Вы находитесь на странице: 1из 352

Chemistry 351 and 352

Physical Chemistry I and II


Darin J. Ulness
Fall 2006 2007
Contents
I Basic Quantum Mechanics 15
1 Quantum Theory 16
1.1 The Fall of Classical Physics . . . . . . . . . . . . . . . . . . . . 16
1.2 Bohrs Atomic Theory . . . . . . . . . . . . . . . . . . . . . . . . 17
1.2.1 First Attempts at the Structure of the Atom . . . . . . . . 17
2 The Postulates of Quantum Mechanics 22
2.1 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 27
3.1 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 31
4.1 The 1D Particle in a Box Problem . . . . . . . . . . . . . . . . . . 31
4.2 Implications of the Particle in a Box problem . . . . . . . . . . . 34
5 The Harmonic Oscillator 38
5.1 Interesting Aspects of the Quantum Harmonic Oscillator . . . . . 40
i
5.2 Spectroscopy (An Introduction) . . . . . . . . . . . . . . . . . . . 42
II Quantum Mechanics of Atoms and Molecules 45
6 Hydrogenic Systems 46
6.1 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 55
7.1 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 60
8.1 Molecular Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
8.1.1 The Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . 61
8.1.2 The BornOppenheimer 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 67
9.1 Molecular Orbital Theory . . . . . . . . . . . . . . . . . . . . . . 67
9.2 Symmetry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
10 Molecular Orbital Diagrams 72
10.1 LCAOLinear Combinations of Atomic Orbitals . . . . . . . . . 72
10.1.1 Classication of Molecular Orbitals . . . . . . . . . . . . . 73
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 ScatteringWhy the Sky is Blue 79
11.1 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 89
12.1 Statistics and Entropy . . . . . . . . . . . . . . . . . . . . . . . . 89
12.1.1 Combinations and Permutations . . . . . . . . . . . . . . . 90
12.2 Fluctuations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92
13 The Boltzmann Distribution 94
13.1 Partition Functions . . . . . . . . . . . . . . . . . . . . . . . . . . 96
13.1.1 Relation between the Q and \ . . . . . . . . . . . . . . . 97
13.2 The Molecular Partition Function . . . . . . . . . . . . . . . . . . 99
14 Statistical Thermodynamics 103
15 Work 107
15.1 Properties of Partial Derivatives . . . . . . . . . . . . . . . . . . . 107
15.1.1 Summary of Relations . . . . . . . . . . . . . . . . . . . . 107
15.2 Denitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 108
15.2.1 Types of Systems . . . . . . . . . . . . . . . . . . . . . . . 108
15.2.2 System Parameters . . . . . . . . . . . . . . . . . . . . . . 109
15.3 Work and Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
15.3.1 Generalized Forces and Displacements . . . . . . . . . . . 110
15.3.2 1\ work . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
16 Maximum Work and Reversible changes 113
16.1 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 119
17.1 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 124
18.1 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 Debyes Law . . . . . . . . . . . . . . . . . . . . . . . . . . 129
18.3 Times Arrow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130
IV Basics of Thermodynamics 134
19 Auxillary Functions and Maxwell Relations 135
19.1 The Other Important State Functions of Thermodynamics . . . . 135
19.2 Enthalpy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 136
19.2.1 Heuristic denition: . . . . . . . . . . . . . . . . . . . . . . 137
19.3 Helmholtz Free Energy . . . . . . . . . . . . . . . . . . . . . . . . 137
19.3.1 Heuristic denition: . . . . . . . . . . . . . . . . . . . . . . 138
19.4 Gibbs Free Energy . . . . . . . . . . . . . . . . . . . . . . . . . . 138
19.4.1 Heuristic denition: . . . . . . . . . . . . . . . . . . . . . . 139
19.5 Heat Capacity of Gases . . . . . . . . . . . . . . . . . . . . . . . . 139
19.5.1 The Relationship Between C
1
and C
\
. . . . . . . . . . . 139
19.6 The Maxwell Relations . . . . . . . . . . . . . . . . . . . . . . . . 140
20 Chemical Potential 142
20.1 Spontaneity of processes . . . . . . . . . . . . . . . . . . . . . . . 142
20.2 Chemical potential . . . . . . . . . . . . . . . . . . . . . . . . . . 144
20.3 Activity and the Activity coe!cient . . . . . . . . . . . . . . . . . 146
20.3.1 Reference States . . . . . . . . . . . . . . . . . . . . . . . 147
20.3.2 Activity and the Chemical Potential . . . . . . . . . . . . 148
21 Equilibrium 151
21.0.3 Equilibrium constants in terms of 1
C
. . . . . . . . . . . . 153
21.0.4 The Partition Coe!cient . . . . . . . . . . . . . . . . . . . 153
22 Chemical Reactions 156
22.1 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 1
o
. . . . . . . . . . . . . . . . . . . 159
22.4 Extent of Reaction . . . . . . . . . . . . . . . . . . . . . . . . . . 160
23 Ionics 161
23.1 Ionic Activities . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
23.1.1 Ionic activity coe!cients . . . . . . . . . . . . . . . . . . . 162
23.2 Theory of Electrolytic Solutions . . . . . . . . . . . . . . . . . . . 163
23.3 Ion Mobility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164
23.3.1 Ion mobility . . . . . . . . . . . . . . . . . . . . . . . . . . 165
24 Thermodynamics of Solvation 169
24.1 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 181
26.1 Particle in a Box . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
26.2 The 3D Particle in a Box Problem . . . . . . . . . . . . . . . . . . 183
27 Operators 187
27.1 Operator Algebra . . . . . . . . . . . . . . . . . . . . . . . . . . . 187
27.2 Orthogonality, Completeness, and the Superposition Principle . . 191
28 Angular Momentum 192
28.1 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
29 Addition of Angular Momentum 201
29.1 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 207
30.1 Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . 207
30.2 Variational method . . . . . . . . . . . . . . . . . . . . . . . . . . 209
31 The Two Level System and Quantum Dynamics 211
31.1 The Two Level System . . . . . . . . . . . . . . . . . . . . . . . . 211
31.2 Quantum Dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . 214
VI Symmetry and Spectroscopy 220
32 Symmetry and Group Theory 221
32.1 Symmetry Operators . . . . . . . . . . . . . . . . . . . . . . . . . 222
32.2 Mathematical Groups . . . . . . . . . . . . . . . . . . . . . . . . . 222
32.2.1 Example: The C
2
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 228
33.1 Molecular Vibrations . . . . . . . . . . . . . . . . . . . . . . . . . 228
33.1.1 Normal Modes . . . . . . . . . . . . . . . . . . . . . . . . 229
33.1.2 Normal Modes and Group Theory . . . . . . . . . . . . . . 229
34 Vibrational Spectroscopy and Group Theory 231
34.1 IR Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . . . . 231
34.2 Raman Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . . . 233
35 Molecular Rotations 235
35.1 Relaxing the rigid rotor . . . . . . . . . . . . . . . . . . . . . . . . 236
35.2 Rotational Spectroscopy . . . . . . . . . . . . . . . . . . . . . . . 236
35.3 Rotation of Polyatomic Molecules . . . . . . . . . . . . . . . . . . 237
36 Electronic Spectroscopy of Molecules 240
36.1 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 FranckCondon activity . . . . . . . . . . . . . . . . . . . . . . . 243
36.2.1 The FranckCondon principle . . . . . . . . . . . . . . . . 243
37 Fourier Transforms 245
37.1 The Fourier transformation . . . . . . . . . . . . . . . . . . . . . 245
VII Kinetics and Gases 249
38 Physical Kinetics 250
38.1 kinetic theory of gases . . . . . . . . . . . . . . . . . . . . . . . . 250
38.2 Molecular Collisions . . . . . . . . . . . . . . . . . . . . . . . . . . 252
39 The Rate Laws of Chemical Kinetics 254
39.1 Rate Laws . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
39.2 Determination of Rate Laws . . . . . . . . . . . . . . . . . . . . . 258
39.2.1 Dierential methods based on the rate law . . . . . . . . . 259
39.2.2 Integrated rate laws . . . . . . . . . . . . . . . . . . . . . . 259
40 Temperature and Chemical Kinetics 261
40.1 Temperature Eects 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 269
41.1 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 Coe!cients . . . . . . . . . . . . . . . 273
42 Behavior of Gases 274
42.1 1. \ and 1 behavior . . . . . . . . . . . . . . . . . . . . . . . . . 274
42.1.1 c and i
T
for an ideal gas . . . . . . . . . . . . . . . . . . . 275
42.1.2 c and i
T
for liquids and solids . . . . . . . . . . . . . . . . 275
42.2 Heat Capacity of Gases Revisited . . . . . . . . . . . . . . . . . . 276
42.2.1 The Relationship Between C
1
and C
\
. . . . . . . . . . . 276
42.3 Expansion of Gases . . . . . . . . . . . . . . . . . . . . . . . . . . 279
42.3.1 Isothermal and Adiabatic expansions . . . . . . . . . . . . 279
42.3.2 Heat capacity C
\
for adiabatic expansions . . . . . . . . . 280
42.3.3 When 1 is the more convenient variable . . . . . . . . . . 281
42.3.4 Joule expansion . . . . . . . . . . . . . . . . . . . . . . . . 282
42.3.5 Joule-Thomson expansion . . . . . . . . . . . . . . . . . . 283
43 Entropy of Gases 286
43.1 Calculation of Entropy . . . . . . . . . . . . . . . . . . . . . . . . 286
43.1.1 Entropy of Real Gases . . . . . . . . . . . . . . . . . . . . 288
VIII More Thermodyanmics 292
44 Critical Phenomena 293
44.1 Critical Behavior of uids . . . . . . . . . . . . . . . . . . . . . . 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 1 and 1 . . . . . . . . . . . . 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 301
45.1 Diusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 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 308
46.1 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 Raoults Law . . . . . . . . . . . . . . . . . . . . . . . . . 312
46.3.3 Ideal Solutions (RL) . . . . . . . . . . . . . . . . . . . . . 314
46.3.4 Henrys Law . . . . . . . . . . . . . . . . . . . . . . . . . . 316
46.4 Colligative Properties . . . . . . . . . . . . . . . . . . . . . . . . . 318
46.4.1 Freezing Point Depression . . . . . . . . . . . . . . . . . . 318
46.4.2 Osmotic Pressure . . . . . . . . . . . . . . . . . . . . . . . 319
47 Entropy Production and Irreverisble Thermodynamics 322
47.1 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
Chemistry 351: Physical
Chemistry I
1
1
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 arent 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
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 dont have to work on an overwhelming number
of problems at a time. Try to whip-o 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 e!cient at helping you learn.
3. You can do the problem. I dont assign problems that you cannot do. If you
think you cant do the problem then maybe you need try a dierent 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 problem
is 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 gure out what mathematical techniques you
need to express the solution to the problem.
3
Do the math; either you will be able to do this or you wont. It might
take 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 is
correct. If you are not sure, start with a related concept that is better
known by you.
Look for self-consistency. Does you nal answer jive with what you
know.
Problems Dealing With Quantum Mechanics
Problem: What is the periodicity of the following functions
,(r) = sin
2
r
,(r) = cos r
,(r) = c
2ia
Solution: For the rst function it is easiest to see the periodicity by writing the
function as ,(r) = (sin r)(sinr). We know that this function will repeat zeros
when ever sin r = 0. This occurs at r = ::, : = 0. 1. 2 . . ., so the periodicity
is :. The second function we should remember from trig as having a period of 2:.
Finally for the last function it is best to used Eulers identity and write
c
2ia
= cos 2r +i sin2r (1)
The real part of this function, cos 2r. has a period of : as does the imaginary
part, sin 2r. Therefore the entire function has a period of :.
4
Problem: Which of the following functions are eigenfunction of the momentum
operator, j
a
= i~
o
oa
.
(r) = c
iIa
(r) = c
ca
2
(r) = cos /r
Solution: We need to determine if j
a
(r) = `(r) where ` is a constant. If
this equation is true then the function is an eigenfunction with eigenvalue `. 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 rst function
j
a
(r) = i~
d(r)
dr
= i~
dc
iIa
dr
= ~/c
iIa
= ~/(r). (2)
so, yes, this function is an eigenfunction of the momentum operator.
For the second function
j
a
(r) = i~
d(r)
dr
= i~
dc
ca
2
dr
= 2i~crc
ca
2
= 2i~c

r(r). (3)
so, no, this function is not and eigenfunction of the momentum operator.
For the last function
j
a
(r) = i~
d(r)
dr
= i~
d cos /r
dr
= i~/
6=cos Ia
z }| {
sin/r. (4)
so, no, this function is not an eigenfunction of the momentum operator.
Problem: A quantum object is described by the wavefunction (r) = c
ca
2
.
What is the probability of nding the object further than c away from the origin
( r = 0)?
5
Solution: First of all we do not know if this wavefunction is normalized, so we
should assume that it isnt. We could normalize this wavefunction, but we wont.
We are interested in nding the probability that the object is outside of the region
c < r < c. To do this using an unnormalized wavefunction we must evaluate
1(|r| c) =
R
c
o
|(r)|
2
dr +
R
o
c
|(r)|
2
dr
R
o
o
|(r)|
2
dr
. (5)
The rst integral in the numerator gives the probability that the object is at a
position r < c and the second integral in the numerator gives the probability
for r c. 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
1(|r| c) =
R
c
o
c
2ca
2
dr +
R
o
c
c
2ca
2
dr
R
o
o
c
2ca
2
dr
. (6)
Mathematica can assist with these integrals to give the nal answer of
1(|r| c) = erfc[
_
2c
3
2
]. (7)
Problem: A quantum object is described by the wavefunction (r) = c
a
over
the range 0 _ r < . Normalize this wavefunction.
Solution: Following our general procedure from the notes if we have some unnor-
malized wavefunction,
unnorm
we know that this function must simply be some
constant ` multiplied by the normalized version of this function:

unnorm
= `
norm
(8)
We have shown generally that ` is given by
` =
s
Z
space
|
unnorm
(r)|
2
dr. (9)
6
Which for this case is
` =
s
Z
o
0
|c
a
|
2
dr =
s
Z
o
0
c
2a
dr =
r
1
2
(10)
So nally we get the normalized wavefunction by rearanging
unnorm
= `
norm
:

norm
(r) =
p
2c
a
. (11)
Problem: A quantum object is described by the wavefunction (r) = c
a
over
the range 0 _ r < . What is the average position of the object?
Solution: We need to work with the normalized wavefunction that we found in
the previous problem, (r) =
_
2c
a
. Generally and average is calculated as
h oi =
Z
space

+
(r) o(r). (12)
which in this case is
h ri =
Z
o
0
p
2c
a
r
p
2c
a
dr = 2
Z
o
0
rc
2a
dr =
1
2
. (13)
So on average you will nd the object at r =
1
2
.
Problem: What is the probability of nding an electron in the 1s state of hydrogen
further than one Bohr radius away from the nucleus?
Solution: We need to evaluate
1(: c
0
) =
Z
2
0
Z

0
Z
o
o
0
|
1c
|
2
:
2
sinod:dodc. (14)
Remember the extra :
2
sin o is needed when integrating in spherical polar coordi-
nates. The normalized 1: wavefunction is

1c
=
1
p
:c
3
0
c
vo
0
. (15)
7
We can do this integral by hand or have Mathematica help us to give
1(: c
0
) =
5
c
2
= 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 =
~
2
2n
\
2
. Express the wavefunction (in Cartesian coordinates) as a product
state.
Solution: This problem appears hard at rst 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, [(r. . .) 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 =
~
2
2:

J
2
Jr
2
+
J
2
J
2
+
J
2
J.
2

. (17)
We see that indeed the Hamiltonian is a sum of term that depends only on r,
a term depending only on and a term that depends only on .. Therefore the
appropriate product state is
[(r. . .) = (r)()(.). (18)
Problem: Expand the Morse potential in a Taylors series about 1
cq
. Verify that
the coe!cient for the linear term is zero. What is the force constant associated
with the Morse potential?
Solution: The Morse potential is
\ (r) = 1
c

1 c
o(11
cq
)

. (19)
8
The Taylor series about 1
cq
for this function is
\ (r) = \ (r)|
1
cq
| {z }
= 0
+
d\ (r)
dr

1
cq
| {z }
= 0
(11
cq
) +
1
2!
d
2
\ (r)
dr
2

1
cq
| {z }
= o
2
1c
(1 1
cq
)
2
+ . (20)
So, yes the coe!cient of the linear term (the term involving (1 1
cq
) to the
rst 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
coe!cient of the quadratic term so in this case / = ,
2
1
c
.
Problem: Without performing any calculations, compare h1i as a function of
the 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 h1i = 1
cq
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 h1i increases as the
quantum number increases.
Problems Dealing With Statistical Mechanics and Thermo-
dynamics
Problem: A vial containing 10
20
benzene molecules is at 300K. How many mole-
cules are in the rst excited state of the ring breathing mode (992 cm
1
)? How
9
many are in the rst excited state of the symmetric CH vibrational mode (3063
cm
1
)?
Solution: This is a problem that deals with the Boltzmann distribution. So,
`
vb
=1
=

2 sinh
992
2 208

3992
2208

10
20
= 8.41 10
17
(21)
and
`
CH
=1
=

2 sinh
3063
2 208

33063
2208

10
20
= 4.02 10
13
. (22)
We see that about
8.4110
17
10
20
100% = 0.841% of the benzene molecules are in the
rst vibrational excited state for the ring breathing mode and
4.0210
13
10
20
100% =
0.0000402% of the benzene molecules are in the rst excited state for the CH
stretching mode.
Problem: Consider a linear chain of ` atoms. Each of the atoms can be in one
of three states . 1 or C, except that an atom in state can not be adjacent to
an atom in state C. Find the entropy per atom for this system as ` . To
solve this problem it is useful to dene the set of three dimensional column vectors
\
())
such that the three elements are the total number of allowed congurations of
a ,-atom chain having the ,
th
atom in state , 1 or C. For example,
\
(1)
=
5
9
7
1
1
1
6
:
8
. \
(2)
=
5
9
7
2
3
2
6
:
8
. \
(3)
=
5
9
7
5
7
5
6
:
8
. . (23)
The \
()+1)
can be found from the \
())
vector using the matrix equation,
\
()+1)
= `\
())
. (24)
where for this example
` =
5
9
7
1 1 0
1 1 1
0 1 1
6
:
8
. (25)
10
The matrix ` is the so-called transfer matrix for this system. It can be shown
that the number of congurations \ = Tr[`
.
]. Now for large `. Tr[`
.
] - `
.
max
,
where `
max
is the largest eigenvalue of `. So
\ = lim
.o
`
.
max
. (26)
1. 1. Use ` to nd \
(4)
2. Verify \
(3)
explicitly by drawing all the allowed 3-atom congurations.
3. Verify \ = Tr[`
.
] for ` = 1 and ` = 2.
4. Use Boltzmanns equation to nd the entropy per atom for this chain
as ` goes to innity.
Solution: For part (a) we simply use the transfer matrix as directed in the
problem (we are given \
(3)
):
\
(4)
=
5
9
7
1 1 0
1 1 1
0 1 1
6
:
8
5
9
7
5
7
5
6
:
8
=
5
9
7
12
17
12
6
:
8
.
For part (b) we need to list all states for the case of ` = 3 and verify the we get
the same result as calculated using the transfer matrix. Remembering that \
(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 States ending in 1 States ending in C
1 1C
1 11 11C
1 11 1CC
11 111 C1C
C1 1C1
C11
CC1
5 states
_
7 states
_
5 states
_
.
11
States like C are not allowed because and C are neighbors.
For part (c) we evaluate \ = Tr[`
.
] for ` = 1 and 2. For ` = 1. \ =
Tr[`] = 3 This corresponds to the three distinguishable microstates , 1. and
C. For ` = 2.
\ = Tr[`
2
] = Tr
5
9
7
5
9
7
1 1 0
1 1 1
0 1 1
6
:
8
5
9
7
1 1 0
1 1 1
0 1 1
6
:
8
6
:
8
= Tr
5
9
7
5
9
7
2 2 1
2 3 2
1 2 2
6
:
8
6
:
8
= 7 (27)
This corresponds to the seven distinguishable microstates . 1. 1. 11. 1C.
C1 and CC (Remember C and cannot be neighbors).
For part (d) we use
o
`
=
/
`
ln\ = lim
.o
/
`
ln `
.
max
= lim
.o
/
`
` ln `
max
= / ln `
max
. (28)
So, we simply need to nd the maximum eigenvalue of the Transfer matrix. Using
Mathematica we nd `
max
= 1 +
_
2. Therefore the limiting entropy per atom
is
o
`
= / ln

1 +
_
2

. (29)
Problem: Using the classical theory of light scattering, calculate the positions of
the 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 500nmgreen light).
Solution: Since there are two vibrational modes we expect two Stokes lines to
the 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
Problem: A simple model for a crystal is a gas of harmonic oscillators. De-
termine , o, and l from the partition function for this model.
Solution: For this model the crystal is modelled as a collection of harmonic
oscillators so we need the partition function for the harmonic oscillator.
Q
crystal
=
.
1O
=

1
2 sinh
o~.
2
!
(30)
From our formulas for statistical thermodynamics
= /1 ln Q
crystal
= +`11 ln

2 sinh
,~.
2

. (31)
where we used properties of logs to pull the ` out front and move the sinh term
from to the numerator,
o = /,
JQ
crystal
J,
+/ lnQ
crystal
(32)
=
`/,~.
2
coth
,~.
2
/ ln

2 sinh
,~.
2

and
l =
JQ
crystal
J,
=
`~.
2
coth
,~.
2
. (33)
Problem: Express the equation of state for internal energy for a Berthelot gas.
Solution: The equation representing a Berthelot gas is
1 =
:11
\ :/

:
2
c
1\
2
. (34)
We are interesting in an equation of state for l(1. \ ). Writing out the total
derivative of l(1. \ ) we get
dl =

Jl
J1

\
d1 +

Jl
J\

T
d\. (35)
13
Now

0l
0T

\
is just heat capacity, C
\
, but

0l
0\

T
is nothing convenient so we must
proceed. We employ the useful relation

Jl
J\

T
= 1

J1
J1

\
1 (36)
to eliminate l in favor of 1 so that we can use the equation of state for a Berthelot
gas. One obtains
1

J1
J1

\
1 = 1

:1
\ :/
+
:
2
c
1
2
\
2

:11
\ :/
+
:
2
c
1\
2
=
2:
2
c
1\
2
. (37)
Hence the equation of state for internal energy of a Berthelot gas is
dl = C
\
d1 +
2:
2
c
1\
2
d\ (38)
Problem: Use the identities for partial derivatives to eliminate the

01
0T

\
factor
in
C
j
= C

+1

J\
J1

J1
J1

\
(39)
so that all derivatives are at constant pressure or temperature.
Solution: Here we either remember an identity or turn to our handout of partial
derivative identities to employ the cyclic rule to

01
0T

\
:

J1
J1

\
=

J1
J\

J\
J1

1
. (40)
This eliminates the constant \ term and so,
C
j
= C

1

J\
J1

2
1

J1
J\

T
. (41)
14
Part I
Basic Quantum Mechanics
15
15
1. Quantum Theory
The goal of science is unication.
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 deed explanation by means of the current understanding of physics. Among
these experiments were
1. The photoelectric eect
2. Low temperature heat capacity
3. Atomic spectral lines
4. Black body radiation and the ultraviolet catastrophe
16
16
5. The two slit experiment
6. The Stern-Gerlach experiment
+ + See Handouts + +
1.2. Bohrs 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 balanced
by the repulsive centrifugal force.
Flaws of the solar system model
Newton: OK
_
Maxwell: problem
_
17
As the electron orbits the nucleus, the atom acts as an oscillating dipole
The classical theory of electromagnetism states that oscillating dipoles
emit radiation and thereby lose energy.
The system is not stable and the electron spirals into the nucleus. The
atom collapses!
Bohrs model: Niels Bohr (18851962)
18
Atoms dont 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.
: =
4:c
0
`
2
~
2
2:
c
c
2
(1.1)
where 2 is the atomic number, :
c
and c are the mass and charge of the
electron respectively and c
0
is the permittivity of free space. ` is a positive
real integer called the quantum number. ~ = /,2: is Plancks constant
divided by 2:.
The constant quantity
4c
0
~
2
ncc
2
appears often and is given the special symbol c
0
=
4c
0
~
2
ncc
2
= 0.52918 and is called the Bohr radius.
The total energy of the Bohr atom is related to its quantum number
1
.
= 2
2

c
2
2c
0

1
`
2
. (1.2)
Tests of the Bohr atom
Ionization energy of Hydrogen atoms
The Ionization energy for Hydrogen atoms (2 = 1) is the minium
energy required to completely remove an electron form it ground state,
i.e., ` = 1 ` =
1
ionize
= 1
o
1
1
=
2
2
c
2
2c
0

2

1
1
2

=
c
2
2c
0
(1.3)
19
1
ionize
=
c
2
2o
0
= 13.606 eV= 109,667 cm
1
=R. Ris called the Rydberg
constant.
1
ionize
experimentally observed from spectroscopy is 13.605 eV (very
good agreement)
Spectroscopic lines fromHydrogen represent the dierence in energy between
the quantum states
Bohr theory: Dierence energies
1
)
1
I
=
c
2
2c
0

1
`
2
)

1
`
2
I

= R

1
`
2
)

1
`
2
I

(1.4)
Initial state `
I
Final States `
)
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
=
41
/c
(1.5)
Failure of the Bohr model
No ne structure predicted (electron-electron coupling)
No hyperne structure predicted (electron-nucleus coupling)
No Zeeman eect predicted (response of spectrum to magnetic eld)
20
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 Schrdinger (18871961): Wave mechanics
Werner Heisenberg (19021976): Matrix mechanics
Paul Dirac (19021984): Abstract vector space approach
21
2. The Postulates of Quantum
Mechanics
2.1. Postulate I
Postulate I: The state of a system is dened by a wavefunction, . which con-
tains all the information that can be known about the system.
We will normally take to be a complex valued function of time and coordi-
nates: (t. r. . .) and, in fact, we will most often deal with time independent
stationary states (r. . .)
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 represents a probability amplitude and is not directly observ-
able.
However the mod-square of the wavefunction,
+
= ||
2
. represents a probability
distribution which is directly observable.
That is, the probability of nding a particle which is described by (r. . .) at the
position between r and r+dr. and +d and . and .+d. is |(r. . .)|
2
drdd.
(or |(:. o. c)|
2
:
2
sin od:dodc in spherical coordinates).
22
22
Properties of the wavefunction
Single valueness
continuous and nite
continuous and nite rst derivative

R
space
|(r. . .)|
2
drdd. <
Normalization of the wavefunction
In order for |(r. . .)|
2
to be exactly interpreted as a probability dis-
tribution, (r. . .) must be normalizable.
That is,
unnorm
= `
norm
, where ` =
q
R
space
|
unnorm
(r. . .)|
2
drdd.
This assures that
R
space
|
norm
|
2
drdd. = 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,
unnorm
we know that this function
must simply be some constant ` multiplied by the normalized version of this
function:

unnorm
= `
norm
. (2.1)
Now, we take the mod-square of both sides and then integrate both sides of this
equation over all space
Z
space
|
unnorm
|
2
drdd. =
Z
space
|`
norm
|
2
drdd.. (2.2)
23
The ` is just a constant so it can be pulled out of both the mod-square and the
integral
Z
space
|
unnorm
|
2
drdd. = `
2
Z
space
|
norm
|
2
drdd.. (2.3)
but
Z
space
|
norm
|
2
drdd. = 1 (2.4)
because that is the very denition of a normalized wavefunction. Thus wherever
we see
R
space
|
norm
|
2
drdd. we can replace it with 1. So,
Z
space
|
unnorm
|
2
drdd. = `
2
1 = `
2
. (2.5)
This gives us an expression for `. Taking the square root of both sides gives.
` =
s
Z
space
|
unnorm
(r. . .)|
2
drdd.. (2.6)
So nally we get the normalized wavefunction by reagranging
unnorm
= `
norm
:

norm
=
1
`

unnorm
. (2.7)
Notice that no where did we ever specify what
unnorm
or
norm
actually were,
therefore this is a general procedure that will work for any wavefunction.
To nd the probability for the particle to be in a nite region of space we simple
evaluate (here a 1D case)
1(r
1
< r < r
2
) =
R
a
2
a
1
|(r)|
2
dr
R
o
o
|(r)|
2
dr
if (a)
==
normalized
Z
a
2
a
1
|(r)|
2
dr (2.8)
2.3. Postulates II and II
Postulate II: Every physical observable is represented by a linear (Hermitian)
operator.
24
An operator takes a function and turns it into another function

C,(r) = q(r) (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
r = r
j
a
= i~
0
0a
and of course the analogous operators for the other coordinates (. .) 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 that
is one of the eigenvalues of the corresponding operator.
There is a special operator equation called the eigenvalue equation which is

C,(r) = `,(r) (2.10)


where ` is just a number.
For a given operator only a special set of function satisfy this equation. These
functions are called eigenfunctions.
25
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

C in the eignevalue equation be the operator that takes the derivative:

C =

d =
o
oa
.
So we want a solution to

d,(r) = `,(r) (2.11)


d,(r)
dr
= `,(r)
So, we ask ourselves what function is proportional to its own derivative? =
,(r) = c
Aa
.
So the eigenfunctions are the set of functions ,(r) = c
Aa
and the eigenvalues are
the numbers `
26
3. The Setup of a Quantum
Mechanical Problem
3.1. The Hamiltonian
The most important physical observable is that of the total energy 1.
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

H = 1. (3.1)
This equation is the (time independent) Schrdinger 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. Dene the classical Hamiltonian for the system.
27
27
The total energy for a classical system is
1
c|
= 1 +\. (3.2)
where 1 is the kinetic energy and \ is the potential energy.
The kinetic energy is always of the form
1 =
1
2:

j
2
a
+j
2
j
+j
2
:

(3.3)
The potential energy is almost always a function of coordinates only
\ = \ (r. . .) (3.4)
Note: Some quantum systems dont have classical analogs so the Hamil-
tonian operator must be hypothesized.
2. Use Postulate II to replace the classical variables, r. j
a
etc., with their
appropriate operators. Thus,

1 =
~
2
2:

\
2
=
~
2
2:
\
2
. (3.5)
where \
2
=
0
2
0a
2
+
0
2
0j
2
+
0
2
0:
2
. and

\ = \ ( r. . .) = \ (r. . .). (3.6)


So,

H =

1 +

\ =
~
2
2:
\
2
+\ (r. . .) (3.7)
3. Solve the Schrdinger equation,

H = 1. which is now a second order
dierential equation of the form

~
2
2:
\
2
+\ (r. . .)

