per unit time for a chemical reaction is a function of both

the chemical affinity and of the reaction rate

diS

dt=a

T

dt=a

Tv ≥ 0 (47.20)

We see that for a spontaneous process the entropy production per unit time is

positive. This is because a = − (4rxnA)T,V is positive as is v.

329

Page 342: 457454 Physical Chemistry Quantum Chemistry

Simultaneous Reactions

For N simultaneous chemical reactions, the entropy production per unit time

generalizes todiS

dt=1

T

NXj=1

ajvj ≥ 0. (47.21)

The second law requires that the total entropy production for simultaneous reac-

tions is positive. It says nothing about the entropy production of the individual

component reactions other then the sum of all the component entropy productions

must be positive.

For example in a system of two coupled reactions we could have a1v1 < 0, a2v2 > 0

such that a1v1 + a2v2 > 0.

47.4. Thermodynamic Coupling

Processes may be what is called thermodynamically coupled such that a process

that normally is not thermodynamically favored can be coupled to another process

that is thermodynamically favored so as to allow for the unfavorable process to

proceed spontaneously.

We just saw an example of such a situation with the discussion of simultaneous

reactions.

Thermodynamic coupling need not be confined to coupling between the same

types of processes.

That is, diffusion is the flux of matter down a concentration gradient. The so-

called Soret effect is flux of matter down a temperature gradient. Conversely, the

so-called Dufour effect is heat flux down a concentration gradient

330

Page 343: 457454 Physical Chemistry Quantum Chemistry

The following table lists a number of thermodynamically coupled phenomena

Flux q m material Q (charge)

Gradient

T Thermoconductivity Thermomechanicaleffect Soret effect Seebeck effect

P Mechanocaloriceffect

Hydrodynamicflow

Reverseosmosis Potential of flow

C Dufour effect Osmosis Diffusion Nernst Potential

ε Peltier effect Electrophoresis Migration Electoconductivity

47.5. Echo Phenonmena

Consider an ensemble that is perturbed away from thermal equilibrium by some

means such as by applying a field.

If the perturbation is released the system will begin to evolve in time as it heads

back towards the thermalized equilibrium state.

The ensemble evolves in two ways

• Reversibly

— A second perturbation can “undo” or reverse the evolution.

• Irreversibly

— The evolution towards equilibrium cannot be undone–it is irreversible

Example: The spin echo in pulsed NMR

• A radio frequency pulse prepares an ensemble of nuclear spins such that

they are all spinning coherently.

331

Page 344: 457454 Physical Chemistry Quantum Chemistry

• A strong signal is seen because all the spinning nuclei cooperate.

• Each nucleus is in a slightly different environment so each spin frequency isslightly different.

• The different environment (spin frequencies) cause the ensemble spinningnuclei to dephase

• Dephasing causes a decrease in the observed signal because now not all nucleiare cooperating.

• Now a radio pulse with the opposite phase is applied to make the nuclei spinin the opposite direction

• This undoes or reverses the dephasing process and the signal regains strength

• The full signal is not recovered however since all the while random ther-

malization is taking place to irreversibly destroy the coherence among the

nuclei.

• This cannot be undone with the second radio pulse.

332

Page 345: 457454 Physical Chemistry Quantum Chemistry

Key Equations for Exam 4

Listed here are some of the key equations for Exam 4. This section should not

substitute for your studying of the rest of this material.

The equations listed here are out of context and it would help you very little to

memorize this section without understanding the context of these equations.

The equations are collected here simply for handy reference for you while working

the problem sets.

Equations

• The Clapeyron Equation is

dP

dT=4φHm

T4φVm. (47.22)

• The Clausius-Clapeyron equation is

d(lnP )

d(1/T )= −4φHm

R(47.23)

• Fick’s first law of diffusion is

1

A

dn

dt= −DdC

dx(47.24)

333

333

Page 346: 457454 Physical Chemistry Quantum Chemistry

• Fick’s second law of diffusion:∂C

∂t= D

∂2C

∂x2. (47.25)

• Relation between the viscosity and the diffusion constant:

D =kT

f

f=6πηr=

kT

6πηr. (47.26)

• Fourier’s law of heat conduction is1

A

dq

dt= Qf = −κ

dT

dx. (47.27)

• Mixing

4mix = properties of soln −X

properties of pure. (47.28)

• Chemical potentialμ = μª +RT ln a (47.29)

• Raoult’s Law:Pi = XiP

•i (47.30)

• Raoult’s law reference

a(RL)i = γ

(RL)i Xi, γ

(RL)i → 1 as Xi → 1 (47.31)

• Henry’s Law:Pi = kXiXi. (47.32)

where kXiis the Henry’s law constant,

kXi = limXi→0

µPi

Xi

¶. (47.33)