= 1
=
~
2
2:
\
2
+ (\ (r. . .) 1) = 0 (3.8)
28
Note: It is solely the form of \ (r. . .) which determines whether this
is easy or hard to do.
For one-dimensional problems
~
2
2:
d
2
dr
2
+ (\ (r) 1) = 0 (3.9)
3.3. The Average Value Theorem
Postulate III implies that if 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 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
h ci =
Z
space

+
cdrdd. (3.10)
For example,
h ri =
Z
space

+
rdrdd. =
Z
space
r||
2
drdd. (3.11)
and
h j
a
i =
Z
space

+
j
a
drdd. = i~
Z
space

+
J
Jr
drdd. (3.12)
29
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
oco, _
1
2

Dh
c.

,
iE

. (3.13)
where the notation
h
c.

,
i
means the commutator of c and

,. The commutator is
dened as
h
c.

,
i
= c

, c. (3.14)
The most important example of complimentary variables is position and momen-
tum. We see
oj
a
or _
1
2
|h[ j
a
. r]i| =
1
2
|h j
a
r r j
a
i| (3.15)
=
1
2

Z

+
~
i

J
Jr
r r
J
Jr

dr

~
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., oj
a
o = 0. we can
know, for example, the position and the r momentum to arbitrary precision.)
Suppose we know the position of a particle perfectly, what can we say about its
momentum?
30
4. Particle in a Box
We now will apply the general program for solving a quantum mechanical problem
to our rst 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, \ (r). shown in the gure and given by
\ (r) =
;
A
?
A
=
r _ 0
0 0 < r < c
r _ c
. (4.1)
Because of the innities at r = 0 and r = c, we need to partition the r-axis into
the three regions shown in the gure.
31
31
Now, in region I and III, where the potential is innite, the particle can never
exist so, must equal zero in these regions.
The particle must be found only in region II.
The Schrdinger equation in region II is (\ (r) = 0)

H = 1 ==
~
2
2:
d
2
(r)
dr
2
= 1. (4.2)
which can be rearranged into the form
d
2
(r)
dr
2
+
2:1
~
2
(r) = 0. (4.3)
The general solution of this dierential equation is
(r) = sin/r +1cos /r. (4.4)
where / =
q
2n1
~
2
.
Now must be continuous for all r. Therefore it must satisfy the boundary
conditions (b.c.): (0) = 0 and (c) = 0.
32
From the (0) = 0 b.c. we see that the constant 1 must be zero because
cos /r|
a=0
= 1.
So we are left with (r) = sin /r for our wavefunction.
As can be inferred from the following gure, the second b.c., (c) = 0. places
certain restrictions on /.
In particular,
/
a
=
::
c
. : = 1. 2. 3. . (4.5)
The values of / are quantized. So, now we have

a
(r) = sin
::r
c
. (4.6)
The constant is the normalization constant. We obtain from
Z
o
o

+
a
(r)
a
(r) = 1 =
Z
o
0

2
sin
::r
c
sin
::r
c
dr. (4.7)
Letting n =
a
o
. dn =

o
dr. this becomes
1 =
2
c
:
Z

0
sin
2
:ndn =
2
c
: /
: /
2
=

2
c
2
. (4.8)
33
Solving for gives
=
r
2
c
. (4.9)
Thus our normalized wavefunctions for a particle in a box are

a
(r) =
;
A
A
?
A
A
=
0 I
q
2
o
sin
aa
o
II
0 III
. (4.10)
Is this wavefunction OK?
We can get the energy levels from /
a
=
q
2n1
n
~
2
and /
a
=
a
o
:
1
a
=
:
2
:
2
~
2
2:c
2
~=
I
2r
=
:
2
/
2
8:c
2
. (4.11)
4.2. Implications of the Particle in a Box problem
Zero Point Energy
34
The smallest value for : is 1 which corresponds to an energy of
1
1
=
/
2
8:c
2
6= 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,
o j = 0. But we also know that the particle is located within a nite region of
space, so o r 6= .
Hence, o ro j = 0 which violates the uncertainty principle.
Features of the Particle in a Box Energy Levels
The energy level spacing is
41 = 1
a+1
1
a
=
(: + 1)
2
/
2
8:c
2

:
2
/
2
8:c
2
= (:
2
/ + 2: + 1 :
2
/ )
/
2
8:c
2
41 = (2: + 1)
/
2
8:c
2
(4.13)
This spacing increases linearly with quantum level :
This spacing decreases with increasing mass
This spacing decreases with increasing c
It is this level spacing that is what is measured experimentally
The Curvature of the Wavefunction
35
The operator for kinetic energy is

1 =
~
2
2n
o
2
oa
2
. The important part of this is
o
2
oa
2
.
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 c.
We know the wavefunction is zero in regions I and III. We also know that
the wave function is not zero everywhere. Therefore it must do something
between r = 0 and r = c. It must have some curvature and hence some zero
point energy.
As c is increased, the wavefunction is less conned and so the curvature does
not need to be as great to satisfy the boundary conditions. Therefore the
energy levels decrease in energy as does their dierence.
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:
hri =
Z
o
o

+
2
(r)r
2
(r)dr =
2
c
Z
o
0
rsin
2
[
2:
c
r]dr =
c
2
(4.14)
36
yet the probability of nding the particle at r =
o
2
is zero:
2
(
o
2
) = 0. There is
a node at r =
o
2
. 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
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 gure is 1 = /(11
cq
), where /
is the spring constant and 1
cq
is the equilibrium position of the ball.
Setting r = 1 1
cq
we can measure the displacement about the equilibrium
position.
38
38
From Newtons law of motion 1 = :c = :
o
2
a
ot
2
. we get
:
d
2
r
dt
2
= /r =
d
2
r
dt
2
+
/
:
r = 0 (5.1)
This is second order dierential equation which we already know the solutions to:
r = sin.t +1cos .t. (5.2)
where . =
q
I
n
and and 1 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 \ . We know
\ =
Z
1dr =
1
2
/r
2
+C. (5.3)
Since energy is on an arbitrary scale we can set C = 0. Thus \ =
1
2
/r
2
.
By postulate III the Schrdinger equation becomes

H = 1 =
3
E
C
~
2
2:
d
2
dr
2
| {z }
K.E.
+
1
2
/r
2
| {z }
P.E.
4
F
D
= 1. (5.4)
This can be rearrange into the form
~
2
2:
d
2

dr
2
+

1
2
/r
2
1

= 0 (5.5)
This dierential equation is not easy to solve (you can wait to solve it in graduate
school).
39
The equation is very close to the form of a know dierential equation called Her-
mites dierential equation the solutions of which are called the Hermite polynom-
inals.
As it turns out, the solutions (the eigenfunctions) to the Schrdinger equation for
the harmonic oscillator are

a
() =
a
H
a
()c

2
2
. =

/:
~
2
1
4
r.
a
=
1
p
2
a
:!
_
:
. (5.6)
where
a
is the normalization constant for the :
th
eigenfunction and H
a
() are
the Hermite polynomials.
The eigenvalues (the energy levels) are
1
a
= (: +
1
2
)~.. (5.7)
where again . =
q
I
n
.
Note the energy levels are often written as
1
a
= (: +
1
2
)/i
0
. (5.8)
where i
0
=
1
2
q
I
n
and 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
1. Consider the ground state (the lowest energy level)
There is residual energy in the ground state because
1
0
= (0 +
1
2
)~..
Just like for the particle in a box, this energy is called the zero point
energy.
It is a consequence of uncertainty principle
If the ground state energy was really zero, then we would conclude
that the momentum of the oscillator was zero.
On the other hand, we would conclude the particle was located at
the bottom of the potential well (at r = 0)
Thus we would have oj = 0. or = 0, so ojor = 0 Not allowed!
The uncertainty principle forces there to be some residual zero
point energy.
2. Consider the wavefunctions.
The wavefunctions penetrate into the region where the classical particle
is forbidden to go
The wavefunction is nonzero past the classical turning point.
The probability distribution ||
2
becomes more and more like what is
expected for the classical oscillator when .
This is a manifestation of the correspondence principle which
states 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
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 radiationmatter interaction are called
spectroscopies.
Atoms and molecules absorb or emit light only at specic (quantized) energies.
These specic values correspond to the energy level dierence between the initial
and nal states.
42
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).
` =
s
Z
space
|
unnorm
(r)|
2
dr. (5.9)
The normalized wavefunction:

norm
=
1
`

unnorm
. (5.10)
The Schrdinger equation (which should be posted on your refrigerator),

H = 1. (5.11)
43
43
The Schrdinger equation for 1D problems as a dierential equation,
~
2
2:
d
2
dr
2
+ (\ (r) 1) = 0. (5.12)
How to get the average value for some property (1D version),
h ci =
Z
space

+
cdr. (5.13)
The momentum operator
j
a
= i~
J
Jr
. (5.14)
Normalized wavefunctions for the 1D particle in a box,

a
(r) =
r
2
c
sin
::r
c
. (5.15)
The energy levels for the 1D particle in a box,
1
a
=
:
2
:
2
~
2
2:c
2
~=
I
2r
=
:
2
/
2
8:c
2
. (5.16)
The energy level spacing for the 1D particle in a box,
41 = (2: + 1)
/
2
8:c
2
(5.17)
The wavefunctions for the harmonic oscillator are

a
() =
a
H
a
()c

2
2
. =

/:
~
2
1
4
r.
a
=
1
p
2
a
:!
_
:
. (5.18)
where
a
is the normalization constant for the :
th
eigenfunction and H
a
()
are the Hermite polynomials.
The energy levels are
1
a
= (: +
1
2
)~.. . =
r
/
:
(5.19)
44
Part II
Quantum Mechanics of Atoms
and Molecules
45
45
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 Schrdinger equation
for. So this will be the rst atomic system we discuss.
The Schrdinger 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 Li
2+
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
with the origin located on the nucleus.
The classical potential energy for these hydrogenic systems is
\ (:) =
2c
2
(4:c
0
):
. (6.1)
So the Hamiltonian is

H =
~
2
2:
c

\
2
+
2c
2
(4:c
0
) :
. (6.2)
Schrdingers equation (in spherical polar coordinates) becomes
1 =

H (6.3)
1 =

~
2
2:
c

\
2
+
2c
2
(4:c
0
) :

1 =

~
2
2:
c

1
:
2
J
J:
:
2
J
J:
+
1
:
2

1
sin o
J
Jo
sin o
J
Jo
+
1
sin
2
o
J
2
Jc
2

+
2c
2
(4:c
0
):

The Hamiltonian is (almost) the sum of a radial part (only a function of :) and
an angular part (only a function of o and c):

H =

H
rad
+
1
:
2

H
ang
. (6.4)

H
rad
=
~
2
2:
c

1
:
2
J
J:
:
2
J
J:

2c
2
(4:c
0
):

(6.5)
and

H
ang
=
~
2
2:
c

1
sino
J
Jo
sin o
J
Jo
+
1
sin
2
o
J
2
Jc
2

(6.6)
Since the Hamiltonian is the sum of two terms, must be a product state.
(:. o. c) =
rad
(:)
ang
(o. c) (6.7)
It turns out that solving the Schrdinger equation,

H
ang

ang
(o. c) = 1
ang
(o. c). (6.8)
47
yields

ang
(o. c) = 1
|n
(o. c). (6.9)
where the 1
|n
(o. c)s are the spherical harmonic functions characterized by quan-
tumnumbers | and :. 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

H
ang
is found to be
1 = 1
|
=
|(| + 1)~
2
2:
c
. (6.10)
So,

H
ang

ang
(o. c) =
|(| + 1)~
2
2:
c

ang
(o. c) (6.11)
Now lets denote the radial part of the wavefunction as
rad
(:) = 1(:).
The full Schrdinger equation becomes

H(:. o. c) = 1(:. o. c) (6.12)

H1(:)1
|n
(o. c) = 11(:)1
|n
(o. c)


H
rad
+
1
:
2

H
ang

1(:)1
|n
(o. c) = 11(:)1
|n
(o. c).
Operating with

H
ang
we get


H
rad
+
|(| + 1)~
2
2:
c
:
2

1(:)1
|n
(o. c) = 11(:)1
|n
(o. c) (6.13)
48
The 1
|n
(o. c) can now be cancelled to leave a one dimensional dierential equation:
~
2
2:
c

1
:
2
J
J:
:
2
J
J:

2c
2
4:c
0
:

|(| + 1)
:
2

1(:) = 11(:). (6.14)


This dierential equation is very similar to a known equation called Laguerres
dierential equation which has as solutions the Laguerre polynomials 1
|
a
(r).
In fact, the solutions to our dierential equation are closely related to the Laguerre
polynomials.
1
a|
(o) =
a|

2o
:

|
c
oa
1
2|+1
a+1

2o
:

. (6.15)
where the normalization constant,
a|
. depends on the : and | quantum numbers
as

a|
=
s

22
:c
0

3
(: | 1)!
2:[(: +|)!]
3
(6.16)
The energy eigenvalues, i.e., the energy levels are given by
1
a
=
2
2
R
:
2
(6.17)
Note: The energy levels are determined by : alone| 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 briey discussed later) is

a|n
(:. o. c) = 1
a|
(:)1
|n
(o. c) (6.18)
6.2. Discussion of the Wavefunctions
We are now very close to having the atomic orbitals familiar from freshman chem-
istry.
49
We have explicitly derived the physicists picture of the atomic orbitals
orbital : | : wavefunctions (o = :,c
0
)
1s 1 0 0
1c
=
100
= c
o
2s 2 0 0
2c
=
200
=

1
o
2

c
o2
2p 2 1 0
2j
0
=
210
= oc
o2
cos o
2 1 1
2j
1
=
211
= oc
o2
sin oc
i
3d 3 2 0
3o
0
=
320
= o
2
c
o3
(3 cos
2
o 1)
3 2 1
3o
1
=
321
= 1
32
(:) cos o sinoc
i
3 2 2
3o
2
=
322
= 1
32
(:) sin
2
oc
i2
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
1c
.
2c
.
2j
0
.
3o
0
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
orbital : | : wavefunctions (o = :,c
0
)
1s 1 0 0
1c
=
1c
2s 2 0 0
2c
=
2c
2p 2 1 0
j
:
=
2j
0
2 1 1
2j
i
=
1
_
2

2j
1
+
2j
31

2 1 1
2j

=
1
i
_
2

2j
1

2j
31

3d 3 2 0
3o
:
2
=
3o
0
3 2 1
3o
i:
=
1
_
2

3o
1
+
3o
31

3 2 1
3o
:
=
1
i
_
2

3o
1

3o
31

3 2 2
3oi
=
1
_
2

3o
2
+
3o
32

3o
i
2
3
2
=
1
i
_
2

3o
2

3o
32

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 o a more detailed discussion until
next semester.
For now we must be satised with the following:
There are two quantum numbers associated with spin: : and :
c
: is the spin quantum number and for an electron : = 1,2 (always).
:
c
is the spin orientation quantum number and :
c
= 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
We simply denote the spin wavefunction generally as
c,ns
and tack it on as
another factor of the complete wavefunction.
When a particular spin state is needed a further notation is commonly used:
c = 1
2
,
1
2
(the spin-up state) and , = 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 eects)
The full wave function is
[
a,|,n,c,n
s
=
a,|,n

c,ns
(6.19)
= 1
a|
(:)1
|,n
(o. c)
The energy is given by
1
a
=
2
2
R
:
2
. (6.20)
where recall. Again note that for a free hydrogenic system the total energy depends
only on the principle quantum number :.
The quantum numbers of the hydrogenic system
The principle quantum number, :: determines the total energy of the sys-
tems and the atomic shells.
The principle quantum number, :. can take on values of 1,2,3. . .
The angular momentum quantum numbers, |: determines the total angular
momentum of the system. It also determines the atomic sub-shells
52
The angular momentum quantum number, |. can take on values of 0,
1. . . . (: 1)
For historical reasons | = 0 is called :, | = 1 is called j. | = 2 is called
d. | = 3 is called , etc.
The orientation quantum number, :: determine the projection of the an-
gular momentum onto the .-axis. It also determines the orientation of the
atomic sub-shells
The magnetic quantum number, :. can take on values of 0, 1. . . . |.
The spin quantum number, :: determines the total spin angular momentum.
For electrons : = 1,2.
The spin orientation quantum number, :
c
: determines the projection of the
spin angular momentum onto the .-axis (i.e., spin-up or spin-down).
For electrons :
c
= 1,2
We have accomplished quite a bit. We have determined all that we can about the
hydrogen atom within Schrdingers theory of quantum mechanics.
This is not the full story however. The Schrdinger theory is a non-relativistic
one; that is, it can not account for relativistic eects which show up in spectral
data. We also had to add spin in an ad hoc manner to account for what we know
experimentallyspin did not fall out of the theory naturally.
Dirac, in the late 1920s, developed a relativistic quantum theory in which the
well established phenomenon of spin arose naturally. His theory also made the
53
bold prediction of the existence anti-matter that has now been veried 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 Schrdinger theory.
54
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 cant be solved exactly.
The helium atom is an example of the three-body-problemdi!cult to handle
even in classical mechanicsone can not get a closed form solution.
The Hamiltonian for helium is

H =
~
2
2:
c
\
2
1
| {z }
K.E of electron 1

~
2
2:
c
\
2
2
| {z }
K.E of electron 2

2c
2
4:c
0
:
1
| {z }
P.E of electron 1

2c
2
4:c
0
:
2
| {z }
P.E of eletcron 2
+
c
2
4:c
0
:
12
| {z }
elec.elec. repulsion
. (7.1)
where :
12
= |:
1
:
2
| is the distance between the electrons.
The electronelectron repulsion term is responsible for the di!culty 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
Perturbation theory (next semester)
Variational theory (next semester)
Ignore the electronelectron 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 ||
2
is directly measurablenot itself.
All this implies that
(1. 2) =
;
A
?
A
=
+(2. 1). symmetric
or
(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
The total wavefunction is
[ = (1. 2)(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).
[ = (1)(2)
| {z }
spatial part
(1)(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 lets blindly list all possibilities for the ground state wave function of helium
[
o
=
1c
(1)c(1)
1c
(2)c(2) (7.5)
[
b
=
1c
(1)c(1)
1c
(2),(2)
[
c
=
1c
(1),(1)
1c
(2)c(2)
[
o
=
1c
(1),(1)
1c
(2),(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
[
o
has symmetric spatial and spin parts and is there for symmetric. It must
be excluded.
Similarly for [
o
.
[
b
and [
c
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
The appropriate linear combination is
c(1),(2) c(2),(1). (7.6)
So the ground state wave function for helium is
[
j
=
1c
(1)
1c
(2) [c(1),(2) c(2),(1)] . (7.7)
Consequences of the Pauli exclusion principle
No two electrons can have the same ve 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:
[ =

1c
(1)c(1)
1c
(1),(1)
a
(1)c(1)
a
(1),(1)

1c
(2)c(2)
1c
(2),(2)
a
(2)c(2)
a
(2),(2)
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

1c
(`)c(`)
1c
(`),(`)
a
(`)c(`)
a
(`),(`)

(7.8)
58
The reason one can be sure that this wavefunction is the antisymmeterized is that
we knowfromlinear 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:
1c
c.

1c
,. and
2c
c (or
2c
,).
We construct the Slater determinant as
[
1
=

1c
(1)c(1)
1c
(1),(1)
2c
(1)c(1)

1c
(2)c(2)
1c
(2),(2)
2c
(2)c(2)

1c
(3)c(3)
1c
(3),(3)
2c
(3)c(3)

(7.9)
or
[
2
=

1c
(1)c(1)
1c
(1),(1)
2c
(1),(1)

1c
(2)c(2)
1c
(2),(2)
2c
(2),(2)

1c
(3)c(3)
1c
(3),(3)
2c
(3),(3)

(7.10)
The short hand notation for these states is (1:)
2
(2:)
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 =
.
X
i=1
"
~
2
2:
c
\
2
i

2c
2
4:c
0
:
i
+
X
)i
c
2
4:c
0
:
i)
#
(7.11)
59
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 : electrons requires that 3:+6 coordinates be specied.
Three of these describe the center of mass position.
3: of these describe the position of the : electrons.
This leaves three degrees of freedom (1. o. c) which describe the position of the
nuclei relative to the center of mass. 1 determines the internuclear separation
and o and c determine the orientation.
60
60
8.1.1. The Hamiltonian
In the center of mass coordinates the Hamiltonian for a diatomic molecule is

H =

1
.
+

1
c
+

\
..
+

\
.c
+

\
cc
. (8.1)

1
.
is the nuclear kinetic energy operator and is given by

1
.
=
~
2
2j

\
2
.
=
~
2
2j1
2
J
J1

1
2
J
J1
+
~
2
2j

J
2
. (8.2)
where

J is angular momentum operator for molecular rotation and j =
n
1
n
2
n
1
+n
2
is
the reduced mass of the diatomic molecule.

1
c
=
P
i

~
2
2nc

\
2
c
.
is the kinetic energy operator for the electrons.

\
..
=
Z
/
Z
Tc
c
2
4c
0
1
is the nuclearnuclear potential energy operator.

\
.c
=
P
i
h
Z
/
c
2
4c
0
v
/.
+
Z
T
c
2
4c
0
v
T.
i
is the nuclearelectron potential energy operator.

\
cc
=
P
i)
c
2
4c
0
v
.
is the electronelectron potential energy operator.
61
8.1.2. The BornOppenheimer Approximation
The BornOppenheimer approximation: The nuclei move much slower than
the electrons. (classical picture)
We put the BornOppenheimer approximation to work by rst dening an eec-
tive Hamiltonian

H
c))
=

1
c
+

\
..
+

\
.c
+

\
cc
. (8.3)
The approximation comes in by treating 1 as a parameter rather than an operator
(or variable). So one writes

H
c))

c
(1. {:
i
}) = 1
c
(1)
c
(1. {:
i
}). (8.4)

c
is the so-called electronic wavefunction.
Now the Schrdinger equation for the diatomic molecule is

1
.
+

H
c))

(1. {:
i
}) = 1(1. {:
i
}). (8.5)
Since the Hamiltonian is a sum of two terms, one can write the wavefunction
(1. {:
i
}) as a product wavefunction
=
.

c
. (8.6)
where
.
is the so-called nuclear wavefunction.
Substituting the product wavefunction into the Schrdinger equation gives

1
.
+

H
c))

c
= 1
.

c
(8.7)

1
.
+1
c
(1)

c
/ = 1
.

c
/

1
.
+1
c
(1)

.
= 1
.
.
62
The last equation is exactly like a Schrdinger equation with a potential equal to
1
c
(1).
One now models 1
c
(1) or determines it experimentally.
8.2. Molecular Vibrations
As stated earlier 1 is the internuclear separation and o and c determine the
orientation. Consequently, 1 is the variable involved with vibration whereas o
and c are involved with rotation.
Considering only the 1 part of the Hamiltonian (under the BornOppenheimer
approximation), we have

~
2
2j
J
2
J1
2
+1
c
(1)

vib
= 1
vib

vib
. (8.8)
It is convenient at this point to expand 1
c
(1) in a Taylor series about the equi-
librium position, 1
eq
:
1
c
(1) = 1
0
+

J1
J1

1
eq
(11
eq
) +
1
2!

J
2
1
J1
2

1
eq
(11
eq
)
2
+ . (8.9)
Now 1
0
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,

01
01

1eq
must be zero, so the linear term vanishes.
One denes

0
2
1
01
2

1eq
= /
c
as the force constant.
The remaining terms in the expansion can collective be dened as C[(11
eq
)
3
] =
\
anh
. the anharmonic potential.
63
As a rst approximation we can neglect the anharmonicity. With this, the Schrdinger
equation becomes

~
2
2j
J
2
J1
2
+
1
2
/
c
(11
eq
)
2

vib
= 1
vib

vib
. (8.10)
If we let r = (11
eq
) this becomes

~
2
2j
J
2
Jr
2
+
1
2
/
c
r
2

vib
= 1
vib

vib
. (8.11)
which is exactly the harmonic oscillator equation. Hence

vib,a
=
a
H
a
(
_
cr)c
ca
2
2
. (8.12)
where c =
q
Icj
~
.
And
1
vib,a
= /c .
c
(: +
1
2
). (8.13)
where .
c
=
1
2
q
I
c
j
.
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 potentialone that will allow of bond
dissociation.
The Morse potential
1
c
(1) = 1
c
[1 c
o(11eq )
]
2
. (8.14)
64
where 1
c
is the well depth and , = 2:c .
c
q
j
21
c
is the Morse parameter. Note:
this expression for the Morse potential has the zero of energy at the bottom of
the well (i.e. 1 = 1
eq
. ;1
c
(1
eq
) = 0).
The Morse Potential can also be written as
1
c
(1) = 1
c
[c
2o(11
eq
)
2c
o(11
eq
)
]. (8.15)
Now the zero of energy is the dissociated state (i.e. 1 . ;1
c
(1 ) = 0).
We approach this quantum mechanical problem exactly like all the other.
The Schrdinger equation is

~
2
2j
J
2
J1
2
+1
c
[1 c
o(11
eq
)
]
2

vib
= 1
vib

vib
(8.16)
This is another dierential equation that is di!cult to solve.
As it turns out, this Schrdinger equation can be transformed into a one of a broad
class of known dierential equations called conuent hypergeometric equations
the solutions of which are the conuent hypergeometric functions,
1
1
1
.
Doing this yields the wavefunctions of the form

vib,a
(.) = .
jn
c
:
1
1
1
(:. 1 + 2j
a
. 2.). (8.17)
. =
_
21
c
j
,/
c
oa
.
=
_
2j
,/
.
j
a
=
p
1
c
+

1
2
:

and energy levels of the form


1
vib,a
= 1
c
+/c .
c
(: +
1
2
) /c .
c
r
c
(: +
1
2
)
2
. (8.18)
65
where .
c
r
c
together is the anharmonicity constant, with r
c
=
Ic .
c
41
c
.
+ + + 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 =
41
Ic
(in units of wavenumbers, cm
1
).
The transition from the : = 0 to the : = 1 state is called the fundamental
transition.
Transitions from : = 0 to : = 2. 3. 4 are called overtone transitions.
Transitions from : = 1 to 2. 3. 4 , : = 2 to 3. 4. 5 . etc. are called hot
transitions (or hot bands)
Since the energy levels depend on mass, isotopes will have a dierent transition
energy and hence appear in a dierent place in the spectrum. Heavier isotopes
have lower transition energies.
66
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 bondingVSEPR (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 rst example, let us consider the molecular hydrogen ion H
+
2
.
The Hamiltonianfor H
+
2
is

H =

1
.
+

1
c|
+

\
.c|
+

\
..
(9.1)
67
67
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 eective Hamiltonian becomes

H =

1
c|
+

\
.c|
(9.2)
=
~
2
2:
c
\
2

c
2
4:c
0
:


c
2
4:c
0
:
1
.
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 orbitals
are built from the hydrogen atomic orbitals.
Molecular orbitals: The hydrogen molecular ion is the prototype and all
other molecular orbitals are built from the hydrogen molecular ion molecular
orbitals.
There is one signicant dierence 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 .-axis of the center of mass
coordinate system.
68
Inversion symmetry
The potential eld of the hydrogen molecular ion is cylindrically symmetric
about the .-axis.
Because of the symmetry the electron density at (r. . .) must equal the
electron density at (r. . .).
The above symmetry therefore requires that the molecular orbitals be eigen-
functions of the inversion operator, :. That is
:(r. . .) = (r. . .) = c(r. . .). (9.3)
Moreover the eigenvalue c can be either +1 or 1.
If c = +1 the molecular wavefunction is even with respect to inversion and
is called gerade and labelled with a q: :
j
=
j
If c = 1 the molecular wavefunction is odd with respect to inversion and
is called ungerade and labelled with a n: :
&
=
&
The terms gerade and ungerade apply only to systems that posses inversion
symmetry.
Cylindrical symmetry
69
The cylindrical symmetry implies that the potential energy can not depend
on the c.
The molecular wavefunction is described by an eigenvalue ` = 0. 1. 2. . . .
We use ` to label the molecular orbitals as shown in the table
` 0 1 2
label o : o
Mirror plane symmetry
70
There is also a symmetry about the r- plane called horizontal mirror plane
symmetry: operator o
I
.
Thus the molecular wavefunction must be an eigenfunction of o
I
with eigen-
value 1.
If the eigenvalue is +1 (even with respect to o
I
) the molecular orbital
is called a bonding orbital.
If the eigenvalue is 1 (odd with respect to o
I
) the molecular orbital
is called an antibonding orbital.
There are also vertical mirror plane symmetries, but we will put that dis-
cussion o for the time being.
71
10. Molecular Orbital Diagrams
10.1. LCAOLinear Combinations of Atomic Orbitals
Now that we know what symmetry the molecular orbitals must posses, we need
to nd some useful approximations for them.
Useful can mean qualitatively useful or quantitatively useful.
Unfortunately we cant 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
and the other labelled 1.
Linear combination of the 1: atomic orbital from each H atom is used for the
molecular orbital of H
+
2
:
(1:

) = /c
v
/
o
0
(10.1)
and
(1:
1
) = /c
v
T
o
0
(10.2)
72
72
We construct two molecular orbitals as
x
+
= C
+
(1:

+ 1:
1
) (10.3)
and
x

= C

(1:

1:
1
) (10.4)
The normalization condition is
Z
x

dl = 1 (10.5)
As can be seen from the above gure, x
+
represents a situation in which the
electron density is concentrated between the nuclei and thus represents a bonding
orbital.
Conversely x

represents a situation in which the electron density is very low


between the nuclei and thus represents an antibonding orbital
10.1.1. Classication of Molecular Orbitals
With atoms we classied 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
The classication according to angular momentum is as follows.
` 0 1 2
orbital symbol o : o
Atomic orbitals with : = 0 form o type molecular orbitals, e.g., : =o. j
:
=o.
Those with : = 1 form : type molecular orbitals, e.g., j
a
=: etc.
The classication according to inversion symmetry is simply a subscript q or
n. For example, o
j
or o
&
etc.
The classication according to bonding or antibonding is an asterisk is used to
denote antibonding. For example, o
j
is a bonding orbital and o
+
&
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 innitely far apart. The ground state of the system would
consist of two separate hydrogen molecules in their ground atomic states: (1:)
1
74
As the atom are brought closer together, their respective : 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
o
j
= 1:

+ 1:
1
(10.6)
and
o
+
&
= 1:

1:
1
. (10.7)
75
It can be shown mathematically that the energy level associated with o
j
is lower
than o
+
&
.
We can intuit this qualitatively however since the o
+
&
orbital must have a node
whereas the o
j
does not.
It is also to be expected since we know H
2
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 H
2
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
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 1: orbital
may combine with a low lying 2: orbital to form a o 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 dened as
BO =
1
2
(# of bonding electrons # of antibonding electrons) (10.8)
Examples follow in the supplement.
77
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 Schrdinger equation for a molecule as follows.
(Next semester will we look at the details of this for polyatomic molecules)

H
mol
[
mol
= 1
mol
[
mol
(10.9)


H
ele
+

H
vib
+

H
rot

ele

vib

rot
= (1
ele
+1
vib
+1
rot
)
ele

vib

rot
78
11. An Aside: Light ScatteringWhy
the Sky is Blue
This chapter addresses the topic of light scattering from two dierent 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 eld part of light with a dipole.
When the light interacts with the molecule an electric dipole is induced according
to
j = c1. (11.1)
where c is the polarizability of the molecule describing the exibility of its
electron cloud.
79
79
For light, the electric eld part is
1(t) = 1
0
cos .t. (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:
c(t) = c
0
+c
1
cos .

t (11.3)
(here for simplicity we assume only one vibrational mode).
Thus the lightmatter interaction is described as
j(t) = c(t)1(t) = (c
0
+c
1
cos .

t) 1
0
cos .t (11.4)
= c
0
1
0
cos .t +c
1
1
0
cos .

t cos .t
= c
0
1
0
cos .t
| {z }
Rayleigh
+
c
1
1
0
2
5
7
cos(. .

)t
| {z }
Stokes Raman
+ cos(. +.

)t
| {z }
AntiStokes Raman
6
8
where a trig identity was used in the last step.
According to classical electrodynamics an oscillating dipole emits an electromag-
netic eld at the oscillation frequency.
In this case we see the dipole oscillates at three distinct frequencies: .. . .

and . +.

as part of three terms in the above expression.


The rst 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
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
1 =
.
4
3c
3
j
2
0
. (11.5)
where j
0
= c
0
1
0
for the case of Rayleigh scattering and j
0
= c
1
1
0
,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 1 b .
4
or alternatively 1 b 1,`
4
. There is a
very 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 at as shown below.
81
We just learned that light scatters as it traverses the atmosphere according to
Rayleighs scattering law: 1(`) b 1,`
4
.
The following gures 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
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
We might expect that clouds should be highly colored since they consist of droplets
of water which scatter light very eectively.
83
The key dierence 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 lightquite the
opposite case as above.
In this limit an entirely dierent analysis is madeone 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
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

a|n
(:. o. c) = 1
a|
(:)1
|n
(o. c) (11.6)
The radial part is.
1
a|
(o) =
a|

2o
:

|
c
oa
1
2|+1
a+1

2o
:

. (11.7)
where the normalization constant,
a|
. depends on the : and | quantum
numbers as

a|
=
s

22
:c
0

3
(: | 1)!
2:[(: +|)!]
3
(11.8)
85
85
The energy levels for the hydrogenic system are given by
1
a
=
2
2
R
:
2
(11.9)
The wavefunctions for the harmonic oscillator are

a
() =
a
H
a
()c

2
2
. =

/:
~
2
1
4
r.
a
=
1
p
2
a
:!
_
:
. (11.10)
where
a
is the normalization constant for the :
th
eigenfunction and H
a
()
are the Hermite polynomials.
The energy levels are
1
a
= (: +
1
2
)~.. . =
r
/
:
(11.11)
The Morse potential is
1
c
(1) = 1
c
[1 c
o(11
eq
)
]
2
. (11.12)
where 1
c
is the well depth and , = 2:c .
c
q
j
21c
is the Morse parameter.
Note: this expression for the Morse potential has the zero of energy at the
bottom of the well (i.e. 1 = 1
eq
. ;1
c
(1
eq
) = 0).
The Morse Potential can also be written as
1
c
(1) = 1
c
[c
2o(11
eq
)
2c
o(11
eq
)
]. (11.13)
Now the zero of energy is the dissociated state (i.e. 1 . ;1
c
(1 ) =
0).
The energy levels for the Morse oscillator are of the form
1
vib,a
= 1
c
+/c .
c
(: +
1
2
) /c .
c
r
c
(: +
1
2
)
2
. (11.14)
where .
c
r
c
together is the anharmonicity constant, with r
c
=
Ic .
c
41
c
.
86
Bond order is dened as
BO =
1
2
(# of bonding electrons # of antibonding electrons) (11.15)
The Rayleigh scattering law is
1(`) b 1,`
4
b .
4
(11.16)
87
Part III
Statistical Mechanics and The
Laws of Thermodynamics
88
88
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 Schrdinger equation for those 100 million molecules, but often Avogadros
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 denitions
Ensemble: A large collection of equivalent macroscopic systems. The sys-
tems are the same except that each one is in a dierent so-called microstate.
89
89
Microstate: The single particular state of one member of the ensemble given
by listing the individual states of each of the microscopic systems in the
macroscopic state.
Conguration: The collection of all equivalent microstates. The number of
possible congurations is dened as \.
Boltzmann developed an equation to connect the microscopic properties of an
ensemble to the macroscopic properties. The Boltzmann equation is
o = / ln \ (12.1)
Where o is entropy and / is Boltzmanns constant.
12.1.1. Combinations and Permutations
Consider a random system that when measured can appear in one of two outcomes
(e.g., ipping coins).
One valuable piece of statistical information about system is knowing how many
dierent ways the system appears j times in, say, outcome 1 after ` measure-
ments.
This is given by the mathematical formula for combinations
C(`. j) =
`!
j!(` j)!
. (12.2)
The number C(`. j) is also called the binomial coe!cient because it gives the
coe!cient for the j
tI
order term in the expansion
(1 +r)
.
=
.
X
j=0
C(`. j)r
j
. (12.3)
90
This formula will allow us to derive a normalization constant so that we can obtain
the probability of obtaining j measurements of state 1.
Set r = 1 in the above. This gives
(1 + 1)
.
=
.
X
j=0
C(`. j)(1)
j
(12.4)
2
.
=
.
X
j=0
C(`. j).
So the probability of any one outcome of ` measurements is
1(`. j) =
1
2
.
C(`. j) =
1
2
.
`!
j!(` j)!
(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
\(`. {`
i
}) =
`!
`
1
!`
2
!`
3
!
(12.6)
where ` is the total number of measurements and `
i
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
This is no problem for small numbers, but when we consider macroscopic systems
(10
20
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 Sterlings ap-
proximation to approximate the value of the factorial.
Sterlings approximation is
ln(`!) ' ` ln ` ` (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 uctuates about the average value. Often
this uctuation is not important, but sometimes it is important.
The uctuation about an average value for any observable property C is described
by the variance which is dened as
o
2
O
= C
2


C
2
. (12.8)
o
O
is consider the range of the observable property.
It can be shown that
o
O

C
-
1
_
`
. (12.9)
where ` is the number of particles. So for example if ` = 10
24
then
1
_
.
= 10
12
92
For ensembles having large numbers of particles measured values of a property are
extremely sharply peaked about the average value.
93
13. The Boltzmann Distribution
Consider a isolated system of ` molecules that has the set {c
i
} energy levels
associated with it.
Since the system is isolated the total energy, 1. and the total number of particles
will be constant.
The total energy is given by
1 =
X
i
`
i
c
i
. (13.1)
where `
i
is the number of particles in energy state i.
The total number of particles is, of course,
` =
X
i
`
i
(13.2)
The number of congurations for the system is then given by the number of
distinct permutations of the system
\ =
`!
`
1
!`
2
!
. (13.3)
A system in equilibrium always tries to maximize entropy and minimize energy
and so the equilibrium conguration is a compromise between these two cases.
94
94
For the moment let us relax the isolation constraint.
Maximizing entropy corresponds to maximizing \ (via o = / ln \). This would
be the situation in which every particle was in a dierent energy state. That is
all `
i
= 1 or 0.
Minimizing energy would be the case where all the particles are in the ground
state (say c
1
).
These two situations are contradictory and some compromise must be obtained.
We start by considering our original systemthat being one with constant energy,
1 and number of particles `
To determine the equilibrium conguration we must nd the maximum \ 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 conguration.
The derivation using Lagrange multipliers arrives at the conguration in which
the
`
i
= `
q
i
c
oc
.
P
)
q
)
c
oc

| {z }
j
.
. (13.4)
where , =
1
IT
and q
)
denotes the degeneracy of states having energy c
)
.
95
The j
i
represents the probability of nding the a randomly chosen particle or
system which has energy c
i
. This is the Boltzmann distribution
1
i
=
q
i
c
oc
.
P
)
q
)
c
oc

(13.5)
Since we started with a isolated system, , and hence 1 are constants. A given
energy 1 will correspond to a unique temperature 1.
The analysis readily generalizes to variable energy i.e., nonisolated systems by
considering 1 as a variable.
13.1. Partition Functions
We have already come across both the partition functions that we will use in this
class.
The rst is \the number of congurations. 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 =
X
)
q
)
c
o1

(13.6)
and is called the canonical partition function.
96
This was rst encountered as the denominator of the Boltzmann distribution and
it is extremely important in statistical mechanics. (Note: the symbol 2 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 o = / ln\. 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 dierent situations but we
will do nothing more than list two important ones here: i) the grand canonical
partition function and ii) the isothermalisobaric partition function
13.1.1. Relation between the Q and \
When we get to connecting quantum mechanics with thermodynamics it will prove
convenient to use Boltzmanns equation (o = / ln\) but as was stated earlier it
is not convenient to use the microcanonical partition function (\).
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
The microcanonical partition function describes a system at xed energy 1. In
fact \ is the number of available states of the ensemble at the particular energy
1. This is essentially the same as the degeneracy of the ensemble q
1
.
Conversely the canonical partition function describes a system with variable en-
ergy.
However, based on our previous discussion of uctuations, even though the energy
of the ensemble is allowed to vary, the number of states with energy equal to the
average energy

1 is overwhelmingly large. That is, almost every state available
to the ensemble has energy

1.
We can express these ideas mathematically to come up with a relation between
\ and Q.
The canonical partition function is
Q =
X
)
q
)
c
oc

. (13.7)
but to a good approximation
Q ' q
1
c
o

1
. (13.8)
Now since the degeneracy is essentially the microcanonical partition function we
have
Q ' \c
o

1
. (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 Boltmanns equation in terms of the canonical
98
partition function:
ln Q ' ln(\c
o

1
) = ln \ + ln(c
o

1
) (13.10)
= ln\
|{z}
SI


1
/1
.
so,
o = / lnQ+

1
1
(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
c = c
ele
+c
vib
+c
rot
. (13.12)
This is a consequence of the Born Oppenheimer approximation
If we include the center of mass translational motion this is
c = c
ele
+c
vib
+c
rot
+c
trans
(13.13)
The i
th
total energy level is
c
i
= c
ele,a
+c
vib,
+c
rot,J
+c
trans,n
. (13.14)
Now if we have a collection of molecules in a macroscopic system. A given con-
guration (say, conguration ,) of that system has total energy 1
)
.
So the canonical partition function is
Q =
X
)
q
)
c
o1

(13.15)
99
But, each 1
)
is made up of the contributions of all of the molecules:
1
)
= c
o
|
+c
b
n
+c
c
a
+ (13.16)
The partition function for the molecule is written as
Q =
X
)
q
)
c
o1

=
X
|,n,a
(q
o
|
q
b
n
q
c
a
)c
o(c
a
I
+c
l
r
+c
c
n
+ )
(13.17)
=
X
|
q
o
|
c
oc
a
I
| {z }
q
mol,a
X
n
q
o
n
c
oc
a
r
| {z }
q
mol,l
X
a
q
o
a
c
oc
a
n
| {z }
q
mol,c

where the
mol,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
mol,i
are the same:

mol,i
=
mol
thus
Q =

.
mol
`!
. (13.18)
This allows us to focus only on a single molecule:

mol
=
X
i
q
i
c
oc
.
=
X
a,,J,n
q
ele,a
q
vib,
q
rot,J
q
trans,n
c
o
(
c
ele,n
+c
vib,r
+c
rot,
+c
trans,r)
(13.19)
X
a
q
ele,a
c
oc
ele,n
| {z }
q
ele
X

q
vib,
c
oc
vib,r
| {z }
q
vib
X
J
q
rot,J
c
oc
rot,
| {z }
q
rot
X
n
q
trans,n
c
octrans,r
| {z }
q
trans
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
The Translational Partition Function

trans
=
\
\
3
(13.20)
where
\ =
/
_
2::/1
(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 H
2
is

rot
-
1
oo
v
(13.22)
where o
v
=
I
2
8
2
1I
(1 is the moment of inertia) and o is the so-called symmetry
number in which o = 1 for unsymmetrical molecules and o = 2 for symmetrical
molecules.
The Vibrational Partition Function

vib
=
c

1
2
o~.
1 c
o~.
=
1
2 sinh
1
2
,~.
(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
so

ele
=
X
i
q
ele,i
c
oc
tele,.
- q
ele,ground
(13.24)
102
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, l. pressure, 1.
entropy, o. and Helmholtz free energy, . 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

C =
1
Q
X
i
C
i
q
i
c
oc
.
. (14.1)
Internal energy
One critical property of an ensemble is the average (internal) energy l.
l =

1 =
1
Q
X
i
c
i
q
i
c
oc
.
. (14.2)
Let us look closer at the above expression. Recall that
Q =
X
i
q
i
c
oc
.
. (14.3)
103
103
Now take the derivative of Q with respect to , gives

JQ
J,

a,\
=

J
J,
"
X
i
q
i
c
oc
.
#!
a,\
=
X
i
q
i

Jc
oc
.
J,

a,\
(14.4)
=
X
i
q
i
c
i
c
oc
.
By comparing this to the expression for l. we see
l =
1
Q

JQ
J,

a,\
=

J ln Q
J,

a,\
. (14.5)
where we used the identity
1
j
0j
0a
=
0 lnj
0a
.
Pressure
Another important property is pressure.
When the ensemble is in the particular state i, dc
i
= j
i
d\ . So at constant
temperature and number of particles
j
i
=

Jc
i
J\

a,o
(14.6)
Thus the ensemble average pressure is given by
1 = j =
1
Q
X
i
q
i

Jc
i
J\

a,o
c
oc
.
. (14.7)
Multiplying by ,,, we get
1 =
1
,Q
X
i
q
i

Jc
i
J\

a,o
,c
oc
.
. (14.8)
Using the chain rule in reverse, i.e.,
Jc
oc
.
J\
=
oc
3c
.
z }| {

Jc
oc
.
Jc
i

Jc
i
J\

Jc
i
J\

,c
oc
.
(14.9)
104
we proceed as
1 =
1
,Q
X
i
q
i

Jc
oc
.
J\

a,o
=
1
,Q

J
J\
X
i
q
i
c
oc
.
!
a,o
(14.10)
=
1
,Q

JQ
J\

a,o
=
1
,

J ln Q
J\

a,o
.
Entropy
We have already obtained the expression for entropy. It is
o =
l
1
+/ ln Q (14.11)
= /,

J lnQ
J,

a,\
+/ ln Q
105
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: = l 1o
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 Gibbs free energy (G) later as well.
The Helmholtz free energy has the most direct relation to the partition function
as can be seen from
= l 1o =

J ln Q
J,

a,\
+/1,

J lnQ
J,

a,\
/1 lnQ (14.12)
= /1 lnQ
Any thermodynamic property can now be obtained from the above functions as
we shall see in the following chapters.
106
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 procient with
partial derivatives.
+ + + See Handout + ++
15.1.1. Summary of Relations
1. The total derivative of .(r. ):
d. =

J.
Jr

j
dr +

J.
J

a
d (15.1)
2. The chain rule for partial derivatives:

J.
Jr

j
=

J.
Jn

Jn
Jr

j
(15.2)
107
107
3. The reciprocal rule:

J.
Jr

Jr
J.

j
= 1 (15.3)
4. The cyclic rule:

J.
Jr

j
=

J.
J

J
Jr

:
(15.4)
5. Finally

J.
Jr

&
=

J.
Jr

j
+

J.
J

J
Jr

&
(15.5)
15.2. Denitions
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
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 of
matter.
For example, volume, mass, heat capacity.
Intensive parameters (or properties): properties that are independent of the
amount of matter.
For example, temperature, pressure, density.
Extensive properties can be converted to intensive properties through ratios:
Extensive property
Extensive property
Intensive property. (15.6)
For example
mass
volume
= density,
volume
moles
= molar volume,
heat capacity
mass
= specic 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
Convention
Work, n. is positive (n 0) if work is done on the system.
Work is negative (n < 0) if work is done by the system.
Heat, . is positive ( 0) if heat is absorbed by the system.
Heat is negative ( < 0) if heat is released from the system.
15.3.1. Generalized Forces and Displacements
In physics you learned that an innitesimal change in work is given by the product
of force, 1, times and innitesimal change in position, dr:
dn = 1dr. (15.7)
For thermodynamics, we need a more general denition if innitesimal 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 innitesimal amount of work done on the system is then given by
dn = Ada. (15.8)
or more generally as
dn =
X
i
A
i
da
i
(15.9)
110
if more than one set of parameters change.
The following table gives some examples of generalized forces and displacements
Generalized Force, A Generalized Displacement, a Contribution to dn
Pressure, 1 Volume, d\ 1d\
Stress, o Strain, d od
Surface tension, Surface area, dA dA
Voltage, E Charge, dQ EdQ
Magnetic Field, H Magnetization, dM HdM
Chemical Potential, j Moles, d: jd:
Gravity, :q Height, d/ :qd/
15.3.2. 1\ 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
dn = 1d\. (15.10)
this is called 1\ 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 \
1
to \
2
against
some constant external pressure 1 = 1
ex
(see gure)
111
The force exerted on a gas by a piston is equal to the external pressure times the
area of the piston: 1 = 1
ex
= 1
ex
= 1,.
Recall from physics that work is the (path) integral over force: n =
R
a
2
a
1
1dr.
This can be manipulated as
n =
Z
a
2
a
1
1dr =
Z
a
2
a
1
1

|{z}
1
ex
dr
|{z}
o\
=
Z
\
2
\
1
1
ex
d\ (15.11)
If 1
ex
is independent of \ then
n =
Z
\
2
\
1
1
ex
d\ = 1
ex
Z
\
2
\
1
d\ = 1
ex
4\ (15.12)
112
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. Maximal Work: Reversible versus Irreversible changes
The value of n depends on 1
ex
during the entire expansion.
In the gure
n

=
Z
\
2
\
1
1
atm
d\ = 1
atm
(\
2
\
1
) (16.1)
113
113
and
n
1
= n
1
+n
2
. (16.2)
where
n
1
=
Z
\
.
\
1
1
atm+2W
d\ = 1
atm+2W
(\
i
\
1
) (16.3)
and
n
2
=
Z
\
2
\
.
1
atm
d\ = 1
atm
(\
2
\
i
) (16.4)
Hence it is clear that |n
1
| |n

| .
Now consider case in the gure below
The expansion is reversible. That is, there is always an intermediate equilibrium
throughout the expansion. Namely 1
gas
= 1
ex
. So,
n
rev
=
Z
\
2
\
1
1
gas
d\ (16.5)
This is the limiting case of path 1 in the previous gure. Thus n
rev
is the maxi-
mum possible work that can be done in an expansion. n
rev
= n
max
.
114
16.2. Heat Capacity
Temperature and heat are dierent.
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(1) =
d
d1
(16.6)
n.b., heat capacity is a function of 1; it is not a constant.
From this equation
d = C(t)d1. (16.7)
That is, when the temperature of a substance having a heat capacity C(t) is
changed by d1. d amount of heat energy is transferred.
The heat capacity also depends on the conditions during the temperature change,
e.g., C
\
(1) =

oq
oT

\
and C
1
(1) =

oq
oT

1
are 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
C
\ n
(1) =
1
:

d
d1

\
(16.8)
2. divide by mass to get specic heat
c
\
=
1
:

d
d1

\
(16.9)
We will discuss heat capacity more later.
115
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 elds) 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., 1, 1 independent then heat capacity is a function of 1
and 1, C(1. 1).
16.3.1. Example 1: The Ideal Gas Law
The equation of state for volume of an ideal gas is
1\ = :11. (16.10)
where 1 is the gas constant (8.315 J K
1
mol
1
) and : is the number of moles.
The ideal gas equation of state can be expressed in terms of intensive variables
only
1\
n
= 11. (16.11)
where \
n
=
\
a
.
The equation of state can also be expressed in terms of density j =
n
\
(and molar
mass :,:)
j =
:1
:11
=
`1
11
. (16.12)
116
16.3.2. Example 2: The van der Waals Equation of State
A more realistic equation of state was presented by van der Waals:
1 =
:11
\ :/

:
2
c
\
2
. (16.13)
The parameter c attempts to account for the attractive forces among the particles
The parameter / attempts to account for the repulsive forces among the particles
/ originates from hard sphere collisions (see gure):
117
In term of intensive variables
1 =
11
\
n
/

c
\
2
n
. (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
Berthelot
1 =
:11
\ :/

:
2
c
1\
2
=
11
\
n
/

c
1\
2
n
(16.15)
Dieterici
1 =
:11c

an
TT1
\ :/
=
11c

a
TT1r
\
n
/
(16.16)
Redlich-Kwang
1 =
:11
\ :/

:
2
c
_
1\ (\ :/)
=
11
\
n
/

c
_
1\
n
(\
n
/)
(16.17)
118
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 rst 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) ow.
Heat ows from high 1 to Low 1.
Temperature scales
Celsius: A relative scale based on water (1 = 0