• Henry’s law reference

a(HL)i = γ

(HL)i Xi, γ

(HL)i → 1 as Xi → 0. (47.34)

334

Page 347: 457454 Physical Chemistry Quantum Chemistry

Index

absorption spectroscopy 241activity 146, 311

mathematical definition of 146activity coefficient 146, 312adiabatic expansion 280

and heat capacity 280adiabatic wall 120angular momentum

addition of 202classical 192eigenfunctions for 199, 219jj coupling 202LS coupling 202quantum numbers 199, 219spin 201

angular momentum quantum num-ber 52

antibonding orbital 71Arrhenious activation energy 261Arrhenious equation 261, 291

temperature corrected 262atomic orbitals 49

chemists picture 50physicists picture 50

aufbau principle 58average value theorem 29Berthelot gas 13, 270binominal coefficient 90blue sky 81Bohr model 18

Bohr radius 19Boltzmann distribution 10, 96, 131Boltzmann’s equation 90, 97, 124,

131bond order 77bonding orbital 71Born model 170

corrections to 175enthalpy of solvation 174entropy of solvation 174free energy of solvation 173, 178partition coefficient 174

Born—Oppenheimer approximation62, 99, 235, 240

and the Franck—Condon princi-ple 243

bosons 56Boyle temperature 272chain rule

for partial derivatives 107character table

for the C2v group 225chemical affinity 328chemical potential 144

for a salt 161relation to activity 148relation to Gibbs free energy

145relation to Helmhotz free en-

ergy 145

335

335

Page 348: 457454 Physical Chemistry Quantum Chemistry

Clapeyron equation 298, 300, 333Clausius-Clapeyron equation 299,

333coefficient of thermal expansion 274coexistence curve 293colligative properties 318commutator 30, 189completeness 191complimentary variables 30compressibility factor

at the critical point 295compressibilty factor 270, 291configuration 90confluent hypergeometric functions

65correspondence principle 41critical point 300cyclic rule 14, 108cylindrical symmetry 69Debye—Huckel limiting law 164, 178Debye—Huckel theory 163Debye—Huckel—Guggenheim equation

164Debye’s law 129, 133degeneracy 186

of the ensemble 98diathermic wall 120diatomic molecules

electron-electron potential en-ergy operator for 61

electronic kinetric energy oper-ator for 61

electronic wavefunction for 62Hamiltonian for 61nuclear kinetic energy operator

for 61nuclear-electron potential energy

operator for 61

nuclear-nuclear potential energyoperator for 61

Schrodinger equation for 62Dieterici gas 270diffusion 301diffusion constant 302eigenfunction 5eigenvalue 5eigenvalue equation 190electric dipole approximation 79, 231electrolytes

strong 161electrophoretic effect 167elementary reactions 255

and stoichiometry 256molecularity 256

emission spectroscopy 241enemble 89ensemble average 103, 132enthalpy 136entropy 105

change for changes in temper-ature 286

change for isothermal expansion286

change for mixing 287of real gases 288

entropy production 322, 323due to chemical reactions 328due to heat flow 326

equation of state 116for a Berthelot gas 118for a Dieterici gas 118for a Redlich—Kwang gas 118for a van der Waals gas 117for an ideal gas 116for gases 269

equilibrium constant 135

336

Page 349: 457454 Physical Chemistry Quantum Chemistry

equlibrium constant 153Euler’s identity 4expansion

of gases 111reversible 114

extent of reaction 328Eyring’s equation 265, 291fermions 56Fick’s first law 302, 333Fick’s second law 302, 334first law of thermodynamics 121,

133flipping coins 90fluctuation 92fluorescence 242

stokes shift 242Fourier’s law of heat conduction 306,

334Franck—Condon integral 243Franck—Condon principle 243free energy

Gibbs 138Helmholtz 137

fugacity 147fundamental transistions 66general equlibrium 151generalized displacement 110generalized force 110gerade 69Gibb’s free energy 106Gibbs-Duhem equation 163good theory 16group

mathematical definition of 222multiplication table 223

group theory 221Hamiltonian operator 27Hamitonian

classical 27harmonic oscillator 38

energy levels for 40, 44, 86potential energy 39Schrodinger equation for 39

heat 109sign convention 110

heat capacity 115, 133Heisenberg uncertainty principle 30

and the harmonic oscillator 41helium 55

electron-electron repulsion term55

Hamiltonian 55Helmholtz free energy 106Henry’s law 311, 316, 334Henry’s law constant 316, 334Hermite polynominals 40hot bands 66Hund’s rule 205hydrogen atom

ioniztion energy of 19hydrogen molecule 74hydrogenic systems 46

energy levels for 49, 86Hamiltonian 47normalization constant 49, 85potential energy for 47Schrodinger equation for 47wavefunction (no spin) 49wavefunction (with spin) 52

ideal solutionRaoult’s law 314

immiscible solutions 153infrared spectroscopy 66internal energy 103, 121intramolecular vibrational relaxation