C for melting ice and


1 = 100

C for boiling water)


Kelvin: An absolute temperature scale based on the ideal gas law. The
temperature at which (for xed \ and :) the pressure is zero is dened as
1 = 0 K
1(Kelvin) = 1(Celsius) + 273.15
119
119
Standard conditions
standard temperature and pressure (STP): 1 = 273.15 K and 1 = 1 atm.
(\
n
(STP) = 22.414 L/mol)
standard ambient temperature and pressure (SATP): 1 = 298.15 K and
1 = 1 bar. (\
n
(SATP) = 24.789 L/mol)
Diathermic wall : A wall that allows heat to ow through it.
Adiabatic wall : A wall the does not allow heat to ow through it.
Thermal equilibrium: If two systems are in contact along a diathermic wall and
no heat ows across the wall, then the systems are in thermal equilibrium.
The zeroth law of thermodynamics
Mathematical statement:
If 1

= 1
1
and 1
1
= 1
C
. then 1

= 1
C
(17.1)
This the mathematical statement of transitivity
Verbal statement: If system is in thermal equilibrium with system 1
and system 1 is in thermal equilibrium system C then system 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 as when it is in thermal contact along a diathermic
wall with system 1.
120
17.2. The First Law of Thermodynamics
Denitions:
State: the state of a system is dened by specifying a minimum number in
intensive variables
State Function: A function of the chosen independent variables that de-
scribes a property of the state (e.g., \ (1. 1)). 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 (n) and the heat supplied to the system (). The energy
of a system is called the internal energy (l) of the system.
The rst law of thermodynamics:
Mathematical statement:
4l = +n (17.2)
or in dierential form
dl = d +dn (17.3)
Verbal statement: The change in internal energy of a system is equal to the
amount of work done on the system plus the amount of heat provided to the
system.
So for a system where all the work is 1\ work the rst law becomes
4l =
Z
\
2
\
1
1
ex
d\ (17.4)
121
in dierential form this is
dl = d 1
ex
d\ (17.5)
Although l can be expressed as a function of any two state variables, the most
convenient at this time are \ and 1. l l(1. \ ).
The total dierential of l(1. \ ) is
dl =

Jl
J1

\
d1 +

Jl
J\

T
d\ (17.6)
Consider adding heat at a constant volume then
dl =

Jl
J1

\
d1 +

Jl
J\

T
d\ = d 1
ex
d\. (17.7)
So,

Jl
J1

\
d1 = d ==

Jl
J1

\
=
d
d1
= C
\
(17.8)
Hence the slope

0l
0T

\
is the heat capacity.
The other slope,

0l
0\

T
. is called the internal pressure (it has no standard symbol).
A useful relation (derivation to come) is

Jl
J\

T
= 1

J1
J1

\
1 (17.9)
Example: A van der Waals gas
1 =
:11
\ :/

:
2
c
\
2
=

J1
J1

\
=
:1
\ :/
(17.10)
122
so the useful relation becomes

Jl
J\

T
= 1
:1
\ :/
1 =
:11
\ :/

:11
\ :/
+
:
2
c
\
2
= +
:
2
c
\
2
(17.11)
The equation of state for l: Express l in terms of 1. \. and 1.
Start with the total dierential of l
dl =

Jl
J1

\
d1 +

Jl
J\

T
d\ (17.12)
but

0l
0T

\
= C
\
and

0l
0\

T
= 1

01
0T

\
1 (useful relation). Hence
dl = C
\
d1 +

1

J1
J1

\
1

d\ (17.13)
is the equation of state for l.
A useful approximation is 4l = C
\
41 which is valid for
i) heat capacity nearly constant over 41 and with no phase transitions.
ii) ideal gas or at constant volume.
123
18. The Second and Third Laws of
Thermodynamics
18.1. Entropy and the Second Law of Thermodynamics
We learned from statistical mechanics that entropy, o. is a measure of the disorder
of the system and is expressed via Boltzmanns equation o = / ln \ (where \ is
the micocanonical partition function)
We expressed Boltzmanns law in terms of the more convenient canonical partition
function as
o =

1
1
+/ lnQ. (18.1)
Now, the average energy of the system

1 is in fact what we call internal energy:
l =

1.
Furthermore we derived the simple relation between the Helmholtz free energy
and the canonical partition function as = /1 ln Q.
Hence,
o =
l
1


1
=
1
1
(l ). (18.2)
Since l. . and 1 are state functions, o is also a state function .
So we may write
do =
1
1
(dl d) (18.3)
124
124
for an isothermal process.
Recall the denition of Helmholtz free energythe 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 d = dn
rev
.
For now let us limit the discussion to reversible processes. Then
do =
1
1
(dl dn
rev
) =
1
1
(d
rev
+dn/
rev
dn/
rev
) (18.4)
=
d
rev
1
. (Reversible process)
Note: An alternative approach to thermodynamics which makes no reference to
molecules or statistical mechanics is to simply begin by dening entropy as do =
oqrev
T
The principle of Clausius
The entropy of an isolated system will always increase in a spontaneous
process
Mathematical statement: (do)
l,\
_ 0
For a general process: dl = d 1
ex
d\
For a reversible process 1
ex
= 1 and d = 1do so dl = 1do 1d\
125
Since l. o. 1. 1. and \ are state functions, dl = 1do 1d\ holds for any
process, but in general, 1do is not heat and 1d\ is not work. (see gure)
1do is heat and 1d\ is work only for reversible processes.
For some dl.
d 1
ex
d\ = 1do 1d\ =1do = d 1
ex
d\ +1d\ (18.5)
1do = d + (1 1
ex
) d\
Case i) 1
ex
1 then (spontaneous) d\ is negative so (11
ex
)d\ is positive.
Case ii) 1 1
ex
then (spontaneous) d\ is positive so (11
ex
)d\ is positive.
Case iii) 1 = 1
ex
then (spontaneous) d\ is zero so (1 1
ex
)d\ is zero.
Thus for any spontaneous process 1do _ d.
This is a mathematical statement of the second law of thermodynamics
126
18.1.1. Statements of the Second Law
Unlike the rst 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 rst law for a reversible change at constant volume.
dl = d +dn = d 1
ex
d\ (18.6)
From our earlier discussion of heat capacity d = C
\
d1 (C
\
since constant vol-
ume). So,
dl = C
\
d1 (18.7)
but also dl = 1do. So
do =
C
\
d1
1
==4o =
Z
T
2
T
1
C
\
1
d1. (18.8)
127
A very similar derivation can be done for a reversible change at constant pressure
(we can not do it quite yet) to yield
4o =
Z
T
2
T
1
C
1
1
d1 (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
4o =
Z
T
2
T
1
C
1
1
d1 (18.10)
now let 1
1
0
4o = o
2
o
0
=
Z
T
2
0
C
1
1
d1 (18.11)
Hence the mathematical statement of the third law is
o(1
2
) =
Z
T
2
0
C
1
1
d1 +o
0
(18.12)
From a macroscopic point of view o
0
is arbitrary. However, a microscopic point of
view suggests o
0
= 0 for perfect crystals of atoms or of totally symmetric molecules
(e.g., Ar, O
2
etc.). o
0
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 1 0.
For o
0
to have signicance
C
T
T
must be nite (not innite) as 1 0. Thus C
1
0.
128
But C
1
=
oq
oT
0 implies
oT
oq
.
In other words, an innitesimal amount of heat causes an innite change in tem-
perature.
In view of what we have learned about uctuations, the ever present random
uctuations in energy provide the innitesimal amount of heat and so you can
never reach absolute zero corresponding to an average energy of zero.
18.2.2. Debyes Law
Heat capacity data only goes down so far. So one needs a theoretical extrapolation
down to 1 = 0. (Debye)
Postulate: C
1n
= c1
3
. That is at low temperatures heat capacity goes as the
cube of the temperature.
C
+
1n
. 1
+
are the lowest temperature data points. So, c = C
+
1n
,1
+3
.
129
The molar entropy is
o
n
(1
+
) =
Z
T
W
0
C
1
1
d1
C
Tr
=oT
3
=
C
+
1n
1
+3
Z
T
W
0
1
2
d1 (18.13)
=
C
+
1n
1
+3
1
3
3

T
W
0
=
C
+
1n
3
.
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 rst.
The interesting thing is that each molecule in a macroscopic system obeys time
invariant dynamics. Both Newtons laws and Quantum dynamics (next semester)
are the same if you replace t with t.
Yet, the behavior of the macrosystem denitely 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
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 is
o = / ln \. (18.14)
The Boltzmann distribution :
q
i
c
oc
.
P
)
q
)
c
oc

(18.15)
The canonical partition function is
Q =
X
)
q
)
c
o1

(18.16)
131
131
The relation between the partition function and the molecular partition
function is
Q =

.
mol
`!
. (18.17)
The Translational Partition Function

trans
=
\
\
3
(18.18)
where
\ =
/
_
2::/1
(18.19)
is the thermal de Broglie wavelength.
The Rotational Partition Function (linear molecules) is

rot
-
1
oo
v
. (18.20)
where o
v
=
I
2
8
2
1I
(1 is the moment of inertia) and o is the so-called sym-
metry number in which o = 1 for unsymmetrical molecules and o = 2 for
symmetrical molecules
The Vibrational Partition Function is

vib
=
1
2 sinh
1
2
,~.
. (18.21)
The ensemble average of any property is given by

C =
1
Q
X
i
C
i
q
i
c
oc
.
. (18.22)
The relations between the canonical partition function and the thermody-
namics variables are
Helmholtz Free Energy = /1 ln Q
Internal energy l =
1
Q

0Q
0o

a,\
=

0 lnQ
0o

a,\
Entropy o = /,

0 lnQ
0o

a,\
+/ ln Q
Pressure 1 =
1
oQ

0Q
0\

a,o
=
1
o

0 lnQ
0\

a,o
132
1\ work is
dn = 1d\. (18.23)
Heat capacity:
d = C(t)d1. (18.24)
General forms of the rst law:
4l = +n. (18.25)
in dierential form this is
dl = d 1
ex
d\. (18.26)
Also,
dl = 1do 1d\. (18.27)
The second law
1do _ d. (18.28)
The third law
o(1
2
) =
Z
T
2
0
C
1
1
d1 +o
0
(18.29)
Debyes law for entropy at very low temperatures
o
n
(1
+
) =
C
+
1n
3
. (18.30)
where C
+
1n
is the molar heat capacity at the lowest temperature for which
there is data.
133
Part IV
Basics of Thermodynamics
134
134
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, 1. (which you are familiar
with from general chemistry) serves are a common point which connects thermo-
dynamics, electrochemistry, and kineticstopics 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 systemthe internal energy and, briey the Helmholtz free energy.
135
135
From the rst law as stated as
dl = 1do 1d\ (19.1)
we say that the natural (most convenient) variables for the equation of state for
l are o and \ . This is l = l(o. \ )
Unfortunately o can not be directly measured and most often 1 is a more conve-
nient variable than \
Because of this fact, it is handy to dene state functions which have dierent 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 (o and 1), (1 and \ ) and (1 and 1)
The table below lists these state functions
State function Symbol Natural variables Denition Units
Internal Energy l o and \ energy
Enthalpy H o and 1 H = l +1\ energy
Helmholtz free energy 1 and \ = l 1o energy
Gibbs free energy G 1 and 1 G = H 1o energy
We consider each of these functions in turn
19.2. Enthalpy
We want a state function whose natural variables are o and 1
Let us try the denition H = l +1\.
136
Now formally
dH = dl +d(1\ ) = dl +1d\ +\ d1. (19.2)
but dl = 1do 1d\. so
dH = 1do 1d\/ +1d\/ +\ d1 (19.3)
= 1do +\ d1.
Hence Enthalpy does indeed have the desired natural variables.
19.2.1. Heuristic denition:
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 1\ work
done. If the process occurs at constant pressure then the enthalpy change is the
heat given o or taken in.
For example, consider an reversibly expanding gas under constant pressure (d1 =
0) and adiabatic (d = 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 ow 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 changeit is the internal energy adjusted by an amount of energy equal
to the 1\ work done by the system. As Freshmen we learn this as 4H =
j
.
19.3. Helmholtz Free Energy
Now we want a state function whose natural variables are 1 and \
137
Let us try the denition = l 1o.
Formally
d = dl d(1o) = dl 1do od1. (19.4)
but dl = 1do 1d\. so
d = 1do / 1d\ 1do / od1 (19.5)
= 1d\ od1.
Hence Helmholtz free energy does indeed have the desired natural variables.
19.3.1. Heuristic denition:
As we have said before Helmholtz free energy is the energy of the system which is
available to do workIt 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 1 and 1
Let us try the denition G = H 1o.
Now formally
dG = dH +d(1o) = dH 1do od1. (19.6)
but from above dH = 1do +\ d1. so
dG = 1do / +\ d1 1do / od1 (19.7)
= \ d1 od1.
Hence Gibbs free energy does indeed have the desired natural variables.
138
19.4.1. Heuristic denition:
Gibbs free energy is the energy of the system which is available to do non 1\
workIt 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 1\ work.
19.5. Heat Capacity of Gases
19.5.1. The Relationship Between C
1
and C
\
To nd how C
1
and C
\
are related we begin with
dH = 1do +\ d1 (19.8)
at constant pressure and reversible conditions
dH = 1do (19.9)
dH = d
but
d = C
1
d1 (19.10)
The constant pressure heat capcity can then be expressed in terms of enthalpy as
C
1
=

JH
J1

1
. (19.11)
So,
C
1
=

J (l +1\ )
J1

1
=

Jl
J1

1
+1

J\
J1

1
(19.12)
note

0l
0T

1
is not C
\
we need

0l
0T

\
. Use an identity of partial derivatives

Jl
J1

1
=

Jl
J1

\
+

Jl
J\

J\
J1

1
(19.13)
139
thus
C
1
=

Jl
J1

\
+

Jl
J\

J\
J1

1
+1

J\
J1

1
(19.14)
= C
\
+

J\
J1

Jl
J\

T
+1

.
Recall the expression for internal pressure

0l
0\

T
= 1

01
0T

\
1. Then
C
1
= C
\
+

J\
J1

1

J1
J1

\
1/ +1/

(19.15)
Finally
C
1
= C
\
+1

J\
J1

J1
J1

\
(19.16)
Example: Ideal gases
1. Ideal gas (equation of state: 1\ = :11): This equation is easily made
explicit in either 1 or \ so we dont need any of the above replacements
C
1
= C
\
+1

J\
J1

J1
J1

\
(19.17)
= C
\
+1
:1
1
:1
\
=
:11
1\
:1
Thus C
1
= C
\
+:1 or
C
1n
= C
\ n
+1 (19.18)
19.6. The Maxwell Relations
Summary of thermodynamic relations weve seen so far
Denitions and relations:
H = l +1\
140
= l 1o
G = H 1o
C
\
=

0l
0T

\
. C
1
=

01
0T

1
basic equations Maxwell relations working equations
dl = 1do 1d\

0T
0\

S
=

01
0S

\
dl = C
\
d1 +

1

01
0T

\
1

d\
dH = 1do +\ d1

0T
01

S
=

0\
0S

1
dH = C
1
d1

1

0\
0T

1
\

d1
d = 1d\ od1

0S
0\

T
= +

01
0T

\
do =
C
1
T
d1 +

01
0T

\
d\
dG = \ d1 od1

0S
01

T
=

0\
0T

1
do =
C
T
T
d1

0\
0T

1
d1
We will get plenty of practice with derivations based on these equations and on
the properties of partial derivatives. (See handout and Homework)
141
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 dierential of is ( = l 1o)
d = dl 1do od1 = d 1
ex
d\ 1do od1 (20.1)
For constant 1 and \. (d)
T,\
= d 1do
From the second law, 1do _ d for a spontaneous process, (d)
T,\
_ 0 for a
spontaneous process.
Hence at equilibrium (d)
T,\
= 0.
For chemistry it is most often more convenient to use Gibbs free energy
The total dierential of G is
dG = dH 1do od1 = d 1
ex
d\ +1d\ +\ d1 1do od1 (20.2)
142
142
For constant 1 and 1 = 1
ex
. (dG)
T,1
= d 1do
Again from the second law, 1do _ d for a spontaneous process, (dG)
T,1
_ 0 for
a spontaneous process.
Hence at equilibrium (dG)
T,1
= 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 to
equilibrium.
The free energy functions are the workhorses of applied thermodynamics so we
want to get a feel for them.
Returning to the total dierentials of free energy,
d = dl 1do od1 (20.3)
and
dG = dH 1do od1. (20.4)
Expressing dl and dH generally as dl = 1do 1d\ and dH = 1do + \ d1
(remember that in general 1do cannot be identied with d and 1d\ cannot be
identied with n).
Plugging these into the total dierentials of free energy gives
d = od1 1d\ (20.5)
and
dG = od1 +\ d1 (20.6)
143
These expressions are quite general, but i) only 1\ work and ii) closed systems.
The total dierential of is also
d = d +dn 1do od1. (20.7)
For a reversible process d = 1do and work is maximal.
Hence (d)
T
= dn
max
==(4)
T
= n
max
. As we have stated in words a number
of times before.
The total dierential of G is also
dG = d +dn +1d\ +\ d1 1do od1. (20.8)
In general dn = dn
0
1
ex
d\ where dn
0
is the non-1\ work.
The total dierential of G becomes
dG = d +dn
0
1
ex
d\ +1d\ +\ d1 1do od1. (20.9)
For constant 1 and 1 = 1
ex
. (dG)
T,1
= d +n
0
1do.
For reversible processes = 1do and this becomes
(dG)
T,1
= dn
0
max
==(4G)
T,1
= n
0
max
(20.10)
So, as stated earlier, the Gibbs free energy is the energy of the system available
to do non-1\ work.
20.2. Chemical potential
What if the amount of substance can change?
144
Extensive properties depend on the amount of stu
For example (1. \ ) now becomes (1. \. :) and the total dierential becomes
d(1. \. :) =

J
J1

\,a
d1 +

J
J\

T,a
d\ +

J
J:

\,T
d: (20.11)
Lets focus on the slope

0
0a

\,T
.
This is a measure of the change in Helmholtz free energy of a system (at
constant 1 and \ ) with the change in the amount of material.
Physically, this is a measure of the potential to change the amount of mate-
rial.
It denes the chemical potential j =

0
0a

\,T
.
So we can also write
d = od1 1d\ +jd: (20.12)
What about the relation of the chemical potential to Gibbs free energy?
G = H 1o = l 1o
| {z }
=
+1\ = +1\ so,
dG = d+1d\ +\ d1 (20.13)
= od1 1/ d\/ +1/ d\/ +\ d1 +jd:
= od1 +\ d1 +jd:.
but from
dG =

JG
J1

1,a
d1 +

JG
J1

T,a
d1 +

JG
J:

1,T
d: (20.14)
145
we see that
j =

JG
J:

1,T
. (20.15)
So, j is also a measure of the change in Gibbs free energy of a system (at constant
1 and 1) with the change in the amount of material and it still has the same
physical meaning.
The Gibbs free energy per mole (G
n
) for a pure substance is equal to the chemical
potential. (G
n
= j)
20.3. Activity and the Activity coe!cient
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
coe!cient.
Activity is hard to dene in words and indeed it has an awkward mathematical
denition as we will soon see.
The activity coe!cient has a more convenient denition 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 denition of activity c
i
of some species i is implicitly stated as
lim

c
i
q()
= 1 (20.16)
where q() is any reference function (e.g., pressure, mole fraction, concentration
etc.), and

is the value of at the reference state.


146
This implicit denition is awkward so for convenience one denes the activity
coe!cient as the argument of the above limit,

i
=
c
i
q()
(20.17)
which we can rearrange as
c
i
=
i
q(). (20.18)
The denition of activity implies that
i
= 1 at q(

) (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 dene 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 (1 = :11,\ ) as our reference function and the
reference state being when 1 = 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
Our reference function is very simple: q() = = 1, so
=
c
1
=c = 1. (20.19)
Thus the activity of our real gas is given by the activity coe!cient times the
pressure of an ideal gas under the same conditions.
Based on the condition that 1 as we approach the reference state (1 = 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 denition).
j
i
j

i
= 11 lnc
i
. (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
j
i
= j

i
+11 ln c
i
(20.21)
and for the ideal state
j
io
i
= j

i
+11 ln c
io
i
=j

i
= j
io
i
11 ln c
io
i
. (20.22)
148
Thus,
j
i
= j
io
i
11 lnc
io
i
+11 ln c
i
(20.23)
j
i
j
io
i
= +11 lnc
i
11 lnc
io
i
= 11 ln
c
i
c
io
i
Example: Real and ideal gases at constant temperature, but any pressure.
Starting from the begining
dj
io
= dG
n
= o
n
=0
z}|{
d1 +\
n
d1 (20.24)
dj
io
= \
n
d1
dj
io
=
11
1
d1.
Now we integrate from the reference state to the current state of interest
Z
j

dj
io
=
Z
1

11
1
d1. (20.25)
This gives
j
io
j

= 11 ln
1
1

. (20.26)
The usual standard state is the ideal gas at 1

= 1. so
j
io
= j

+11 ln 1. (20.27)
(Note that as 1 1. j
io
j

).
Lets say our gas is not ideal, then at a given pressure
j = j

+11 ln c. (20.28)
For gases activity is usually called fugacity and given the symbol ,, so c = , for
real gasses. Thus
j = j

+11 ln ,. (20.29)
149
Lets say that instead of referencing to the ideal gas at 1 = 1, we want to reference
to the ideal gas at the current pressure 1.
This is easily done by using j

= j
io
11 ln 1 in the above equation for j.
j = j
io
11 ln1 +11 ln ,
j = j
io
+11 ln
,
1
.
Example: The barometric equation for an ideal gas.
We have an ideal gas so,
j
io
= j

+11 ln1 (20.30)


where we will take the reference state to be at sea level, i.e. 1

= 1 atm.
So at sea level
j
io
(0) = j

+
=0
z }| {
11 ln1 = j

(20.31)
and at elevation /
j
io
(/) = j

+11 ln 1
I
(20.32)
The gas elds the gravitational force which gives it a potential energy per mole
of `q/ at height /. We add this energy per mole term to the chemical potential
(which is free energy per mole) thus at equilibrium
j
io
(0) = j
io
(/) +`q/ (20.33)
Referencing to the reference state we get
j

/ = j

/+11 ln 1
I
+`q/ (20.34)
11 ln 1
I
= `q/
1
I
= c
3LI
TT
The last line is the barometric equation and it shows that pressure is exponentially
decreasing function of altitude.
150
21. Equilibrium
First let us consider the equilibrium 1.
Since and 1 are in equilibrium their chemical potentials must be equal
j

= j
1
(21.1)
Now,
j

= j

+11 ln c

(21.2)
and
j
1
= j

1
+11 lnc
1
So the equilibrium condition becomes
j

+11 ln c

= j

1
+11 lnc
1
(21.3)
4j

= j

1
= 11 ln c
1
11 lnc

4j

= 11 ln
c
1
c

Since chemical potential is free energy per mole, if we multiply the above by :
moles we have
4G

= :11 ln
c
1
c

as a consequence of the equilibrium condition.


The quantity
o
T
o
/
denes the equilibrium constant, 1
o
, for this process.
151
151
Say the system 1 is not in equilibrium then we can not write
j

= j
1
but we can write
j

+
4j
z }| {
j
1
j

= j
1
(21.4)
Proceeding as above we get
j

+11 lnc

+4j = j

1
+11 ln c
1
(21.5)
4j = j

1
j

+11 ln c
1
+11 ln c

4j = 4j

+11 ln
c
1
c

.
Again multiplying by : gives
4G = 4G

+:11 ln
c
1
c

.
If the 4G < 0 then the transition 1 proceeds spontaneously as written.
Consider a more complicated equilibrium
c+/1 cC +d1. (21.6)
The equilibrium condition is
cj

+/j
1
= cj
C
+dj
1
. (21.7)
In a manner similar to the above
cj

+c11 lnc

+/j

1
+/11 ln c
1
= cj

C
+c11 ln c
C
+dj

1
+d11 lnc
1
(21.8)
Rearranging gives
=4rxn G

z }| {
cj

+/j

1
cj

C
dj

1
= 11 ln
c
c
C
c
o
1
c
o

c
b
1
(21.9)
152
the equilibrium constant is
1
o
=
c
c
C
c
o
1
c
o

c
b
1
==4G

= 11 ln1
o
(21.10)
Note: : is absent in the above since the molar values are implied by the stoi-
chiometry.
21.0.3. Equilibrium constants in terms of 1
C
Equilibrium constant in terms of a condensed phase concentration:
1
0
C
=
[C]
c
[1]
o
[]
o
[1]
b
. (21.11)
which is related to 1
o
by
1
o
= 1
0
C

c
C

o
1

b
1

. (21.12)
If the reactants are solutes then as the solution is diluted all the activity coe!cients
go to unity and 1
0
C
1
o
.
21.0.4. The Partition Coe!cient
Up to now we have only considered miscible solutions.
We now consider the problem of determining the equilibrium concentrations of a
solute in both phases of an immiscible mixture.
153
The equilibrium equation is

c

o
(21.13)
The equilibrium expression for this process is
4G
co
= 0 = 4G

co
:11 ln 1
o
. (21.14)
where, 4G

co
= G

o
G

c
. The equilibrium constant for this process has a special
name; it is called the partition coe!cient, 1
oc
= 1
oc
part
, for species in the c,
mixture.
We can solve for the partition coe!cient to yield
1
oc
=
c
o

c
c

= c

4C

o<c
nTT
. (21.15)
For low concentrations
1
oc
'
[]
o
[]
c
. (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
For most drugs
0 < 1
ow
part
< 4 (21.17)
Partition coe!cient Delivery mechanism
low 1
ow
part
(likes water) injection
medium 1
ow
part
oral
high 1
ow
part
(likes oil) skin patch/ointment
Factors other than the partition coe!cient inuence the drug delivery choice. For
example, can the drug handle the acidic environment of the stomach?
155
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 nal product
thermodynamic state that is in general dierent from the initial reactant state.
For any extensive property
4
rxn
(Property) = property of products property of reactants
Example
Reaction: cA+/B= cC+dD
4
rxn
o = co
n,C
+do
n,D
co
n,A
/o
n,B
22.1. Heats of Reactions
Exothermic reaction: heat is given o to the surroundings
Endothermic reaction: heat is given taken in from the surroundings
At constant pressure (1
ca
= 1
= 4
rxn
l n = 4
rxn
l 14
rxn
\ = 4
rxn
H (22.1)
4
rxn
H < 0 for Exothermic reactions.
4
rxn
H 0 for Endothermic reactions.
156
156
22.1.1. Heats of Formation
Hesss Law of heat summation: 4
rxn
H is independent of chemical pathway
Example: C
2
H
2
+H
2
= C
2
H
4
.
This direct reaction is not easy but it can be done in steps
C
2
H
2
+
5
2
O
2
2CO
2
+H
2
O(liq) 4
rxn
H

= 1299.63 kJ
2CO
2
+2H
2
O(liq)C
2
H
4
+ 3O
2
4
rxn
H

= +1410.97 kJ
H
2
+
1
2
O
2
H
2
O(liq) 4
rxn
H

= 285.83 kJ
C
2
H
2
+H
2
= C
2
H
4
4
rxn
H

= 174.49 kJ
The heat of formation 4
)
H

is the 4
rxn
H at STP in forming a compound from
its constituent atoms in their natural states.
O
2
. H
2
. C(graphite) are examples of atoms in their natural state.
Example: Formation of water
H
2
+
1
2
O
2
= H
2
O not 2H
2
+O
2
= 2H
2
O
4
rxn
H =
P
i
i
i
4
)
H(i), where i
i
is the stoichiometric factor of the i
th
com-
ponent.
Example: H
2
O(liq)H
2
O(gas) at SATP
H
2
+
1
2
O
2
= H
2
O(gas) 4
)
H

= 241.818 kJ
H
2
+
1
2
O
2
= H
2
O(liq) 4
)
H

= 285.830 kJ
H
2
O(liq)H
2
O(gas) 4
rxn
H

= 241.818 (285.830) = 44.012 kJ


22.1.2. Temperature dependence of the heat of reaction
4
rxn
H(1
2
) = 4
rxn
H(1
1
) +
Z
T
2
T
1
4
rxn
C
1
d1 (22.2)
157
22.2. Reversible reactions
Recall the requirement for a spontaneous change: 4G < 0 for constant 1 and 1.
4
rxn
G = G(products) G(reactants) =
X
i
i
i
j
i
. (22.3)
(remember j
i
= G
n,i
for pure substance i).
As we saw before j
i
can be dened in terms of activity
j
i
= j

i
+11 ln c
i
. (22.4)
So,
4
rxn
G =
4
rxn
G

z }| {
X
i
i
i
j

i
+11
X
i
i
i
lnc
i
. (22.5)
Using the property of logarithms: c lnr + / ln = ln(r
o

b
) the above expression
becomes
4
rxn
G = 4
rxn
G

+11 ln
Y
i
c
i
.
i
(22.6)
4
rxn
G = 4
rxn
G

+11 ln Q.
where Q =
Q
i
c
i
.
i
is the activity quotient.
At equilibrium, 4
rxn
G = 0 and Q = 1
o
(Thermodynamic equilibrium constant).
1
o
depends on 1 but is independent of 1.
For the reaction cA + /B = cC + dD
1
o
=
c
c
C
c
o
D
c
o
A
c
b
B
(22.7)
Note that the activity of any pure solid or liquid is for all practical purposes
equal to 1.
158
For ideal gases, c
i
=
1
.
1

=
A
.
1
1

(1

= 1 bar) This leads to the sometimes


useful relation
1
1
=
1
c
C
1
o
D
1
o
A
1
b
B
=
(1

c
C
)
c
(1

c
D
)
o
(1

c
A
)
o
(1

c
B
)
b
= 1
o

c+oob
. (22.8)
or more generally 1
1
= 1
o
(1

)
S
.
i
.
.
So at equilibrium, 4
rxn
G = 4
rxn
G

+11 ln Q becomes
0 = 4
rxn
G

+11 ln1
o
=4
rxn
G

= 11 ln 1
o
. (22.9)
22.3. Temperature Dependence of 1
d
Starting with G = H 1o or G,1 = H,1 o.
From this

0(GT)
0(1T)

1
= H.
Applying this to
4
rxn
G

1
=
4
rxn
H

1
4
rxn
o (22.10)
gives

J(4
rxn
G

,1)
J(1,1)

1
= 4
rxn
H

(22.11)
Using 4
rxn
G

= 11 ln1
o
. we get

J ln1
o
J(1,1)

1
ind.
=
of 1
d ln1
o
d(1,1)
=
4
rxn
H

1
(22.12)
or (using
o
o(1T)
=
oT
o(1T)
o
oT
= 1
2 o
oT
)
d ln1
o
d1
=
4
rxn
H

11
2
(22.13)
Integration gives
ln 1
o
(1
2
) = ln 1
o
(1
1
) +
1
1
Z
T
2
T
1
4
rxn
H

n
1
2
(22.14)
For a reasonably small range 1
2
1
1
this is well approximated by
ln 1
o
(1
2
) = ln 1
o
(1
1
)
4
rxn
H

n
1

1
1
2

1
1
1

(22.15)
159
22.4. Extent of Reaction
There are other equilibrium constants that are used in the literature.
From 1
)
= A
)
1, 1
A
= 1
1
1
4u

From :
)
= 1
)
\
1T
(ideal gas approximation), 1
a
= 1
1

1T
\

4u
From concentration C
)
=
a

\
=
1

1T
. 1
C
= 1
1
(11)
4u
Equilibrium constants
constants expression relation to 1
o
situation used
1
o
activity(products)
activity(reactants)
when an exact answer is needed
1
1
partial pressure(products)
partial pressure(reactants)
1
a
1

34r
gas reactions
1
A
mole fraction(products)
mole fraction(reactants)

1a
1

34r

1
34r
when eq. 1 is known
1
a
moles(products)
moles(reactants)

1
a
1

34r

1T
\

4u
when \ is known and constant
1
C
concentration(products)
concentration(reactants)

1
a
1

34r

(11)
4u
when concentration known
160
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
`

+
A

3

+
`
:
+
(aq) +

A
:
3
(aq). (23.1)
where
+
(

) is the number of cations (anions) and .


+
(.

) 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:
j
salt
=
+
j
+
+

(23.2)
To determine the activity we start with
lnc
)
=
j
)
j

)
11
. , = + or (23.3)
and
lnc
salt
=
j
salt
j

salt
11
. (23.4)
161
161
Substituting the expression for j
salt
into this gives
lnc
salt
=

+
j
+

+
+
j

11
(23.5)
=

+
j
+

+
j

+
11
| {z }

+
lno
+
+

11
| {z }

3
lno
3
So,
lnc
salt
=
+
ln c
+
+

lnc

(23.6)
or, alternatively,
c
salt
= c

+
c

3
(23.7)
It is the case that 1 mole of salt behaves like =
+
+

moles of nonelectrolytes
in terms of the colligative properties. This suggests that the interesting quantity
is
j
salt

:
j
salt

=
j

salt

+11 ln c
1
salt
. (23.8)
We see that
c
1
salt
= (c

+
c

3
)
1
= c

. (23.9)
The quantity c

is the mean ionic activity.


23.1.1. Ionic activity coe!cients
The activity coe!cients for ionic solutions can also be dened via
c
+
=
+
:
+
. c

. (23.10)
where :
+
=
+
: and :

:.
The mean ionic activity coe!cient is

= (

+
+

)
1
. (23.11)
162
The quantities c
+
. c

.
+
and

cannot be measured individually.


One can use the colligative properties to measure the ionic activity coe!cients.
It is convenient to redene the osmotic coe!cient as
c =
1000 g/kg
:`
1
lnc
1
. (23.12)
where the subscript 1 refers to the solvent.
Similarly freezing point depression is redened as
o = c1
)
:. (23.13)
So, c corresponds to the empirical factor i discussed earlier.
Recall how was calculated from the Gibbs-Duhem equation:
ln

= ,
Z
n
0
,
:
0
d:
0
. (23.14)
where , = 1 c.
23.2. Theory of Electrolytic Solutions
Ionic strength is dened as
1 =
1
2
X
i
.
2
i
:
i
. (23.15)
where . is the charge of the ion and : its concentration.
Results from DebyeHckel theory: point charge in a continuum
The DebyeHckel equation:
ln

=
c|.
+
.

|
_
1
1 1c
0
_
1
. (23.16)
163
where
c =
c
3
(/1)
32

2:j

1
1000

12
. (23.17)
1 =
8:1c
2
j

1000/1
. (23.18)
c
0
is the radius of closest approach, c is the charge on the electron, j

is the
density of the pure solvent, is the dielectric constant for the pure solvent and 1
is Avogadros number.
Notice that the parameters c and 1 depend only on the solvent.
One important approximation to this equation is to neglect the 1 term to get the
DebyeHukel limiting Law (DHLL):
ln

= c|.
+
.

|
_
1. (23.19)
This gives the dependence of ln

for dilute solutions (: 0). It is seen that


the DHLL correctly predicts the
_
: dependence of ln

, which is observed ex-


perimentally (recall 1 =
1
2
P
i
.
2
i
:
i
).
A useful empirical approximation is to set 1c
0
= 1 and to add an empirical
correction to get the :
ln

=
c|.
+
.

|
_
1
1
_
1
+ 2,:

2
+
+
2

+
+

. (23.20)
This equation works well to ionic strengths of about 1 = 0.1
23.3. Ion Mobility
Current, 1 is given by the rate of change (in time) of charge, Q:
1 =
dQ
dt
(23.21)
164
(Electrical) work, n. is required to move a change through a potential (or voltage),
:
n = Q (23.22)
Power is given by the product of the voltage and the current:
j = 1 (23.23)
Resistance is given by the ratio of the voltage to current:
1 =

1
Conductance is the inverse of the resistance (1
1
).
Some relevant constants
charge of an electron c = 1.602177 10
19
C.
Faradays constant F = 1c = 96485 C/mol (Avogadros number of electrons)
23.3.1. Ion mobility
165
The total current passing through an ionic solution is determined by the sum of
the current carried by the cations and by the anions
1 = 1
+
+1

(23.24)
Now
1
i
=
dQ
i
dt
= |.
i
| c
d`
i
dt
. (23.25)
where i = +. .
For uniform ion velocity (
i
) the number of ions arriving at the electrode during
any given time interval 4t is
4`
i
=
`
i
\

i
4t ==
d`
i
dt
=
`
i
\

i
(23.26)
so
1
i
= |.
i
| c
`
i
\

i
(23.27)
Recall Coulombs law
1
i
= .
i
c1. (in vacuum) (23.28)
where 1 is the electric eld, 1 =
o.
oa
.
Also recall Newtons law
1
i
= :c
i
= :
d
i
dt
= .
i
c1. (23.29)
The moving ions experience a viscous drag , that is proportional to their velocities.
So the total force on the ions is a sum of the Coulomb force and the viscous drag
1
i
= .
i
c1 ,
i
(in solution). (23.30)
The ions quickly reach terminal velocity, i.e., the viscous drag equals the Coulomb
force. Hence 1
i
= 0.
.
i
c1 = ,
i
==
i
=
.
i
c1
,
. (23.31)
The drag , has three basic origins.
166
1. Stokes Law type force
spherical ion moving through a continuous medium
this contribution is independent of the other ions
2. Electrophoretic eect.
oppositely charged ions pull at each other
3. Relaxation eects
solvation shell must re-adjust as ion moves. a dressed ion.
167
A more fundamental quantity than ion velocity is the ion mobility, n
i
which is the
ions velocity per eld,
n
i
=

i
1
. (23.32)
For the case for parallel plate capacitors 1 =
.
|
. where | is the separation of the
plates. So,
n
i
=

i
|

. (23.33)
Here the current carried by ion i is
1
i
= |.
i
| c
`
i
\

n
i

|
. (23.34)
Suppose a salt has a degree of dissociation c (c = 1 for strong electrolytes) to
produce i
+
cations and i

anions, then each mole of salt gives: `


+
= ci
+
1:
and `

= ci

1:.
The current then becomes
1
i
= |.
i
| c
ci
i
1:
\

n
i

|
F=1c
= ci
i
:|.
i
| n
i
F

\ |
(23.35)
It is of interest to determine the ratio of the current carried by the cation versus
the anion.
1
+
1

=
c / i
+
: / |.
+
| n
+
/ F /
.
\ |
/
c / i

: / |.

| n

/ F /
.
\ |
/
=
=1
z }| {
i
+
|.
+
|
i

|.

|
n
+
n

=
n
+
n

(23.36)
Thus the ratio of the currents is determined by simply the ratio of the mobilities.
168
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 ionsolvent 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 4G
vs
= G
ion in solv.
G
ion in vac
.
Since Gibbs free energy corresponds to non-1\ work, 4G
vs
can be determined
by calculating the reversible work done in transferring an ion into the bulk of the
solvent.
169
169
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

s
170
4G
vs
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 gure)
Begin with the state in which the charged sphere (the ion) is in a vacuum.
Determine the work, n
dis
. done in discharging the sphere.
Assume the uncharged sphere can pass from the (neutral) vacuum to the
neutral solvent without doing any work, n
tr
= 0. (This is an approximation).
Determine the work, n
ch
. done in charging the sphere which is now in the
solvent.
171
So,
4G
vs
= n
dis
+n
tr
+n
ch
= n
dis
+n
ch
(24.1)
Work done in discharging the sphere:
The act of discharging a sphere involves bringing out to innity from the surface
innitesimal 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
This is expressed mathematically as
n
dis
=
Z
0
:c
Z
o
v
.
o
4:c
0
:
2
d:do (24.2)
=
Z
0
:c
o
4:c
0
:
i
do
=
(.c)
2
8:c
0
:
i
.
where . is the oxidation state of the ion, c is the charge of the electron, :
i
is the
radius of the sphere (ion) and c
0
is the permittivity of free space.
Work done in charging the sphere:
The only dierence in charging the sphere is that the sign of the work will be dier-
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,
n
ch
= +
(.c)
2
8:c
0

c
:
i
(24.3)
24.1.1. Free Energy of Solvation for the Born Model
Combining the above two expression for work gives
4G
vs
=
(.c)
2
8:c
0
:
i
+
(.c)
2
8:c
0

c
:
i
(24.4)
=
(.c)
2
8:c
0
:
i

c
1

The above expression is 4G


vs
,ion. For : moles of ions (:1 = `)
4G
vs
=
` (.c)
2
8:c
0
:
i

c
1

(24.5)
173
The dielectric constant of any solvent is always greater than unity so
1
.
s
1 is
always negative hence 4G
vs
< 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 c and , having dielectric constants

c
and
o
.
Since Gibbs free energy is a state function we can write the change in free energy
for transfer of an ion form the , phase to the c phase as
4G
oc
=
=4G
v<c
z }| {
4G
ov
+4G
vc
(24.6)
=
` (.c)
2
8:c
0
:
i

o
1

+
` (.c)
2
8:c
0
:
i

c
1

=
` (.c)
2
8:c
0
:
i

c

1

The Partition Coe!cient


We can now write the partition coe!cient for the Born model as
1
co
i
= c
34C

c<o
nTT
= c

1(:c)
2
8rr
.

0
TT

1
so

1
s
c

(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
From

JG
J1

1
= o =

J4G
vs
J1

1
= 4o
vs
. (24.8)
we nd entropy to be
4o
vs
=
J
J1
"
` (.c)
2
8:c
0
:
i

c
1

#
. (24.9)
The only variable in the above equation that has a temperature dependence is the
dielectric constant of the solvent so,
4o
vs
=
` (.c)
2
8:c
0
:
i
J
J1

=
` (.c)
2
8:c
0
:
i

2
c
J
c
J1
. (24.10)
Enthalpy is obtained via the relation:
4H
vs
= 4G
vs
+14o
vs
(24.11)
=
` (.c)
2
8:c
0
:
i

c
1

+
` (.c)
2
1
8:c
0
:
i

2
c
J
c
J1
=
` (.c)
2
8:c
0
:
i

c
+
1

2
c
J
c
J1
1

24.2. Corrections to the Born Model


The Born model is very valuable because of its simplicityqualitative 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
1. The solvophobic eect: a cavity must form in the solvent to accommodate
the ion.
2. Changes in solvent structure: the local environment of the ion has a dierent
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. Specic interactions: any interaction energy specic 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
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 = l +1\ dH = 1do +\ d1
= l 1o d = od1 1d\
G = H 1o dG = od1 +\ d1
The chemical potential equation
j
i
= j

i
+11 ln c
i
(25.1)
The 4G equation (this should be posted on your refrigerator)
4G = 4G

+11 lnQ. (25.2)


177
177
At equilibrium 4G = 0 and
4G

= 11 ln 1
o
(25.3)
For an ideal gas
C
1n
= C
n
+1 (25.4)
The DebyeHukel limiting Law (DHLL):
ln

= c|.
+
.

|
_
1. (25.5)
The ratio of the current carried by the cation versus the anion in terms of
ion mobility is
1
+
1

=
n
+
n

(25.6)
The chemical potential equation
j
i
= j

i
+11 ln c
i
(25.7)
The 4G equation (this should be posted on your refrigerator)
4G = 4G

+11 lnQ. (25.8)


At equilibrium 4G = 0 and
4G

= 11 ln 1
o
(25.9)
4G for the Born model:
4G
vs
=
` (.c)
2
8:c
0
:
c

c
1

(25.10)
4G for transfer of an ion form the , phase to the c phase,
4G
oc
=
` (.c)
2
8:c
0
:
i

c

1

(25.11)
178
Chemistry 352: Physical
Chemistry II
179
179
Part V
Quantum Mechanics and
Dynamics
180
180
26. Particle in a 3D Box
We now return to quantum mechanics and investigate some of the important
models that we omitted from the rst 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, \ (r). is given by
\ (r) =
;
A
?
A
=
r _ 0
0 0 < r < c
r _ c
. (26.1)
Because of the innities at r = 0 and r = c, we need to partition the r-axis into
the three regions shown in the gure.
181
181
Now, in region I and III, where the potential is innite, the particle can never
exist so, must equal zero in these regions.
The particle must be found only in region II.
The Schrdinger equation in region II is (\ (r) = 0)

H = 1 ==
~
2
2:
d
2
(r)
dr
2
= 1. (26.2)
The general solution of this dierential equation is
(r) = sin/r +1cos /r. (26.3)
where / =
q
2n1
~
2
.
Now must be continuous for all r. Therefore it must satisfy the boundary
conditions (b.c.): (0) = 0 and (c) = 0.
From the (0) = 0 b.c. we see that the constant 1 must be zero because
cos /r|
a=0
= 1.
So we are left with (r) = sin /r for our wavefunction.
182
The second b.c., (c) = 0. places certain restrictions on /.
In particular,
/
a
=
::
c
. : = 1. 2. 3. . (26.4)
The values of / are quantized. So, now we have

a
(r) = sin
::r
c
. (26.5)
The constant is the normalization constant.
Solving for gives
=
r
2
c
. (26.6)
Thus our normalized wavefunctions for a particle in a box are (in region II)

a
(r) =
r
2
c
sin
::r
c
. (26.7)
We found the energy levels to be
1
a
=
:
2
:
2
~
2
2:c
2
~=
I
2r
=
:
2
/
2
8:c
2
. (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
\ (r. . .) =
(
0. 0 < r < c. 0 < < /. 0 < . < c
. else
. (26.9)
183
Now the Schrdinger equation is

H = 1 =
~
2
2:
\
2
= 1
=
~
2
2:

J
2

Jr
2
+
J
2

J
2
+
J
2

J.
2

= 1. (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
(r. . .) =
a
(r)
j
()
:
(.). (26.11)
This lets us perform a mathematical trick which is sometimes useful in solving
partial dierential equations.
Subbing the product wavefunction into the Schrdinger equation we get
~
2
2:

J
2

:
Jr
2
+
J
2

:
J
2
+
J
2

:
J.
2

= 1
a

:
(26.12)
~
2
2:

:
J
2

a
Jr
2
+

a

:
J
2

j
J
2
+

a

j
J
2

:
J.
2

= 1
a

:
.
We now divide both sides by
a

:
to get
~
2
2:

a
J
2

a
Jr
2
+
1

j
J
2

j
J
2
+
1

:
J
2

:
J.
2

= 1. (26.13)
This equation is now of the form
,(r) +q() +/(.) = C. (26.14)
where C is a constant.
If we take the derivative with respect to r we get
d
dr

,(r) +q() +/(.) = C.


d,(r)
dr
+
dq()
dr
+
d/(.)
dr
=
dC
dr
.
d,(r)
dr
= 0. (26.15)
184
So, ,(r) is a constant. Similarly for q() and /(.)
Applying this to our Schrdinger equation means that we have converted our
partial dierential equation into three independent ordinary dierential equations,
~
2
2:
1

a
d
2

a
dr
2
= 1
a
==
~
2
2:
d
2

a
dr
2
= 1
a

a
(26.16)
~
2
2:
1

j
d
2

j
d
2
= 1
j
==
~
2
2:
d
2

j
d
2
= 1
j

j
~
2
2:
1

:
d
2

:
d.
2
= 1
:
==
~
2
2:
d
2

:
d.
2
= 1
:

:
which we recognize as the 1D particle in a box equations.
Hence we immediately have

a
=
r
2
c
sin
:
a
:r
c
. (26.17)

j
=
r
2
/
sin
:
j
:
/
.

:
=
r
2
c
sin
:
:
:.
c
and
1
a,ai
=
:
2
a
/
2
8:c
2
. (26.18)
1
j,a

=
:
2
j
/
2
8:/
2
.
1
:,a
:
=
:
2
:
/
2
8:c
2
.
The total wavefunction is
=
2
_
2
_
c/c
sin
:
a
:r
c
sin
:
j
:
/
sin
:
:
:.
c
(26.19)
and the total energy is
1 = 1
a,a
i
+1
j,a

+1
:,a
:
. (26.20)
185
Degeneracy
The 3D particle in a box model brings up the concept of degeneracy.
When :( 1) states have the same total energy they are said to be :-fold degen-
erate.
Let the 3D box be a cube (c = / = c) then the states
(:
a
= 2. :
j
= 1. :
:
= 1). (26.21)
(:
a
= 1. :
j
= 2. :
:
= 1).
(:
a
= 1. :
j
= 1. :
:
= 2)
have the same total energy and thus are degenerate.
186
27. Operators
27.1. Operator Algebra
We now take a mathematical excursion and discuss the algebra of operators.
Denitions
Function: A function, say ,, describes how a dependent variable, say . is
related to an independent variable, say r: = ,(r)
e.g., = r
2
, = sin r. etc.
Operator: An operator, say

C. transforms a function, say ,, into another
function, say q:

C,(r) = q(r).
Algebra: An algebra is a specic collection of rules applied to a set of objects
and 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 of
real numbers
187
187
Note: Commutivity is not a requirement of an algebra
+ example 1: multiplication on the set of real number is commutive:
c/ = /c
+ example 2: multiplication on the set of : : matrices is not com-
mutive: c/ 6= /c 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 c =

,. then c,(r) = q(r) =

,,(r) (27.3)
2. Addition:
if c,(r) = q(r) and

,,(r) = /(r). (27.4)
then ( c +

,),(r) = c,(r) +

,,(r) = q(r) +/(r)
3. Multiplication:
c

,,(r) = c

,,(r)

(27.5)

, c,(r) =

, ( c,(r)) .
but in general c

,,(r) 6=

, c,(r).
4. Inverse:
if c,(r) = q(r) and

,q(r) = ,(r) (27.6)
then

, = c
1
and is said to be c inverse
188
Linear operators:
A special and important class of operators
They obey all of the above properties in addition to
c(,(r) +q(r)) = c,(r) + cq(r). and
c(`,(r)) = ` c,(r). where ` 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. r: r,(r) = r,(r)


d:

d,(r) =
o
oa
,(r)


d
2
:

d
2
,(r) =

d

d,(r)

=

d

o
oa
,(r)

=
o
oa

o
oa
,(r)

=
o
2
oa
2
,(r)
:: :,(r. . .) = ,(r. . .)


\:

\,(r. . .) =

0
0a
c
a
+
0
0j
c
j
+
0
0:
c
:

,(r. . .)


\
2
:

\
2
,(r. . .) =

0
2
0a
2
+
0
2
0j
2
+
0
2
0:
2

,(r. . .)
Commutators:
We have seen that in general c

, 6=

, c. This leads to the construction of the
commutator, [. ]:
h
c.

,
i
= c

, c. (27.7)
189
If c

, =

, c, then
h
c.

,
i
= 0 and c and

, are said to commute with one another.
The eigenvalue equation:
If c,(r) = q(r) and q(r) = c,(r). then the operator equation, c,(r) = q(r)
becomes the eigenvalue equation
c,(r) = c,(r). (27.8)
The eigenvalue equation is of fundamental importance in quantum theory. We
shall see that eigenvalues of certain operator can be identied as experimental
observables.
Commuting operators and simultaneous sets of eigenfunctions.
If c,(r) = c,(r) and

, and c commute, then

,,(r) = /,(r).
The proof goes as follows: On the one hand,

, ( c,) =

, (c,) = c

,, (27.9)
because , is an eigenfunction of c.
On the other hand,

, ( c,) = c

,,

(27.10)
because

, and c commute.
Thus
c

,,

= c

,,

. (27.11)
which states that

,, is an eigenfunction of c with eigenvalue c. The only way for
this to be true is if

,, = /,.
190
27.2. Orthogonality, Completeness, and the Superposition
Principle
Theorem 1: The eigenfunctions of a Hermitian operator corresponding to dier-
ent eigenvalues are orthogonal:
Z
space

+
)

I
= 0. , 6= /. (27.12)
Theorem 2: The eigenfunctions of a Hermitian operator form a complete set
Corollary (the superposition principle): Any arbitrary function in the
space of eigenfunctions {,
i
} can be written as a superposition of these eigenfunc-
tions:
=
X
i
c
i
,
i
(27.13)
191
28. Angular Momentum
We will encounter several dierent 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 rst review the classical treatment.
28.1. Classical Theory of Angular Momentum
The classical angular momentum,

1, is given by

1 = r j (28.1)
The vector cross-product can be computed by nding the following determinant:

1 =

c
a
c
j
c
:
r .
j
a
j
j
j
:

=
1
i
z }| {
(j
:
.j
j
)c
a
+
1
z }| {
(.j
a
rj
:
)c
j
+
1
:
z }| {
(rj
j
j
a
)c
:
(28.2)
Hence,
1
a
= (j
:
.j
j
) . (28.3)
1
j
= (.j
a
rj
:
) . (28.4)
1
:
= (rj
j
j
a
) . (28.5)
Another quantity that we will nd useful is
1
2
=

1

1 = 1
2
a
+1
2
j
+1
2
:
(28.6)
192
192
28.2. Quantum theory of Angular Momentum
So, in accordance with postulate II, we replace the classical variables with their
operators. That is,

1
a
= ( j
:
. j
j
) =
~
i

J
J.
.
J
J

. (28.7)

1
j
= ( . j
a
r j
:
) =
~
i

.
J
Jr
r
J
J.

. (28.8)

1
:
= ( r j
j
j
a
) =
~
i

r
J
J

J
Jr

. (28.9)
Recall the basic commutators.

J
Jn
. n

= 1. (28.10)

J
Jn
.

= 0.
where n. = r. . or . and n 6= .
From these basic commutators one can derive
h

1
a
.

1
j
i
= i~

1
:
.
h

1
j
.

1
:
i
= i~

1
a
.
h

1
:
.

1
a
i
= i~

1
j
(28.11)
and
h

1
2
.

1
a
i
=
h

1
2
.

1
j
i
=
h

1
2
.

1
:
i
= 0 (28.12)
It is often convenient to express the angular momentum operators in spherical
polar coordinates as follows.

1
a
= i~

sin c
J
Jo
+ cot o cos c
J
Jc

. (28.13)

1
j
= i~

cos c
J
Jo
cot o sin c
J
Jc

. (28.14)
193

1
:
= i~
J
Jc
(28.15)

1
2
= ~
2

J
2
Jo
2
+ cot o
J
Jo
+
1
sin
2
o
J
2
Jc
2

(28.16)
28.3. Particle on a Ring
Consider a particle of mass j conned to move on a ring of radius 1.
The moment of inertia is 1 = j1
2
The Hamiltonian is given by

H =

1
2
:
21
=
~
2
21
d
2
dc
2
(28.17)
(note that we use d rather than J since the problem is one-dimensional).
The Schrdinger equation becomes
~
2
21
d
2

dc
2
= 1 (28.18)
Notice that this Schrdinger equation is exactly the same form as the particle in
a box. The only dierence is the boundary conditions.
The boundary condition for the particle in a box were was zero outside the box.
Now the boundary condition is that (c) = (c + 2:). The wavefunction must
by 2: periodic.
The allowable wavefunctions are

n
(c) =
;
A
?
A
=
cos :c
sin:c
c
in
. (28.19)
194
: = 0. 1. 2. 3. . . .
These wavefunctions are really the same. It will be most convenient to use

n
(c) = c
in
as our wave functions.
Plugging
n
(c) = c
in
into the Schrdinger equation gives
~
2
21
d
2
c
in
dc
2
= 1
n
c
in
(28.20)
~
2
:
2
21
c
in
= 1
n
c
in
Therefore the energy levels (the eigenvalues) for a particle in a ring are
1
n
=
~
2
:
2
21
=
:
2
/
2
8:
2
1
. (28.21)
Next we need to nd the normalization constant, .
1 =
Z
2
0

+
dc (28.22)
1 =
Z
2
0

2
c
in
c
in
dc
1 =
2
Z
2
0
dc = 2:
2
.
thus
=
r
1
2:
. (28.23)
Hence the normalized wavefunctions for a particle on a ring are
=
1
_
2:
c
in
. (28.24)
28.4. General Theory of Angular Momentum
To discuss angular momentum in a more general way it is convenient to dene
two so-called ladder operators

1
+
=

1
a
+i

1
j
(28.25)
195
and

=

1
a
i

1
j
(28.26)
We collect here the commutators of

1
+
and

1

:
h

1
:
.