(IVR) 242

337

Page 350: 457454 Physical Chemistry Quantum Chemistry

inversion symmetry 69operator 69

ion mobility 166and current 168

ion transfer 174IR spectroscopy 231

and the character table 232isothermal compressibility 274isothermal expansion 279Joule expansion 282Joule-Thomson expansion 283kinetic theory of gases 250Lagrange multipliers 95Laguerre polynominals 49laminar flow 304law of corresponding states 296law of rectilinear diameters 293Legendra polynomials 200linear combinations of atomic or-

bitals (LCAO) 72Lorenz number 307many electron atom

Hamlitonian for 59maximal work 113Maxwell relations 140Maxwell’s distribution of speeds 252,

290mean free path 253, 290mean ionic activity 162mean ionic activity coefficient 162method of initial velocities 259method of isolation 259microstate 90Mie scattering 84mirror plane symmetry 70molar heat capacity 115molecular collisions

simple model for 252

molecular hydrogen ion 67Hamiltonian for 67

molecular orbital diagram 76molecular orbitals 68molecular rotations 235

asymmetric tops 239centrifugal stretching 236linear tops 238polyatomic molecules 237spherical tops 239symmetric tops 238vibrational state dependence of

236molecular vibrations 228molecule

Scrodinger equation for 78momentum operator 5Morse oscillator 64

energy levels for 65, 86Schrodinger equation for 65wavefunction for 65

Morse potential 64, 86, 240force constant associated with

9Taylor series expansion of 8

normal modes 229operator

Hermitian 189ladder 195linear 189symmetry 222

operator algebra 187orientation quantum number 53orthogonality 191overtone transitions 66parameters

extensive 109intensive 109

338

Page 351: 457454 Physical Chemistry Quantum Chemistry

particle in a box 31, 181energy levels 183energy levels for 34, 44, 218features of the energy levels 35normalization constant for 33potenial energy 31Schrodinger equation for 32three dimensional 183three dimensional energy levels

185three dimensional wavefunction

185wavefunction for 183wavefunctions for 34, 44, 218

particle on a ring 194boundary conditions 194energy levels for 195, 218Hamitonian for 194wavefunctions for 195, 218

partition coefficient 154and drug delivery 155for the Born model 174

partition functioncanonical 96, 131electronic 101grand canonical 97isothermal—isobaric 97microcanonical 96molecular 100rotational 101, 132translational 101, 132vibrational 101, 132

Pauli exclusion principle 56consquences of 58

perturbation theory 207example of the quartic oscilla-

tor 208phase diagram 300

Poiseuille’s formula 304polarizability 79postulate I (of quantum mechan-

ics) 22postulate II (of quantum mechan-

ics) 24postulate III (of quantum mechan-

ics) 25pressure 104principle of Clausius 125, 324principle quantum number 52probability amplitude 22probability distribution 22PV work 111, 133Raman scattering 80Raman spectroscopy 66, 233

and the character table 234Raoult’s law 311, 312, 314, 334

deviations from 315reference state 315

rate law 255rate laws 254

determination of 258integrated 259

Rayleigh scattering 80Rayleigh scattering law 81, 82, 87reaction velocity 255, 291reciprocal rule 108red sunsets 82Redlich-Kwang gas 270reference states 147relationship between CP and CV

139, 276relaxation effects 167rigid rotor 200

degeneracy of 235, 248energy 235, 247

rotational energy levels 200, 219

339

Page 352: 457454 Physical Chemistry Quantum Chemistry

degeneracy of 200rotational Hamiltonian 200rule of mutual exclusion 234Rydberg constant 20SATP 120Schrodinger equation

time dependent 214time independent 27

second law“local” formulation 324

second law of thermodynamics 126,133

statements of 127simple collision theory 262Slater determinant 58

for lithium 59solar system model 17solvation 169solvophobic effect 176specific heat 115spherical harmonic functions 48, 200spin 201

quantum number 51, 53wavefunction 51

spin orientationquantum number 51, 53

spin-orbitcoupling 205Hamiltonian 205interaction energy 205

spontaneous process 142state function 121

table of important ones 136Sterlings approximation 92Stoke’s law 167, 304STP 120superposition 191systems

types of 108temperature 115term symbols 204thermal conductivity 301

of gases 306of liquids 306

thermal equilibrium 120third law of thermodynamics 128,

133tips for solving problems 2total derivative 107transfer matrix 11triple point 300two level system 211

‘left’ and ‘right’ states 213, 219Hamiltonian for 212

Tyndall scattering 84ungerade 69van der Waals equation

340