1
+
i
=

1
+
=

1
+

1
:
=

1
:

1
+


1
+
(28.27)
h

1
:
.

1

i
=

=

1

1
:
=

1
:

+

1

(28.28)
Now, since

1
:
and

1
2
commute there must exist a set of simultaneous eigenfunc-
tions {
i
}

1
:

i
= :
i
(28.29)
and

1
2

i
= /
2

i
(28.30)
Physically, /~ represents the length of the angular momentum vector and :~
represents the projection onto the .-axis. (Note: for simplicity in writing we are
hiding the ~ in the wavefunctions.)
On these physical grounds we conclude |:| _ /. i.e., / sets an upper and lower
limit on :.
Lets dene the maximum value of : to be a new quantum number | = :
max
.
(Thus | _ /).
And lets dene the minium value of : to be a new quantum number |
0
= :
min
.
(Thus |
0
_ /)
Now, at least one of the eigenfunctions in the set {
i
} yields the eigenvalue :
max
(or |) when operated on by

1
:
. Lets call that eigenfunction
|
;

1
:

|
= |
|
. (28.31)
Now we can operate on both sides of this equation with

1

1
:

|
=

1

|
|
(28.32)
196
Using the commutator relation

1

1
:
=

1
:

+

1

we get

1
:

+

1

|
= |

|
(28.33)

1
:

|
+

1

|
= |

|
Bringing the second term on the left hand side over to the right hand side gives

1
:

|
= |

|
(28.34)

1
:

|
| {z }

I31
= (| 1)

|
| {z }

I31
We see that

1

|
=
|1
is in fact an eigenfunction of

1
:
(with associated eigen-
value (| 1)) and is thus a member of {
i
} .
The eigenfunction
|1
has an associated eigenvalue that is one unit less then the
maximum value.
The above procedure can be repeated : times so that

1
a

|
=
|a
provided :
does not exceed | |
0
.
The eigenfunction
|a
has an associated eigenvalue that is : units less then the
maximum value, i.e.,

1
:

|a
= (| :)
|a
. (28.35)
The largest value of : is | |
0
. For that case,

1
:

|
0 = (| | +|
0
)
|
0 = |
0

|
0 . (28.36)
Similar behavior is seen for the operator

1
+
. except in the opposite directionthe
eigenvalue is increased by one unit for each action of

1
+
. For example

1
+

1
:

|
0 =

1
+
|
0

|
0 (28.37)

1
:

1
+


1
+

|
0 = |
0

1
+

|
0

1
:

1
+

|
0 = (|
0
+ 1)

1
+

|
0 .
197
The raising and lowering nature of

1
+
and

1

is why they are called ladder


operators.
We can not act with

1
+
and

1

indenitely since we are limited by |we reach


the ends of the ladder. This requires that

|
0 = 0 (28.38)
(we cant go lower than the lowest step) and

1
+

|
= 0 (28.39)
(we cant go higher than the highest step).
Often times the ladder operators appear in tandem either as

1

1
+
or

1
+

so it
is useful list some identities for these products

1
+
=

1
2


1
2
:


1
:
(28.40)
and

1
+

=

1
2


1
2
:
+

1
:
(28.41)
We can use these identities to derive a relation between the quantum numbers /
and |.
We begin with

1
+

|
=

1

1
+

= 0. (28.42)
but from the rst of the above identities

1
+

|
=

1
2


1
2
:


1
:

|
= (/
2
|
2
|)
|
(28.43)
Therefore
/
2
|
2
| = 0 =/ =
p
|(| + 1). (28.44)
198
We we can also consider

1
+

|
0 =

1
+

|
0

= 0 (28.45)
and

1
+

|
0 =

1
2


1
2
:
+

1
:

|
0 = (/
2
|
02
+|
0
)
|
0 . (28.46)
substituting in the relation we just found for / gives
|(| + 1) |
02
+|
0
= 0; (28.47)
simplifying gives
| = |
0
(28.48)
Thus :
max
= |, :
min
= | and so : = |. | 1. | 2. . . . . | + 1 . |.
This also implies that the number of rungs is 2| +1 and that | must be either an
integer or a half-integer.
28.5. Quantum Properties of Angular Momentum
The eigenfunctions of angular momentum are entirely specied by two quantum
numbers | and ::
|n
.

1
2

|n
= |(| + 1)
|n

1
:

|n
= :
|n
(28.49)
If we write out the rst of these explicitly in spherical polar coordinates as a
partial dierential equation we obtain
J
2

|n
Jo
2
+ cot o
J
|n
Jo
+
1
sin
2
o
J
2

|n
Jc
2
+|(| + 1)
|n
= 0 (28.50)
The solutions to this partial dierential equation are known to be the spherical
harmonic functions

|n
= 1
|n
(o. c). (28.51)
199
The spherical harmonics are functions of two variables, but they are a product of
a function only of o and a function only of c.

|n
= 1
|n
(o. c) = 1
|n|
|
(o)c
in
. (28.52)
where the 1
|n|
|
(o) are the Legendra polynomials and 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
2
0
Z

0
1
+
|
0
n
0 (o. c)1
|n
(o. c) sin ododc =
(
1 |
0
= | and :
0
= :
0 |
0
6= | or :
0
6= :
(28.53)
28.5.1. The rigid rotor
Rotational energy
For general rotation in three dimensions the is

H =
~
2
21

1
2
. (28.54)
so the Schrdinger equation is

H
|n
= 1
|n

|n
=
~
2
21

1
2

|n
= 1
|n

|n
=
~
2
21
|(| + 1)
|n
= 1
|n

|n
. (28.55)
Thus
1
|n
=
|(| + 1)~
2
21
=
|(| + 1)/
2
8:
2
1
= 1
|
. (28.56)
There is no : dependence for the energy. In other words, the energy levels are
determined only by the value of |.
We know that there are 2| +1 dierent : values for a particular | value. All 2| +1
of these wavefunctions correspond to the same energy. We say the there is a 2| +1
degeneracy of the energy levels.
200
29. Addition of Angular Momentum
29.1. Spin Angular Momentum
We learned above that | may take on integer or half-integer values.
Systems in which | 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, | = : = 1,2. The values
of : = :
c
are limited to +1/2 and 1,2.
One peculiarity of this system is that the wavefunctions are 4: periodic (and 2:
antiperiodic):

c
(o) =
c
(o + 2:) (29.1)
and

c
(o) =
c
(o + 4:). (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
29.2. Addition of Angular Momentum
In atoms the are a number of sources of angular momentum: The |s and :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 (rst couple of rows of the
periodic table)
nd the total spin angular momentum o = `
c,max
, (`
c
=
P
i
:
c
.
)
nd the total orbital angular momentum 1 = `
max
, (` =
P
i
:
i
)
then J = 1 +o
2. ,, coupling
applies to higher atomic weight atoms
nd subtotal angular momentum for each electron ,
i
= |
i
+:
i
then nd total angular momentum by J =
P
i
,
i
.
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

J
1
and

J
2
(

J
1
and

J
2
could be

1 or

o angular momentum; we use

J when
we speak generally.)
202
The total angular momentum is

J
T
=

J
1
+

J
2
.
The total .-component of the angular momentum is

J
:
T
=

J
:
1
+

J
:
2
The last statement implies that the orientation quantum number of the total
system is simple the sum of that for the components
` = :
1
+:
2
(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 ` by
J
max
= `
max
= :
1max
+:
2max
= ,
1
+,
2
(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,
J
min
= |,
1
,
2
| . (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 = ,
1
+,
2
. ,
1
+,
2
1. . . . . |,
1
,
2
| + 1. |,
1
,
2
| . (29.6)
29.2.2. An Example: Two Electrons
The table below shows the total spin angular momentum o for a two electron
system
203
spin state :
c
1
:
c
1
`
S
o
c(1)c(2)
1
2
1
2
1 1
,(1),(2)
1
2

1
2
1 1
c(1),(2) +,(1)c(2) 0 0 0 1
c(1),(2) ,(1)c(2) 0 0 0 0
Counting states:
The spin degeneracy, q
S
. of the states is given by 2o + 1. In the above example
the degeneracy is q
S
= 3 for the o = 1 states and q
S
= 1 for the o = 0 states.
29.2.3. Term Symbols
We have already seen several term symbols, those being
1
o and
3
o 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 briey 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 q
1
= 21 + 1.
For historical reasons 1 values are associated with a letter like the | values of a
hydrogenic system are.
1 0 1 2 3 4 5
symbol o 1 1 1 G H
.
204
The term symbol for a particular states is constructed from the following general
template
j
S
1
J
.
Many electron atoms have term symbols associated with their states.
Rules:
1. All closed shells have zero spin and orbital angular momentums: 1 = 0,
o = 0. These states are all singlet o states, notated by
1
o
2. An electron and a hole lead to equivalent term symbols.
E.g., j
1
and j
5
have the same term symbol.
3. Hunds 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 1.
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 spinorbit interaction energy 1
SO
.
The spinorbit Hamiltonian is

H
SO
= /c
[

1

o (29.7)

H
SO
=
/c
2

J
2


1
2


o
2

.
205
where is the spinorbit coupling constant.
From the Hamiltonian the spinorbit interaction energy is
1
SO
=
/c
2
[J(J + 1) 1(1 + 1) o(o + 1)] (29.8)
206
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 satised 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 dierence between the two systems as a perturbation to the solv-
able system
Use the solvable systems wavefunctions as a zeroth order approximation to
the wavefunctions for the unsolvable system.
These wavefunctions are used to nd a rst order correction to the energy.
207
207
The rst order energy is then used to make a rst order approximation to
the 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 rst
order energy correction.
The :
tI
state energy in perturbation theory:
1
a
= 1
(0)
a
+1
(1)
a
+. . . . (30.1)
where 1
(0)
a
is the :
tI
state energy for the unperturbed (solvable) system and 1
(1)
a
is the rst order correction. This is given by
1
(1)
a
=
Z
all
space

(0)+
a

H
(1)

(0)
a
dr. (30.2)
where

H
(1)
is the rst order correction to the Hamiltonianthe perturbation.
Example: the quartic oscillator
Consider the quartic oscillator described by the potential \ (r) =
1
2
/r
2
+cr
4
where c is very small and can be treated as a perturbation.
The obvious solvable system is the harmonic oscillator:

H =
~
2
2:
d
2
dr
2
+
1
2
/r
2
. (30.3)
This has energy levels 1
a
= ~.(:+
1
2
) and wavefunctions
a
H
a
(
_
cr)c
ca
2
2
.
where c =
q
In
~
The perturbative part of the Hamiltonian is

H
(1)
= cr
4
. (30.4)
208
For example, the ground state energy correction is then calculation from
1
(1)
0
=
Z
o
o

(0)+
0

H
(1)

(0)
0
dr (30.5)
=
Z
o
o

0
c
ca
2
2
cr
4

0
c
ca
2
2
dr
= c
2
0
Z
o
o
r
4
c
ca
2
dr
=
3
_
:c
2
0
4c
5
2
.
so the rst order ground state energy for a quartic oscillator is
1
0
'
~.
2
+
3
_
:c
2
0
4c
5
2
.
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,
1
trial
. gives a trial wavefunction which is closest to the real wavefunction.
The basis for this is the variation theorem which states
1
trial
_ 1.
We will not prove this theorem here.
The trial energy is calculated by
1
trial
=
R
all
space

+
trial

H
trial
dr
R
all
space

+
trial

trial
dr
(30.6)
The trial energy is now a function of the adjustable parameter, j. that we use to
minimize the trial energy by setting
d1
trial
dj
= 0 (30.7)
209
and solving for j. (Strictly speaking we should check that we have a minimum
and not a maximum or inection point, but with reasonably good trial functions
one is pretty safe in having a minimum.)
210
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,
ring an electron down a particle accelerator.
We shall nally get to quantum dynamics in this chapter, but rst 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
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
1
and
2
separated by energy 4c =
c
2
c
1
as shown below
The states
1
and
2
are orthonormal:
Z
TLS

+
)

I
dl =
(
1 , = /
0 , 6= /
. (31.1)
where
R
TLS
dl means integration over the two level space (which is really just the
sum
P
2
i=1
).
The states
1
and
2
are eigenfunctions of the two level Hamiltonian,

H = c
1
o
1,
+c
2
o
2,
. (31.2)
where o
),
projects out the ,
th
state of the wavefunction being acted on.
212
For example let some arbitrary wavefunction = c
1
+/
2
. then

H = (c
1
o
1,
+c
2
o
2,
) (c
1
+/
2
) (31.3)
= c
1
o
1,
(c
1
+/
2
) +c
2
o
2,
(c
1
+/
2
)
= cc
1

1
+/c
2

2
Another orthonormal set of wavefunctions are the so-called left

1
=
1
_
2

1
+
1
_
2

2
(31.4)
and right

1
=
1
_
2

1
_
2

2
(31.5)
states.
We can invert above equations and solve for
1
and
2
in terms of
1
and
1

1
=
1
_
2

1
+
1
_
2

1
(31.6)
and

2
=
1
_
2

1
_
2

1
. (31.7)
213
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 .
[
a
(r. t) =
a
(r)c

.
~
1
n
t
(31.8)
We can verify this by obtaining the time independent Schrdinger equation from
the more general time dependent
i~
J[
a
(r. t)
Jt
=

H[
a
(r. t) (31.9)
i~
J
a
(r)c

.
~
1nt
Jt
=

H
a
(r)c

.
~
1nt
i~
a
(r)
Jc

.
~
1
n
t
Jt
=

H
a
(r)c

.
~
1
n
t
i~
a
(r)

i
~
1
a

.
~
1
n
t
=

H
a
(r)c

.
~
1
n
t
1
a

a
(r)c

.
~
1nt
= c

.
~
1nt

H
a
(r)
1
a

a
(r) =

H
a
(r) (31.10)
Does this mean the eigenstates are not stationary states? To determine this we
need to calculate the probability of nding the particle in the same eigenstate at
some future time. This is given by
1(r. t) =

Z
[
+
a
(r. 0)[
a
(r. t)dr

2
(31.11)
=

Z

+
a
(r)
a
(r)c

.
~
1nt
dr

2
=

.
~
1
n
t
Z

+
a
(r)
a
(r)dr

2
=

.
~
1
n
t
(1)

2
= 1.
214
so no matter what time t we check we will always nd 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
x(r. t) =
1
_
2
[
1
(r. t) +
1
_
2
[
2
(r. t) (31.12)
exposing the phase factors we get
x(r. t) =
1
_
2

1
(r)c

.
~
1
1
t
+
1
_
2

2
(r)c

.
~
1
2
t
(31.13)
Lets now track the probability of nding the particle in the same superposition
state. Similar to before we calculate
1(r. t) =

Z
x
+
(r. 0)x(r. t)

2
=

Z
1
_
2

+
1
(r) +
1
_
2

+
2
(r)

1
_
2

1
(r)c

.
~
1
1
t
+
1
_
2

2
(r)c

.
~
1
2
t

2
=

1
2
Z


+
1
(r)
1
(r)c

.
~
1
1
t
+
+
1
(r)
2
(r)c

.
~
1
2
t
+
+
2
(r)
1
(r)c

.
~
1
1
t
+
+
2
(r)
2
(r)c

.
~
1
2
t
!
dr

2
. (31.14)
The cross-terms (those of the form
+
1
(r)
2
(r) and
+
2
(r)
1
(r)) are zero when
215
integrated because the eigenfunctions are orthogonal. This leaves
1(r. t) =

Z
x
+
(r. 0)x(r. t)

2
(31.15)
=

1
2
Z

+
1
(r)
1
(r)c

.
~
1
1
t
+
+
2
(r)
2
(r)c

.
~
1
2
t

dr

2
=

1
2

.
~
1
1
t
Z

+
1
(r)
1
(r)dr +c

.
~
1
2
t
Z

+
2
(r)
2
(r)dr

2
=

1
2

.
~
1
1
t
+c

.
~
1
2
t

2
=
1
4

c
+
.
~
1
1
t
+c
+
.
~
1
2
t

.
~
1
1
t
+c

.
~
1
2
t

=
1
4

1 +c
+
.
~
(1
1
1
2
)t
+c

.
~
(1
1
1
2
)t
+ 1

=
1
2

1 + cos
(1
1
1
2
)
~
t

.
The probability of nd in the system in its original superposition states is not one
for all times t.
216
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 .
` =
s
Z
space
|
unnorm
(r. . .)|
2
drdd.. (31.16)
The normalized wavefunction:

norm
=
1
`

unnorm
. (31.17)
How to get the average value for some property,
h ci =
Z
space

+
cdrdd.. (31.18)
217
217
The Laplacian
\
2
=

J
2
Jr
2
+
J
2
J
2
+
J
2
J.
2

. (31.19)
Normalized wavefunctions for the 3D particle in a box,

a
(r) =
2
_
2
_
c/c
sin
:
a
:r
c
sin
:
j
:
/
sin
:
:
:.
c
. (31.20)
The energy levels for the 3D particle in a box,
1
a
i
,a

,a
:
=
:
2
a
/
2
8:c
2
+
:
2
j
/
2
8:/
2
+
:
2
:
/
2
8:c
2
. (31.21)
Orthonormality:
Z
space

+
)

I
=
(
1. , = /
0. , 6= /
. (31.22)
Superpostion:
=
X
i
c
i
,
i
(31.23)
Commonly used comutators of the angular momentum operators are
h

1
a
.

1
j
i
= i~

1
:
.
h

1
j
.

1
:
i
= i~

1
a
.
h

1
:
.

1
a
i
= i~

1
j
(31.24)
and
h

1
2
.

1
a
i
=
h

1
2
.

1
j
i
=
h

1
2
.

1
:
i
= 0. (31.25)
The energy levels for a particle in a ring are
1
n
=
~
2
:
2
21
=
:
2
/
2
8:
2
1
. (31.26)
The normalized wavefunctions for a particle on a ring are
=
1
_
2:
c
in
. (31.27)
218
The eigenfunctions of angular momentum are entirely specied by two quan-
tum numbers | and ::
|n
.

1
2

|n
= |(| + 1)
|n

1
:

|n
= :
|n
(31.28)
The energy levels for the rigid rotor are
1
|
=
|(| + 1)~
2
21
. (31.29)
Degeneracy for general angular momentum is
q
J
= 2J + 1. (31.30)
The rst order energy correction in pertubation theory is
1
(1)
a
=
Z
all
space

(0)+
a

H
(1)

(0)
a
dr. (31.31)
The trial energy in variation theory is calculated by
1
trial
=
R
all
space

+
trial

H
trial
dr
R
all
space

+
trial

trial
dr
(31.32)
In general
[
a
(r. t) =
a
(r)c

.
~
1
n
t
(31.33)
The left and right superposition states are

1
=
1
_
2

1
+
1
_
2

2
(31.34)
and

1
=
1
_
2

1
_
2

2
(31.35)
219
Part VI
Symmetry and Spectroscopy
220
220
32. Symmetry and Group Theory
We now take a short break from physical chemistry to discuss ideas from the
mathematical eld 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 it
belongs.
The mathematical properties of all the possible groups have been worked
out
These mathematical properties translate into a wide variety of variety of
physical properties including
Bonding
Properties of wavefunctions
Vibrational modes
Many more applications
221
221
32.1. Symmetry Operators
Any operator that leaves ||
2
invariant are symmetry operators for that particular
system:

C||
2
= ||
2
. (32.1)
This implies

C = . (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 : and o
I
.
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
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 nd 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 C
2
Group
The C
2
group consists of the symmetry elements

1.

C
2
. o

(in-plane) and o
0

(transverse).
Water is an example of a molecule described by this point group.
The multiplication table for the C
2
group is
C
2

1

C
2
o

o
0

1

1

C
2
o

o
0

C
2

C
2

1 o
0

o
0


1

C
2
o
0

o
0


C
2

1
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
Connecting with the C
2
group example lets consider the wavefunctions for water.
In this case one can collect the eigenvalues (either +1 or 1) for each of the four
symmetry operators as a four component vector. As it turns out there is four
possible sets of eigenvalueshence four dierent vectors:

1
= (1. 1. 1. 1)

2
= (1. 1. 1. 1)
1
1
= (1. 1. 1. 1)
1
2
= (1. 1. 1. 1).
To see where these four vectors come from, consider the following.
The rst value has to be +1 since the only eigenvalue of

1 is 1
The eigenvalue of

C
2
can be +1 or 1
When it is +1 the vectors are labelled
When it is 1 the vectors are labelled 1
The eigenvalue of o

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 o
0

can be either +1 or 1
Finally there is a restriction do to the fact that the eigenvalues must obey
the group multiplication table.
This restriction forces the eigenvalues of o

and o
0

to be the same for


the type vectors and opposite for the 1 type vectors.
224
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 :
C
2

1

C
2
o

o
0

1
1 1 1 1

2
1 1 1 1
1
1
1 1 1 1
1
2
1 1 1 1
+ + + See Handout on Character Tables + ++
32.3.1. Direct Products
The direct product of a two vectors is dened as
(r
1
. r
2
. r
3
. . . .) (
1
.
2
.
3
. . . .) = (r
1

1
. r
2

2
. r
3

3
. . . .) (32.3)
For the example of the C
2
group consider
1
1
1
2
= (1. 1. 1. 1) (1. 1. 1. 1)
= (1. 1. 1. 1) =
2
(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
In doing this we will, for simplicity, consider only proper rotations (C
a
). 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 casethe O group (remember we are considering only proper rotations).
The S vector has the symmetry of a sphere (r
2
+
2
+ .
2
) 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 A
1
.
The P vector is triply degenerate and has the symmetry of r. and . 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 T
1
group.
226
The D vector has a degeneracy of ve and the symmetry of 2.
2
r
2

2
. r.. ..
r. r
2

2
. Looking at the table for the O group we see the degeneracy splits:
two states become E type and the remaining three become T
2
type.
The F states have a degeneracy of 7 and the symmetry of .
3
. r.
2
. .
2
. r..
.(r
2

2
). r(r
2
3
2
) and (3r
2

2
). In an octahedral environment the states
split with one becoming A
2
, three becoming T
1
and three becoming T
2
. 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 D
4
symmetry.
Now the A
1
vector of the O group becomes the A
1
vector of the D
4
group. The
triply degenerate T
1
vector splits into a A
2
state and a doubly degenerate E state.
The E states from the O group become a A
1
type state and a B
1
type state.
The T
1
states from the O group become a A
2
type state and a E type state.
The T
2
states from the O group become a B
2
and a E type state.
227
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 specic 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 ` nuclei rather
than just two nuclei.
To do so we begin with the 3` nuclear degrees of freedom.
As for the diatomic case 3 degrees of freedom determine the center of mass motion.
That leaves us with 3` 3 coordinates to specify.
One must now consider two dierent 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 3` 5 vibrational degrees of freedom for linear poly-
atomic molecules and 3` 6 vibrational degrees of freedom for nonlinear mole-
cules.
228
228
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 C
2
. The character table is
C
2

1

C
2
o

o
0

1
1 1 1 1

2
1 1 1 1
1
1
1 1 1 1
1
2
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
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
1
The same is true for the symmetric stretching mode. It too is associated with
1
.
The asymmetric stretch, however, is associated with 1
1
since

C
2
and o
0

transform
the mode into its opposite and o

leaves it unchanged.
230
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
Lightmatter interaction is dominated by the lightdipole coupling so the
other 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
j = j
a
c
a
+j
j
c
j
+j
:
c
:
(34.2)
where j
a
= r. j
j
= . j
:
= ..
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 same
symmetry as either r, . or . for the particular group.
Example: Water
Recall that the point group symmetry of the water molecule is C
2
.
We now need a column of the character table which we have ignored up to this
point.
The character table is
C
2

1

C
2
o

o
0

Functions

1
1 1 1 1 .. r
2
.
2
. .
2

2
1 1 1 1 r
1
1
1 1 1 1 r. r.
1
2
1 1 1 1 . .
The last column describes the symmetry of several important functions for the
point group.
232
Among these functions are r, . and ..
So we can see immediately that the IR active modes of any molecule having this
point group will be
1
. 1
1
. and 1
2
.
The
2
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
1,
and 1
1
symmetry and hence
are all IR active and appear in the IR spectrum
34.2. Raman Spectroscopy
Raman spectroscopy is somewhat dierent 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 eld.
233
The molecular quantity of interest is the polarizability tensor,

c.
We will not get into tensors in this course except to say the polarizability tensor
elements are proportional to the quadratic functions, r
2
.
2
. .
2
. r. r.. .. (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 : as a symmetry
element) then no modes will be both IR and Raman active.
234
35. Molecular Rotations
Recall that the three degrees of freedom that described the position of the nuclei
about the center of mass were (1. o. c). The 1 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

1
.
=
~
2
2j

\
2
.
=
~
2
2j1
2
J
J1

1
2
J
J1
+
~
2
2j

J
2
. (35.1)
We will now be concerned only with the angular part,

~
2
21

J
2
. (35.2)
Now, under the Born-Oppenheimer approximation, 1is a parameter. For constant
1 the rotational energy is given by
1
rot
=
J(J + 1)~
2
2j1
2
=
J(J + 1)/
2
8:
2
1
. (35.3)
This is the so-called rigid rotor energy.
It is common to dene
1
c
=
/
8:
2
1
(35.4)
as the rotational constant. Then
1
rot
= J(J + 1)/1
c
(35.5)
with a degeneracy of
q
J
= 2J + 1 (35.6)
235
235
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 1 value is dependent on the particular vibrational level.
One denes a rotational interaction constant that depends on the vi-
brational level, :.
1
a
= 1
c

: +
1
2

c
c
. (35.7)
where c
c
is an empirical rotationalvibrational interaction constant.
2. Centrifugal stretching:
Rotation tends to stretch the diatomic distance 1.
This is corrected for by the term
J
2
(J + 1)
2
1
c
. (35.8)
where
1
c
=
41
3
c
.
2
c
(35.9)
is the centrifugal stretching constant.
35.2. Rotational Spectroscopy
A rotational transition can occur in the same vibrational level :. 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
It turns out that typical rotational energy gaps are on the order of a few wavenum-
bers or less.
Thermal energy, /1. 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 specialthe principle axes coordinate system.
The principle axes coordinate system is the one in which the .-axis is taken to be
along the principle symmetry axis.
The total moment of inertia, 1 = 1
aa
+1
jj
+1
::
The Hamiltonian in the principle axes system is

H =
~
2
2
"

J
2
a
1
aa
+

J
2
j
1
jj
+

J
2
:
1
::
#
(35.10)
237
There are four classes of polyatomic molecules regarding rotations
1. Linear (e.g., carbon dioxide)
1
::
= 0. 1
aa
= 1
jj


J
2
=

J
2
a
+

J
2
j
The Hamiltonian is

H =
~
2
21
aa

J
2
(35.11)
The rotational energy is
1
rot
= /1J(J + 1). (35.12)
where
1 =
/
8:
2
1
aa
(35.13)
2. Symmetric tops (e.g., benzene)
1
aa
= 1
jj


J
2
=

J
2
a
+

J
2
j
+

J
2
:
The Hamiltonian is

H =
~
2
2
"

J
2
a
+

J
2
j
1
aa
+

J
2
:
1
::
#
(35.14)
The rotational energy is
1
rot
= /1J(J + 1) +/(1)1
2
. (35.15)
where
=
/
8:
2
1
::
. (35.16)
1 =
/
8:
2
1
aa
(35.17)
and 1 is the quantum number describing the projection of the angular
momentum onto the .-axis
238
3. Spherical tops (e.g., methane)
1
aa
= 1
jj
= 1
::


J
2
=

J
2
a
+

J
2
j
+

J
2
:
The Hamiltonian is

H =
~
2
21
aa

J
2
(35.18)
The rotational energy is
1
rot
= /1J(J + 1). (35.19)
where
1 =
/
8:
2
1
aa
(35.20)
4. Asymmetric tops
1
aa
6= 1
jj
6= 1
::
These are more complicated and we will not discuss them in detail
239
36. Electronic Spectroscopy of
Molecules
The electronic spectra of molecules are quite dierent 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 BornOppenheimer 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, 1
c
, was parameterized by
the internuclear distance, 1.
1
c
as a function of 1 describe the eective potential for the nuclei.
It had a qualitative shape similar to the Morse potential.
240
240
In the gure below the ground and rst excited electronic levels (as a function of
1) are shown.
Note: The potential minima are not at the same value of 1 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
+ + + 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 uorescence 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
36.2. FranckCondon 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 FranckCondon
integral.
36.2.1. The FranckCondon principle
When the BornOppenheimer approximation is applied to spectroscopic transi-
tions, one obtains the FranckCondon principle.
The FranckCondon principle states that the nuclei do not move during an elec-
tronic transition.
Physically this means that for a particular transition to be FranckCondon ac-
tive there must be good overlap of the vibrational wavefunctions involved in the
transition.
Mathematically this means that the strength of a transition from[
i
=
c|,i

ib,i

[
)
=
c|,)

ib,)
is given by

Z
all
space
[
+
)
j
c|
[
i

2
=

Z
c|
space
Z
ib
space

+
c|,)

+
ib,)
j
c|

c|,i

ib,i

2
. (36.1)
243
where j
c|
is the electronic transition dipole. We can separate the integrals as

Z
c|
space

+
c|,)
j
c|

c|,i

2
| {z }
if 6=0, allowed

Z
ib
space

+
ib,)

ib,i

2
| {z }
FranckCondon
. (36.2)
244
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 timefrequency 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 ,(t) will, in this work, by denoted
by a tilde,

,(.). and is given by
=[,(t)] =

,(.) =
Z
o
o
,(t)c
i.t
dt. (37.1)
245
245
The Fourier transformation is unique and it has a unique inverse, =
1
, which is
given by
=
1
h

,(.)
i
= ,(t) =
1
2:
Z
o
o

,(.)c
i.t
d.. (37.2)
The above two relations form the convention used throughout this work.
Other authors use dierent 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 transformation
operation, i.e., =[,(t)] =

,(.). Whereas the symbol =
1
will represent the inverse
Fourier transformation, i.e., =
1
h

,(.)
i
= ,(t).
246
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: 3` 5
not linear: 3` 6
The so-called rigid rotor energy is
1
rot
= J(J + 1)/1
c
. (37.3)
where
1
c
=
/
8:
2
1
(37.4)
is the rotational constant.
247
247
The degeneracy of the rigid rotor is
q
J
= 2J + 1 (37.5)
FranckCondon Factor:

Z
ib
space

+
ib,)

ib,i

2
(37.6)
The Fourier transformation is
=[,(t)] =

,(.) =
Z
o
o
,(t)c
i.t
dt. (37.7)
The inverse Fourier transformation is
=
1
h

,(.)
i
= ,(t) =
1
2:
Z
o
o

,(.)c
i.t
d.. (37.8)
248
Part VII
Kinetics and Gases
249
249
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 (:), : is the molecular (or atomic) mass
Each particle of mass : has velocity , hence a momentum of j = : and
a kinetic energy of 11 =
1
2
: =
1
2
:
2
.
A sample of ` molecules is characterized by its number density :
+
=
.
\
.
From the ideal gas law 1\ = :11 =
.
1
11 (1 is Avogadros number):
.
\
=
11
1T
= :
+
Consider the i
th
particle at position r
i
= (r. . .) in coordinate (position) space.
Its velocity is
i
=
oa
.
ot
=

oa
.
ot
.
oj
.
ot
.
o:
.
ot

. This can represented in velocity space by


the vector
i
= (
a
.
.
j
.
.
:
.
).
The velocities of the particles are characterized by a probability distribution func-
tion for velocities 1(
a
.
j
.
:
. t) which is in general a function of time, t.
250
250
The number of particles, `
\r
. having velocities in a macroscopic volume, \

. in
velocity space is
`
\r
= `
Z
\
r
1(. t)d = `
Z Z Z
\
r
1(
a
.
j
.
:
. t)d
a
d
j
d
:
(38.1)
It is more convenient to switch to spherical polar coordinates in velocity space
(. o. c); n.b., is simply a magnitude (not a vector)it is the speed.
The probability distribution function then becomes 1(. o. c. 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 isotropicindependent
of o and c.
1(. o. c. t) = 1(. t). (38.2)
Furthermore, stationary distributionsthose independent of timeare often en-
countered.
1(. o. c. t) = 1(. o. c). (38.3)
251
We shall consider stationary isotropic distributions 1(). So 1() represents a
distribution of speeds.
It can be shown from rst principles that
1() = 4:

:
2:/
b
1
3
2
c
3rr
2
2I
l
T

2
(38.4)
where /
b
= 1.380658 10
23
is Boltzmanns constant. This is the Maxwells
distribution (of speeds).
38.2. Molecular Collisions
The average speed of a particle can calculated from Maxwells distribution:
hi = =
Z
o
0
1()d =
Z
o
0

3
4:(
:
2:/1
)
3
2
c

rr
2
2IT
d (38.5)
=
s
8/1
::

1
1

1I=1
=
1n=A
r
811
:`
It will be convenient to dene number density as :
+
=
.
\
where ` is the number
of particles. For an ideal gas (\ =
a1T
1
), :
+
=
`
:
|{z}
=1
1
1T
=
11
1T
.
A simple model for molecular collisions:
252
Particles are hard spheres of radius o.
A Particle moving at sweeps out a cylinder of radius o and length 4t ==
\ = :o
2
4t.
+ + + See Handout + ++
The number of collisions equals the number of particles with their centers
in \ :
number of collisions = :
+
:o
2
4t (38.6)
The collision frequency = :
+
:o
2

For the above model we need to nd the average collision frequency. Since the
molecules are moving relative to one another we must nd the average relative
velocity,
12
= h|
1

2
|i
It can be shown that

12
=
r
1611
:`
=
_
2 ==
collision
frequency
=
_
2:
+
:o
2
. (38.7)
From the above expression one denes the mean free path ` to be
` =
/
_
2:
+
:o
2
/
a
W
=
1T
TT
=
11
_
211:o
2
(38.8)
Example: Ar at SATP (1 = 298 K, 1 = 1 bar):
= 380.48
m
s
.
collision
frequency
= 5.25 10
9
s
1
.
` = 72.5 nm
253
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 eld 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
cA + /B = 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]
dt
or
d[C]
dt
or
d[D]
dt
(39.2)
BUT this is ambiguous because c moles of A reacts with / moles of B and c does
not, in general, equal /. We must account for the stoichiometry.
254
254
We dene the reaction velocity as
v =
1

i
d[I]
dt
(39.3)
where
i
= c. /. c or d and I = A, B, C, or D.
This denition is useful but must be used with caution since for complicated
reactions all the vs may not be equal. An example of this is
cA +

/B cC + dD
/
0
B
0
c
0
C
0
+ d
0
D
0
(39.4)
A rate law is the mathematical statement of how the reaction velocity depends
on concentration.
v = ,(conc.) (39.5)
For the most part, rate laws are empirical.
Many, but certainly not all, rate laws are of the form
v = /[A
1
]
a
A
1
[A
2
]
a
A
2
[A
a
]
a
A
n
. (39.6)
The reaction is said to be of order r
A
.
in species A
i
and it is of overall order
P
i
r
A
.
.
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
Let
2A + 2B = C + D (39.8)
be the overall reaction. One possible set of elementary steps could be
elementary rxn molecularity
A + A A
0
Bimolecular
A
0
A
00
Unimolecular
A
00
+ 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 = /[A]
Bimolecular A + A Product v = /[A]
2
Bimolecular A + B Product v = /[A][B]
Trimolecular A + A + A Product v = /[A]
3
Trimolecular A + A + B Product v = /[A]
2
[B]
Trimolecular A + B + C Product v = /[A][B][C]
.
Conversely, rate laws for overall reactions can not be determined by stoichiometry.
Connection to thermodynamics
Consider the overall or elementary reaction
cA + /B
I
]

Ir
cC + dD (39.9)
where /
)
is the rate constant for the reaction to proceed in the forward direction
and /
v
is the rate constant for the reaction to proceed in the reverse direction.
Now, at equilibrium v
)
= v
b
which implies
/
)
[A]
o
[B]
b
= /
v
[C]
c
[D]
o
(39.10)
256
bringing /
v
to the LHS and [A][B] to the RHS we get
/
)
/
v
=
[C]
c
[D]
o
[A]
o
[B]
b
= 1
0
c
(39.11)
where 1
0
c
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 laws
Consider the (overall) reaction between molecular hydrogen and molecular iodine,
H
2
+ I
2
= 2HI. (39.12)
The observed rate laws are v
)
= /
)
[H
2
][I
2
] and v
v
= /
v
[HI]
2
. This suggests that the
reaction is elementary. In fact, the reaction is not elementary. Moral: Kinetics
is very much an empirical science.
Next consider the reaction between molecular hydrogen and molecular bromine,
H
2
+ Br
2
= 2HBr. (39.13)
The observed rate law for this reaction is very complicated,
v =
/[H
2
][Br
2
]
12
1 +
I
0
[HBr]
[Br
2
]
.
this does not obey any common form.
The above two example are seemingly very similar but they have very dierent
observed rate laws. Moral: Kinetics is very much an empirical science.
Objectives of chemical kinetics
257
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 thermodynamicsmore di!cult
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
oc(t)
ot
. That is we must measure c(t) as a
function of time and nd 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 catalysts
to stop a reaction at a given point.
Very fast (sec/subsec) reactions cause problems because the reaction goes
faster than one can mix the reactants.
258
39.2.1. Dierential 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 amplied when taking the derivative.
1. Method of initial velocities
for v = /[A]
a
[B]
j
rate laws.
initially v
0
= /c
a
/
j
where c and / are the initial concentrations of A
and B respectively
taking the log of both sides gives lnv
0
= ln[/c
a
/
j
] = ln / +rlnc+ ln /
c and / can be varied independently so both r and can be determined.
problems
1. if the concentration drops very sharply
2. if there is an induction period
2. Method of isolation
for v = /[A]
a
[B]
j
rate laws
start with initial concentrations c and / equal to the stoichometry; this
gives the overall order of r +
ood with, say, A so v - /c
a
[B]
j
39.2.2. Integrated rate laws
The above dierential methods look directly at the rate law which is a dierential
equation. The dierential equation is not solved.
259
We now solve the dierential equations to yield what are called the integrated
rate law.
The dierential equations (rate law) and their solutions (integrated rate law) are
simply listed here for a few rate laws.
type rate law
o)
integrated rate law
o)
1
st
order
1

.
o[1]
ot
= /[1] [1] = [1
0
]c

.
It
2
nd
order
1

.
o[1]
ot
= /[1]
2 1
[1]
=
1
[1
0
]

i
/t
:
th
order
b) 1

.
o[1]
ot
= /[1]
a 1
[1]
n31
=
1
[1
0
]
n31
(: 1)
i
/t
enyzme
1

.
o[1]
ot
=
I[1]
I
r
+[1]
/
n
ln
[1
0
]
[1]
+ ([1
0
] [1]) =
i
/t
o)
[1] is the concentration of one of the reactants in an elementary reaction and

i
is the stoichiometric factor for [1] (n.b.,
i
is a negative number).
b)
The order need not be an integer. For example : = 3,2 is a three-halves
order rate law.
260
40. Temperature and Chemical
Kinetics
40.1. Temperature Eects on Rate Constants
An empirical rate constant was proposed by Arrhenious:
d ln/
d1
=
1
o
11
2
or (40.1)
d ln /
d(1,1)
=
1
o
1
. (40.2)
where 1
o
is the Arrhenious activation energy.
Integration of the above yields
ln / = ln
1
o
11
==/ = c

T
a
TT
(40.3)
( is the constant of integration). This is the Arrhenious equation
Recall the equilibrium constant can also be obtained from kinetics
1
0
c
=
/
)
/
v
' 1
o
. (40.4)
Now, take the log of this:
ln1
o
= ln

/
)
/
v

= ln/
)
ln /
v
. (40.5)
261
261
Substituting the Arrhenious equation for the rate constants gives
ln1
o
= ln

)
c

Ta
]
TT

ln
h

v
c

T
a
r
TT
i
(40.6)
= ln

+
1
o
r
1
o
]
11
40.1.1. Temperature corrections to the Arrhenious parameters
The Arrhenious parameters and 1
o
are constants.
Theoretical approaches to reaction rates predict rate constants of the form
/ = c1
)
c
1
0
1T
. (40.7)
Forcing this to coincide with the Arrhenious implies
1
o
= 1
0
+,11 (40.8)
and
= c1
)
c
)
(40.9)
We can verify this by starting with the Arrhenious equation and substituting the
above expressions,
/ = c

T
a
TT
= c1
)
c
)
c

T
0
+TT
TT
= c1
)
c
)
/ c
)
/ c
3T
0
TT
= c1
)
c
3T
0
TT
_
(40.10)
40.2. Theory of Reaction Rates
Simple collision theory (SCT)
Bimolecular reactions (A,B)
Reaction rate determined by molecular collisions
262
Collision frequency for AB collisions
.
AB
= :o
AB
1
2
s
811
:1j
[A][B] (40.11)
where j =
n
A
n
B
n
A
+n
B
is the reduced mass and o
AB
is the collision diameter.
The maximum reaction velocity is v
max
=
:
AB
1
. but intuitively the actual
reaction velocity will be less because
the ability to react depends on orientation == a steric factor j
a minimum amount of collisional energy is required== c
1
min
1T
The actual reaction velocity is
v =
j.
AB
c

T
min
TT
1
(40.12)
The rate constant for a bimolecular reaction is
/ =
v
[A][B]
(40.13)
so SCT predicts
/ =
j:
AB
c
3
T
min
TT
1
[A][B]
= j:o
AB
1
s
811
:1j
c

T
min
TT
(40.14)
263
Comparison to the (temperature corrected) Arrhenious equation suggests
= j:o
AB
1
s
811
:1j
c
1
2
(40.15)
and
1
o
= 1
min
+
1
2
11 (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 apply
thermodynamics to it.
For the above example, the equilibrium constant is dened as
1

o
=
c

c
1
low
'
conc.
[]
[A][B]
(40.18)
264
Denition: transmission factor, ,
accounts for the fraction of activated complex that becomes product.
From statistical mechanics, it can be shown that , = /
b
1,/ where /
b
is Boltzmanns constant and / is Plancks constant.
The reaction rate constant for reactants going to products for ACT is
/ = ,1

o
=
/
b
1
/
1

o
(40.19)
Thermodynamics tells us that
4G

= 11 ln1

o
(40.20)
which can be written as
1

o
= c

4C

TT
= c

41

TT
c
4S

T
(40.21)
where 4G

= 4H

14o

.
The ACT reaction rate constant now becomes
/ =
/
b
1
/
c

41

TT
c
4S

T
. (40.22)
This is Eyrings 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 nal products.
Reactions may occur in series or in parallel or both, in what is called a reaction
network.
Parallel reactions:
265
Parallel reactions are of the form, for example,
A + B
1
I
1
C (40.23)
A + B
2
I
2
D
The rate constant for the disappearance of [A] is simply the sum of the two
rate constants: / = /
1
+/
2
Series reactions:
Series reactions necessarily include and intermediate product. They are of
the form
A
I
1
B
I
2
C (40.24)
The concentrations of A, B and C are determined by the system of dieren-
tial equations:

d[A]
dt
= /
1
[A]
d[B]
dt
= /
1
[A] /
2
[B]
d[C]
dt
= /
2
[B].
which, when solved yields
[A] = [A
0
]c
I
1
t
[B] =
/
1
[A
0
]
/
2
/
1

c
I
1
t
c
I
2
t

[C] = [A
0
] [A] [B] = [A
0
]

1
/
2
c
I
1
t
/
1
c
I
2
t
/
2
/
1

See in class animation


266
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
MR.
267
2. Propagation: The radical formed in the initiation step reacts with some so
molecule M
0
to form another molecule M
00
and another radical R
0
. This step
repeats an indenite number of times.
R+M
0
M
00
+ R
0
.
3. Termination: The radicals interact with each other or with the walls of the
container to forma stable molecule
R
0
+R
0
M
000
or
R
0
+wall removed
268
41. Gases and the Virial Series
Unlike liquids and solids, a particular particle has much less signicant interactions
with the other particles.
This simplies 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
1\ = :11. (41.1)
The equation of state can also be expressed in term of density j =
n
\
j =
:1
:11
. (41.2)
The van der Waals gas equation of state
1 =
:11
\ :/

:
2
c
\
2
(41.3)
or
1 =
11
\
n
/

c
\
2
n
. (41.4)
where the parameter c accounts for the attractive forces among the particles
and parameter / accounts for the repulsive forces among the particles
269
269
Berthelot
1 =
:11
\ :/

:
2
c
1\
2
=
11
\
n
/

c
1\
2
n
(41.5)
Dieterici
1 =
:11c

an
TT1
\ :/
=
11c

a
TT1
r
\
n
/
(41.6)
Redlich-Kwang
1 =
:11
\ :/

:
2
c
_
1\ (\ :/)
=
11
\
n
/

c
_
1\
n
(\
n
/)
(41.7)
41.2. The Virial Series
Denition: Compressibility Factor: . =
1\
a1T
=
1\r
1T
.
. is unity for an ideal gas because for such a gas 1\ = :11.
For a real gas . must approach unity upon dilution (
a
\
0).
. can be expended in a power series called the virial series.
The virial series in powers of
a
\
is
. = 1 +1(1)

:
\

+C(1)

:
\

2
+1(1)

:
\

3
+ . (41.8)
or
. = 1 +1(1)

1
\
n

+C(1)

1
\
n

2
+1(1)

1
\
n

3
+ . (41.9)
1(1). C(1). etc. are called the virial coe!cients.
Conceptually 1(1) represents pair-wise interaction of the particles, C(1) repre-
sents triplet interactions, etc.
270
41.2.1. Relation to the van der Waals Equation of State
Recall the van der Waals equation
1 =
11
\
n
/

c
\
2
n
(41.10)
multiply both sides by
\r
1T
to get
1\
n
11
=
\
n
1/ 1 /
1/ 1 /
\
n
/

\/
n
11
c
\
2 /
n
=
\
n
\
n
/

c
11\
n
=
1
1
b
\r

c
11\
n
(41.11)
but
1\r
1T
= . so
. =
1
1
b
\
r

c
11\
n
. (41.12)
The rst term is of the form
1
1a
which has the power series expansion
1
1 r
= 1 +r +r
2
+ . (41.13)
Therefore
. =
c
11\
n
+ 1 +
/
\
n
+

/
\
n

2
+ . (41.14)
the rst term is proportional to
1
\r
and so it can be combined with the
1
\r
term
in the series expansion, hence
. = 1 +

/
c
11

1
\
n
+

/
\
n

2
+ . (41.15)
This series can now be compared termby term to the virial series to give expression
for the virial coe!cients:
1(1) =

/
c
11

. C(t) = /
2
. 1(1) = /
3
. etc. (41.16)
271
41.2.2. The Boyle Temperature
The temperature at which 1(1) = 0 is called the Boyle temperature, 1
b
.
The virial series at 1
b
becomes
.(1 = 1
b
) = 1 + 0

1
\
n

+C(1)

1
\
n

2
+1(1)

1
\
n

3
+
= 1 +C(1)

1
\
n

2
+1(1)

1
\
n

3
+ . (41.17)
The lowest order correction are now

1
\r

2
. The gas behaves more like an ideal
gas at 1
b
then for other temperatures.
41.2.3. The Virial Series in Pressure
One can also expand the compressibility factor in pressure
. = 1 +1
0
(1)1 +C
0
(1)1
2
+1
0
(1)1
3
+ . (41.18)
The relation of this expansion to the one in
1
\
r
can be obtained. One nds (see
homework)
1
0
(1) =
1(1)
11
. (41.19)
272
C
0
(1) =
C(1) 1(1)
2
(11)
2
(41.20)
and
1
0
(1) =
1(1) 31(1)C(1) 21(1)
3
(11)
3
(41.21)
41.2.4. Estimation of Virial Coe!cients
The virial coe!cients can be estimated using empirical equations and tabulated
parameters.
Estimates based on Beattie-Bridgeman constants:
1(1) = 1
0


0
11

c
1
3
. (41.22)
C(1) =

0
c
11
1
0
/
1
0
c
1
3
. (41.23)
1(1) =
1
0
/c
1
3
. (41.24)
where
0
. 1
0
. c. /. c are tabulated constants
Estimates based on critical values (we will discuss critical values shortly, for
now treat them as empirical parameters):
1(1) =
911
c
1281
c

1
61
2
c
1
2

. (41.25)
273
42. Behavior of Gases
42.1. 1. \ and 1 behavior
We shall briey consider the 1. \ and 1 behavior of dense uids (e.g., liquids).
Taking volume as a function of 1 and 1. we consider the total derivative
d\ (1. 1) =

J\
J1

1
d1 +

J\
J1

T
d1. (42.1)
We can change this from a extensive property equation to an intensive property
equation by dividing by \ :
d\
\
=
1
\

J\
J1

1
| {z }
c
d1 +
1
\

J\
J1

T
| {z }
i
T
d1.
c is the coe!cient of thermal expansion.
At a given pressure. c describes the change in volume with temperature.
Positive c means the volume of the uid increases with increasing temper-
ature.
i
T
is the isothermal compressibility
At a given temperature, i
T
describes the change in volume with pressure.
Positive i
T
means the volume of the uid decreases with increasing pressure.
i
T
is dierent from ., the compressibility factor.
274
274
42.1.1. c and i
T
for an ideal gas
As an exercise we shall calculate c and i
T
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: \ =
a1T
1
.
i
T
=
1
\

J\
J1

T
=
1
\

a1T
1

J1
!
T
=
1
\

:11
1
2

(42.2)
=
1
(1\ )
| {z }
=a1T
:11
1
=
1
: / 1/ 1 /
: / 1/ 1 /
1
=
1
1
and
c =
1
\

J\
J1

1
=
1
\

a1T
1

J1
!
1
=
1
\ 1
|{z}
=a/1/T
: / 1/ =
1
1
(42.3)
42.1.2. c and i
T
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, 1
0
.
At constant 1
\ (1) = \
0
+

J\
J1

T
| {z }
\
0
i
T
(1 1
0
) +

J\
J1

2
T
(1 1
0
)
2
+ (42.4)
so,
\ (1) - \
0
[1 i
T
(1 1
0
)] . (42.5)
This approximation is quite good even over a rather large pressure range (11
0
=
100 atm or so).
275
Likewise at constant 1
\ (1) = \
0
+

J\
J1

1
| {z }
\
0
c
(1 1
0
) +

J\
J1

2
T
(1 1
0
)
2
+ (42.6)
so,
\ (1) - \
0
[1 +c(1 1
0
)] . (42.7)
As one nal point, we can apply the cyclic rule for partial derivatives to determine
the ratio
c
i
T
:
c
i
T
=

0\
0T

0\
01

T
cyclic
=
rule

J1
J1

\
(42.8)
42.2. Heat Capacity of Gases Revisited
This section is a review from the rst semester with an additional example beyond
the ideal gas.
42.2.1. The Relationship Between C
1
and C
\
To nd how C
1
and C
\
are related we begin with
C
1
=

JH
J1

1
. H = l +1\ (42.9)
so
C
1
=

J (l +1\ )
J1

1
=

Jl
J1

1
+1

J\
J1

1
(42.10)
note

0l
0T

1
is not C
\
we need

0l
0T

\
. Use an identity of partial derivatives

Jl
J1

1
=

Jl
J1

\
+

Jl
J\

J\
J1

1
(42.11)
276
thus
C
1
=

Jl
J1

\
+

Jl
J\

J\
J1

1
+1

J\
J1

1
(42.12)
= C
\
+

J\
J1

Jl
J\

T
+1

.
Recall the expression for internal pressure

0l
0\

T
= 1

01
0T

\
1. Then
C
1
= C
\
+

J\
J1

1

J1
J1

\
1/ +1/

(42.13)
Finally
C
1
= C
\
+1

J\
J1

J1
J1

\
(42.14)
For solids and liquids:

J\
J1

1
= \ c.

J1
J1

\
=
c
i
T
(42.15)
so
C
1
= C
\
+
c
2
1\
i
T
(42.16)
For gases we need the equation of state which often is conveniently explicit in 1
or \ but not both
1. Explicit in 1: Replace

J\
J1

1
with

01
0T

01
0\

T
(42.17)
2. Explicit in \ : Replace

J1
J1

\
with

0\
0T

0\
01

T
(42.18)
277
Examples
1. Ideal gas (equation of state: 1\ = :11): This equation is easily made
explicit in either 1 or \ so we dont need any of the above replacements
C
1
= C
\
+1

J\
J1

J1
J1

\
(42.19)
= C
\
+1
:1
1
:1
\
=
:11
1\
= :1
Thus C
1
= C
\
+:1 or C
1n
= C
\ n
+1
2. One term viral equation (equation of state: \ =
a1T
1
+:1). This is explicit
in \ so use case 2 above
C
1
= C
\
+1

J\
J1

J1
J1

\
= C
\
1

J\
J1

0\
0T

0\
01

T
(42.20)
The partial derivatives are

J\
J1

1
=
:1
1
+:1
0
.

J\
J1

T
=
:11
1
2
. (42.21)
so

0\
0T

0\
01

T
=
a1
1
+:1
0

a1T
1
2
=
: / 1(1+11
0
)
: / 11
. (42.22)
Thus
C
1
= C
\
+1 /

:1
1
+:1
0

1(1+11
0
)
11 /
!
(42.23)
= C
\
+:1

1 +
11
0
1

2
or
C
1n
= C
\ n
+1

1 +
11
0
1

2
(42.24)
278
42.3. Expansion of Gases
Expanding gases do work:
n =
Z
\
2
\
1
1
ex
d\ (42.25)
As we learned last semester the value of n depends on 1
ex
during the expansion.
Recall that if the expansion is reversible, there is always an intermediate equi-
librium throughout the expansion. Namely 1
gas
= 1
ex
. So,
n
rev
=
Z
\
2
\
1
1
gas
d\ (42.26)
For an ideal gas (1 =
a1T
\
) this becomes
n
rev
=
Z
\
2
\
1
:11
\
d\ = :11 ln

\
2
\
1

(42.27)
Also recall that n
rev
is the maximum possible work that can be done in an
expansion. n
rev
= n
max
.
42.3.1. Isothermal and Adiabatic expansions
We shall consider two limits for the expansion of gases
1. Isothermal expansion 1 is constant
2. Adiabatic expansion = 0.
Isothermal expansion
For the case of a ideal gas, l(1. \ ) = l(1) (independent of \ ). So for
isothermal expansion 4l = 0 = +n == = n.
279
Adiabatic expansion
Since = 0, dl = dn = 1
ex
d\ = 1d\ (reversible).
For an ideal gas
dl = 1d\ =
:11
\
d\ (42.28)
42.3.2. Heat capacity C
\
for adiabatic expansions
Considering an ideal gas going adiabatically from (1
1
. \
1
) to (1
2
. \
2
).
Recall
C
\
=

Jl
J1

\
==dl = C
\
d1 (42.29)
So from above
C
\
d1 =
:11
\
d\ ==
C
\
d1
1
=
:1d\
\
(42.30)
Going from (1
1
. \
1
) to (1
2
. \
2
):
Z
T
2
T
1
C
\
1
d1 =
Z
\
2
\
1
:1
\
d\. (42.31)
If C
\
(1) is reasonably constant over the internal 1
1
to 1
2
then this is approxi-
mately

C
\
ln

1
2
1
1

= :1ln

\
2
\
1

(42.32)
where

C
\
=
1
2
(C
\
(1
1
) +C
\
(1
2
)) . Or, in terms of molar heat capacity

C
\ n
ln

1
2
1
1

= 1ln

\
2
\
1

(42.33)
280
42.3.3. When 1 is the more convenient variable
What if 1 is the more convenient variable? Then use H instead of l
Let us still consider an adiabatic expansion
H = l + 1\. dH = dl + 1d\ + \ d1 (because both 1 and \ can, in general,
change)
dH = d +dn +1/ d\/ +\ d1 (42.34)
dH = \ d1.
Now,
C
1
=

JH
J1

1
==dH = C
j
d1 = \ d1 (42.35)
For an ideal gas this becomes
C
j
d1 =
:11
1
d1 (42.36)
Going from (1
1
. 1
1
) to (1
2
. 1
2
):
Z
T
2
T
1
C
1
1
d1 =
Z
1
2
1
1
:1
1
d1. (42.37)
If C
1
(1) is reasonably constant over the internal 1
1
to 1
2
then this is approxi-
mately

C
1
ln

1
2
1
1

= :1ln

1
2
1
1

(42.38)
where

C
1
=
1
2
(C
1
(1
1
) +C
1
(1
2
)) . Or, in terms of molar heat capacity

C
1n
ln

1
2
1
1

= 1ln

1
2
1
1

(42.39)
From the above two cases
ln

1
2
1
1

=
1

C
1n
ln

1
2
1
1

=
1

C
\ n
ln

\
2
\
1

(42.40)
281
So
ln

1
2
1
1

=

C
1n

C
\ n
| {z }
=
ln

\
2
\
1

(42.41)
hence
ln

1
2
1
1

= ln

\
2
\
1

= ln

\
1
\
2

= ln

\
1
\
2

(42.42)
Thus

1
2
1
1

=

\
1
\
2

=1
2
\

2
= 1
1
\

1
. (42.43)
but 1
i
\

i
are arbitrary so this implies 1\

= constant (** NOTE: The axes


should be reversed **)
42.3.4. Joule expansion
Consider a gas expanding adiabatically against a vacuum (1
ex
= 0). In this case
= 0 (adiabatic) and n = 0 (since dn = 1
ex
d\ ).
282
This implies
4l = +n = 0. (42.44)
Internal energy is constant.
We want to nd

0T
0\

l
.
Identity:

J1
J\

l
=

J1
Jl

\
| {z }
1C
1

Jl
J\

T
=
1
C
\

Jl
J\

T
(42.45)
For an ideal gas

0l
0\

T
= 0 (since l(1. \ ) = l(1)). Thus in as much as the
gas can be considered ideal

0T
0\

l
= 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 gure below
283
The work done on the left is
n
1
= 1
1
4\ = 1
1
(0 \
1
) = 1
1
\
1
. (42.46)
The work done on the right is
n
1
= 1
2
4\ = 1
2
(\
2
0) = 1
2
\
2
. (42.47)
Now,
4l = l
2
l
1
= n
1
+n
1
= 1
1
\
1
1
2
\
2
(42.48)
Thus
l
2
+1
2
\
2
= l
1
+1
1
\
1
=H
2
= H
1
(42.49)
For Joule-Thomson expansion the enthalpy is constant.
We want to nd

0T
0\

1
= j. (the Joule-Thomson coe!cient).
Identity:

J1
J1

1
=

J1
JH

1
| {z }
1C
T

JH
J1

T
=
1
C
1

JH
J1

T
= j (42.50)
284
Recall the useful identity

JH
J1

T
= \ 1

J\
J1

1
(42.51)
Thus
j =
\ +1

0\
0T

1
C
1
(42.52)
Example: The one term virial equation: (equation of state 1\ = :11 +:1)
j =
1
C
1

:11
1
:1 +
:11
1
+:11
0

(42.53)
j =
1 +11
0
C
1n
.
Limts:
Low 1: 1
0
is positive and 1 is negative, so j is positivethe gas cools upon
expansion
High 1: 1
0
is nearly zero and 1 is positive, so j is negativethe gas warms
upon expansion
The Joule-Thomson inversion temperature is the temperature where j = 0.
285
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 \ :
dl = d +dn
oq=C
1
oT
== dl = C
\
d1. but also dl = 1do. So
do =
C
\
1
d1 ==4o =
Z
T
2
T
1
C
\
1
d1. (43.1)
At constant 1: (use H = l +1\ instead of l)
dH = dl+1d\ +\ d1 = d1d\ +1d\ +\ d1. So dH = d
oq=C
T
oT
==
dH = C
1
d1. but also dH
oq=ToS
= 1do. So
do =
C
1
1
d1 ==4o =
Z
T
2
T
1
C
1
1
d1. (43.2)
Isothermal expansion of an ideal gas (1\ = :11):
Recall that for isothermal expansion of an ideal gas dl = 0 = 1do 1d\
= do =
1o\
T
.
286
286
Using the equation of state
do =
:1d\
\
==4o =
Z
\
2
\
1
:1
\
d\ = :1ln
\
2
\
1
. (43.3)
Using the equation of state to express \
1
and \
2
in terms of 1
1
and 1
2
.
do = :1ln
\
2
\
1
= :1ln
a / 1/ T /
1
2
a / 1/ T /
1
1
= :1ln
1
2
1
1
. (43.4)
If two variables change in going from the initial to nal 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:
4o

= :

1ln
\

+\
1
\

. 4o
1
= :
1
1ln
\
1
+\

\
1
(43.5)
and
4o
mix
= 4o

+4o
1
(43.6)
287
Recall Avogadros principle: : b \ for an ideal gas. So.
4o
mix
= 1
3
E
E
E
C
:

ln
:

+:
1
:

| {z }
1A
/
+:
1
ln
:
1
+:

:
1
| {z }
1A
T
4
F
F
F
D
= 1(:

lnA

+:
1
lnA
1
)
(43.7)
43.1.1. Entropy of Real Gases
Consider the question: How does o o
ideal
as 1 0 ?
Use Maxwell relation

0S
01

T
=

0\
0T

1
and single term viral equation, \ =
a1T
1
+:1.
So

Jo
J1

T
=

J\
J1

1
=
:1
1
:1
0
(43.8)
Hence
do =

:1
1
:1
0

d1
U
<
<
==o
2
o
1
= :1ln
1
2
1
1
:1
0
(1
2
1
1
) (43.9)
For an ideal gas 1
0
= 0, so
o
ideal
2
o
ideal
1
= :1ln
1
2
1
1
(43.10)
Thus
o
2
o
1
= o
ideal
2
o
ideal
1
:1
0
(1
2
1
1
) (43.11)
Letting 1
1
0 and 1
2
1
0
(Standard pressure 1 bar), this becomes
o
2
o /
ideal
1
= o
ideal
2
o /
ideal
1
:1
0
(1
2
1
1
) (43.12)
Dening o
ideal
2
. 1
2
1
0
as o
0
. So,
o(1
0
) = o
0
:1
0
1
0
(43.13)
288
The entropy at any 1 and 1 can be obtained expresses as
o(1. 1) = o
ideal
(1. 1) :1
0
1 (43.14)
Thus
o(1. 1) = o
0
(1) :1ln
1
1
0
:1
0
1 (43.15)
289
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 Maxwells distribution of speeds is
1() = 4:

:
2:/
b
1
3
2
c
3rr
2
2I
l
T

2
. (43.16)
The average speed of a particle is
hi =
r
811
:`
(43.17)
The mean free path is
` =
11
_
211:o
2
(43.18)
290
290
The reaction velocity is
v =
1

i
d[I]
dt
(43.19)
The relation between the rate constant and the thermodynamic equilibrium
constant is
1
c
=
/
)
/
v
(43.20)
The Arrhenious equation
/ = c

Ta
TT
(43.21)
Important thermodynamic relation:
4G = 4H 14o (43.22)
Eyrings equation is
/ =
/
b
1
/
c

4C

TT
=
/
b
1
/
c

41

TT
c
4S

T
(43.23)
The van der Waals gas equation of state:
1 =
11
\
n
/

c
\
2
n
. (43.24)
Compressibility Factor:
. =
1\
:11
=
1\
n
11
. (43.25)
The virial series is
. = 1 +1(1)

1
\
n

+C(1)

1
\
n

2
+1(1)

1
\
n

3
+ . (43.26)
Relation between heat capacities for an ideal gas:
C
1n
= C
\ n
+1 (43.27)
291
Part VIII
More Thermodyanmics
292
292
44. Critical Phenomena
44.1. Critical Behavior of uids
The point on the top of the coexistence curve is called the critical point. It is
characterized by a critical temperature, 1
c
, and a critical density j
c
.
Law of rectilinear diameters: The average density [j
ave
=
1
2
(j
liq
+ j
vap
)] is
linear in temperature.
293
293
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 A
1
= A
2
in the above gure.
Simple gas laws do not work well near critical points.
294
44.1.2. Gas Constants from Critical Data
Consider the van der Waals equation at the critical point (1
c
. 1
c
. \
nc
)
1
c
=
11
c
\
nc
/

c
\
2
nc
. (44.1)
There is an inection point (
o1
o\r
= 0.
o
2
1
o\
2
r
= 0) at the critical point. So, setting
the rst and second derivatives at the critical point equal to zero we get
d1
d\
n

c
=
11
c
(\
nc
/)
2
+
2c
\
3
nc
= 0 (44.2)
and
d
2
1
d\
2
n

c
=
211
c
(\
nc
/)
3

6c
\
4
nc
= 0 (44.3)
solving these three equations for 1
c
. 1
c
and \
nc
gives
\
nc
= 3/. (44.4)
1
c
=
8c
27/1
. (44.5)
1
c
=
c
27/
2
. (44.6)
These values can be used to nd the compressibility factor, .. at the critical point
.
c
=
1
c
\
nc
11
c
=
3
8
= 0.375. (44.7)
Notice that both c and / whose values depend on the particular gas have dropped
out. That is (for the van der Waals Equation) .
c
= 0.375 for all gases.
The other equations of state give similar results
van der Waals Berthelot Dieterici Redlich-Kwong
.
c
3,8 = 0.375 3,8 = 0.375 2/c
2
' 0.27 0.33
295
44.2. The Law of Corresponding States
We have found that .
c
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 dene unitless reduced variables 1
v
= 1,1
c
. 1
v
= 1,1
c
. and \
v
= \,\
c
.
Then .
v
=
1
r
\
r
1T
r
.
.
v
is a universal functionit 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 c and , at a constant
1 and \.
Suppose some amount of material, d:. goes from c ,
(d
c
)
T
= 1d\
c
j
c
d:
(d
o
)
T
= 1d\
o
+j
o
d:
(d)
T,\
= 1
= 0 since \ is constant
z }| {
(d\
c
+d\
o
) +

j
o
j
c

d:
For a spontaneous process deceases (d < 0)
At equilibrium d = 0. This implies j
o
= j
c
is the condition for equilibrium.
When c, , denote liquid (or solid) and vapor phases, then for a given 1, the
pressure of the system when j
o
= j
c
is the called the vapor pressure of the
material at temperature 1.
296
For phase changes at constant 1 and 1 then (dG)
T,1
=

j
o
j
c

d:. So again
j
o
= j
c
is the condition for equilibrium.
44.3.1. The chemical potential and 1 and 1
How does j vary with 1 and 1?
Generally for homogeneous substances,
dG = od1 +\ d1 +jd: (44.8)
Now,
o =

JG
J1

1,a
(44.9)
So,

Jo
J:

1,T
=
J
J:
JG
J1
=
J
J1
JG
J:
=

Jj
J1

1
. (44.10)
But o = :o
n
(1. 1) so,

Jj
J1

1
= o
n
. (44.11)
Similarly,

Jj
J1

T
= \
n
. (44.12)
Now the total dierential of j is
dj(1. 1) =
S
r
z }| {

Jj
J1

1
d1 +
\
r
z }| {

Jj
J1

T
d1 (44.13)
dj(1. 1) = o
n
d1 +\
n
d1
297
44.3.2. The Clapeyron Equation
At equilibrium j
o
= j
c
so,
o
nc
d1 +\
nc
d1 = o
no
d1 +\
no
d1 (44.14)
Now
d1
d1
=
o
nc
o
no
\
nc
\
no
=
4

o
n
4

\
n
4S=
41
T
=
4

H
n
14

\
n
(44.15)
This is the Clapeyron Equation
d1
d1
=
4

H
n
14

\
n
(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
4

\ = \
n,vap
\
n,liq
' \
n,vap
(44.17)
Assuming the vapor phase obeys the ideal gas equation of state,
4

\ =
11
1
(44.18)
Substituting this into the Clapeyron equation gives
d1
d1
=
4

H
n
1
1T
1
=
4

H
n
1
11
2
(44.19)
Collecting the 1s on one side of the equation and the 1s on the other we get
d1
1
=
4

H
n
1
d1
1
2
(44.20)
Now we identify
o1
1
= d(ln1) and
oT
T
2
= d(1,1) so this becomes
d(ln 1) =
4

H
n
1
d(1,1) (44.21)
298
Rearranging again leads to
d(ln 1)
d(1,1)
=
4

H
n
1
(44.22)
This is the Clausius-Clapeyron equation.
44.4. Equilibria of condensed phases
Examples
Solidliquid
icewater, most other common liquids
Solidsolid
rhombic sulfurmonoclinic sulfur
grey tinwhite tin
graphitediamond
For example a diamond at STP is metastable with respect to graphite.
A diamond is not forever!
At equilibrium j
c
= j
o
this implies (for incompressible liquids and solids)
j

c
+\
nc
(1 1

) = j

o
+\
no
(1 1

) (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
j

c
j

o
= (\
no
\
nc
) (1 1

) (44.24)
299
Thus for any given 1 only one 1 allows for equilibrium.
Recall the Clapeyron equation
d1
d1
=
4
)
H
n
14
)
\
n
=
H
no
H
nc
1 (\
no
\
nc
)
(44.25)
We make the good approximation that 4
)
H
n
is independent of 1 and solve the
Clapeyron equation
Z

d1
1
=
4
)
\
n
d1
4
)
H
n
=ln
1
)
1

)
=
4
)
\
n
(1 1

)
4
)
H
n
(44.26)
where 1

)
is the freezing temperature at standard pressure (1 bar).
44.5. Triple Point and Phase Diagrams
Denitions
Phase Diagram: A graph of 1 vs. 1 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:
j
solid
= j
liq
= j
vap
(44.27)
300
45. Transport Properties of Fluids
Transport properties of matter deal with the ow (or ux) 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 uids:
1. Diusion: The ux of material down a concentration gradient
2. Viscosity: The ux of momentum down a velocity gradient
3. Thermal Conductivity: The ux of energy down a temperature gradient
+ + See Transport Phenomena handout + +
45.1. Diusion
At equilibrium concentration on a bulk solution will be uniform.
So if there exists a concentration gradient there will be a net ux, J. of material
from high concentration to low concentration so as to establish an equilibrium.
J =
1

d:
dt
(45.1)
301
301
The ux of material through a plane depends on the concentration dierence
J = 1
dC
dr
==
1

d:
dt
= 1
dC
dr
where 1 is the diusion constant
1

d:
dt
= 1
dC
dr
(45.2)
This is Ficks rst law of diusion (in one dimension).
The change in concentration in a lamina between r and dr with time is given by
the ux in minus the ux out of the lamina:
JC
Jt
=
J(r) J(r +dr)
dr
=
JJ
Jr
(45.3)
Using Ficks rst law for J
JC
Jt
=
J
Jr
1
JC
Jr
. (45.4)
If 1 is truly constant we get Ficks second law of diusion:
JC
Jt
= 1
J
2
C
Jr
2
. (45.5)
302
The solution of this partial dierential 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(r. t) =
C
0
2
_
:1t
c

i
2
4TI
(45.6)
2. Step function solution
C(r. t) = C
0
"
1
2

1
_
:
Z i
I
4TI
0
c
j
2
d
#
(45.7)
=
1
2
C
0

1 erf

r
_
41t

=
1
2
C
0
erfc

r
_
41t

where erf and erfc are tabulated functions respectively called the error func-
tion and complementary error function.
45.2. Viscosity
Viscosity, j. is the resistance to dierential uid ow, i.e., The tendency of a
liquid to ow at the same velocity throughout.
303
The frictional (viscous) force is 1 = j
o
oa
. (The units of j are
mass
lenghttime
. 1 poise
=
g
cms
.)
Poiseuilles Formula
Applies to Laminar (nonturbulent) ow
For a liquid owing trough a tube (radius :. length |), the volume of ow
4\ in time 4t is
4\
4t
=
::
4
41
8j|
(45.8)
where 41 is the driving pressure, i.e., the dierence in pressure on either
side of the tube.
For a gas
4\
4t
=
::
2
16j|

1
2
i
1
2
)
1
0

(45.9)
where 1
i
is the inlet pressure, 1
)
is the outlet pressure and 1
0
is the pressure
at which the volume is read.
Stokes law: spheres falling through uids
304
The frictional force (exerted upwards) is proportional to velocity: 1
f
= ,.
Stokes showed , = 6:j:
Gravitational force (exerted downwards): 1
g
=
4v
3
3
(j j
0
)q. where q is the
gravitational acceleration (9.8 m/s
2
).
Terminal velocity is reached when 1
f
+1
g
= 0 giving
,
term
+
4::
3
3
(j j
0
)q = 0 (45.10)

term
=
4::
3
(j j
0
)q
3,
using , = 6:j:

term
=
4: / :
3 /
(j j
0
)q
3

6: / j: /
=
2:
2
(j j
0
)q
9j
(45.11)
Related to diusion constant:
1 =
/1
,
)=6jv
=
/1
6:j:
(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, i, of a material is a measure of the tendency of energy
in the form of heat to ow through the material.
Consider a slab of solid material of area between two large parallel plates a
distance 1 apart. The plates are held at constant but dierent temperatures 1
1
and 1
2
(1
1
1
2
) for a su!ciently long time that a steady state exists.
305
Under such conditions, a linear steady state temperature distribution across the
material is established. And a constant rate of heat ow
oq
ot
is needed to maintain
the temperature dierence 41 = (1
1
1
2
)
1

d
dt
= i
41
1
. (45.13)
If we take the limit where 1 becomes innitesimally small (1 dr) we obtain a
dierential form of this equation:
1

d
dt
= Q
)
= i
d1
dr
. (45.14)
where Q
)
is the heat ux. This is called Fouriers law of heat conduction
(one-dimensional version).
Thermal conductivities are positive quantities so Fouriers law says that heat ow
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
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
di!cult 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 in
hand.
The Wiedemann, Frantz and Lorenz equation relates the thermal conductivity to
electrical conductivity, i
c|
for pure metals:
i
i
c|
1
= L = const. (45.15)
where L is the Lorenz number (typically 22 to 29 10
9
V
2
/K
2
).
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
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 : =
a
1
a
2
mole fraction A
2
=
a
2
a
1
+a
2
. A
1
= 1 A
2
molality : =
1000A
2
A
1
A
1
. where `
1
is the molecular weight of species 1
Molarity c
2
=
a
2
1 solution
46.2. Partial Molar Quantities
Thermodynamic properties, in general change upon mixing
4
mix
= properties of soln
X
properties of pure. (46.1)
For example,
4
mix
\ = \
soln
\
solute
\
solvent
(46.2)
Consider a thermodynamic quantity, say, volume.
308
308
In general, it is a function of 1. 1. :
1
and :
2
: \ (1. 1. :
1
. :
2
). So, the total
derivative is
d\ =

J\
J1

1,a
1
,a
2
d1 +

J\
J1

T,a
1
,a
2
d1 +

J\
J:
1

T,1,a
2
d:
1
+

J\
J:
2

T,1,a
1
d:
2
.
(46.3)

0\
0a
.

T,1,a

=

\
i
, the partial molar volume.
Similarly
dG =

JG
J1

1,a
1
,a
2
d1 +

JG
J1

T,a
1
,a
2
d1 +

JG
J:
1

T,1,a
2
d:
1
+

JG
J:
2

T,1,a
1
d:
2
.
(46.4)

0G
0a
.

T,1,a

= j
i
.
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 \

i
H

i
o

i
G

i
Pure liquid i per mole \

ni
H

ni
o

ni
j

i
Whole solution \ H o G
Solution/(total moles) \
n
H
n
o
n
G
n
Partial molar of i in solution

\
i

H
i

o
i
j
i
Apparent molar (of solute)

\

H
Reference state \

i
H

i
o

i
j

i
309
46.2.2. Partial Molar Volumes
Consider the partial molar volume
For constant 1 and 1
d\ =

\
1
d:
1
+

\
2
d:
2
(46.5)
Now,

\
i
depends on concentration, so change each amount of substance propor-
tional to the amount substance present,
d:
1
= :
1
d`. d:
2
= :
2
d`. (46.6)
So,
d\ =

\
1
:
1
+

\
2
:
2

d`
U
oA
==\ =

\
1
:
1
+

\
2
:
2
(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 specically
4
mix
\ = \ (\

n1
:
1
+\

n2
:
2
) (46.8)
=

\
1
:
1
+

\
2
:
2

(\

n1
:
1
+\

n2
:
2
)
=

\
1
\

n1

:
1
+

\
2
\

n2

:
2
4
mix
\ can be positive, negative or zero.
For example,
1. one unit of baseballs are mixed with one unit of basketballs. 4
mix
\ < 0.
2. one unit of baseballs are mixed with one unit of books. 4
mix
\ 0.
310
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 Raoults law
2. innite dilution limit
1. all neighboring molecules are dierent than the given molecule
2. the ideal state for Henrys law
Raoults law limit Henrys 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
Also recall that the mathematical denition of activity c
i
of some species i is
implicitly stated as
lim

c
i
q()
= 1 (46.9)
where q() is any reference function (e.g., pressure, mole fraction, concentration
etc.), and

is the value of at the reference state.


This implicit denition is awkward so for convenience one denes the activity
coe!cient as the argument of the above limit,

i
=
c
i
q()
(46.10)
which we can rearrange as
c
i
=
i
q(). (46.11)
The denition of activity implies that
i
= 1 at q(

) (the reference state)


That is
i
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 denition).
j
i
j

i
= 11 lnc
i
. (46.12)
46.3.2. Raoults Law
In discussing both Raoults law and Henrys law, we are describing the behavior
of a liquid solution by measuring the vapor (partial) pressures of the components
312
For simplicity we consider here only a two component solution.
dG = j
1
d:
1
+j
2
d:
2
. (46.13)
Take dierential change along a line of constant concentration, so
dG = (j
1
:
1
+j
2
:
2
) d` (46.14)
then
G = j
1
:
1
+j
2
:
2
. (46.15)
Recall that
4
mix
G = G(soln) G(pure components) (46.16)
Hence,
4
mix
G = j
1
:
1
+j
2
:
2
j

1
:
1
j

2
:
2
(46.17)
= (j
1
j

1
) :
1
+ (j
2
j

2
) :
2
.
Now,
j
1
j

1
= 11 ln
c
i
c

i
low 1
' 11 ln
1
i
1

i
. (46.18)
where 1
i
is the vapor pressure of the i
th
component above the solution.
313
Thus
4
mix
G = 11

:
1
ln
c
1
c

1
+:
2
ln
c
2
c

(46.19)
or at low 1
4
mix
G = 11

:
1
ln
1
1
1

1
+:
2
ln
1
2
1

(46.20)
46.3.3. Ideal Solutions (RL)
Raoults Law:
1
i
= A
i
1

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 Raoults law
is
4
id(RL)
mix
G = 11

:
1
ln
A
1
1

1
/
1

1
/
+:
2
ln
A
2
1

2
/
1

2
/
!
(46.22)
4
id(RL)
mix
G = 11 (:
1
ln A
1
+:
2
lnA
2
) (46.23)
Again, this is for an ideal solution in the Raoults Law sense.
From
o =

JG
J1

1
and H =

J (G,1)
J (1,1)

1
. (46.24)
the entropy of mixing for an ideal Raoult solution is
4
id(RL)
mix
o = 1(:
1
ln A
1
+:
2
ln A
2
) (46.25)
and the enthalpy of mixing is
4
id(RL)
mix
H = 0 (46.26)
314
(since G,1 is independent of 1,1).
The Reference State (RL)
Let us apply the denition of activity for the Raoults law reference state.
The reference function is q() = = A
i
. and the reference state is A
i
= 1
So,
lim
A
.
1
c
(RL)
i
A
i
= 1 (46.27)
implies
c
(RL)
i
=
(RL)
i
A
i
. (46.28)
and
(RL)
i
1 as A
i
1
Deviations from Raoults Law
Raoults 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 specic interactions between particles, real solutions
generally deviate from Raoults law.
The physical interpretation of deviation from Raoults law is
positive deviation: the molecules prefer to be around themselves rather than
other 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
It is very important to note that this deviation from Raoults 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 Raoults lawNegative deviation from Raoults law
46.3.4. Henrys Law
Henrys Law:
1
i
= /
A
.
A
i
. (46.29)
where /
A
.
is the Henrys law constant,
/
A
.
= lim
A
.
0

1
i
A
i

(46.30)
Henrys law applies to the solute not to the solvent and becomes more correct for
real solution as the concentration of solute goes to zero (A
i
0), i.e., at innite
dilution.
316
The Reference State (HL)
Referring to the denition of activity again we see that the reference function is
q() = = A
i
. and the reference state is now A
i
= 0
So,
lim
A
.
0
c
(HL)
i
A
i
= 1 (46.31)
implies
c
(HL)
i
=
(HL)
i
A
i
. (46.32)
and
(HL)
i
1 as A
i
0
If instead of mole fraction, molality or molarity is used then
c
(HL)
i
=
(HL)
n
.
:
i
(46.33)
and
c
(HL)
i
=
(HL)
A
.
`
i
(46.34)
respectively.
Comparison of Raoults Law and Henrys Law
Both Raoults law and Henrys law become better approximations for real solu-
tions as the solution becomes pure. But, they apply to opposite species in the
solution. Raoults law applies to the dominant species, A
1
1, whereas Henrys
law applies to the subdominant species A
2
0. So, in summary
Raoults law:
1
1 as A
1
1
Henrys law:
2
1 as A
2
0
317
46.4. Colligative Properties
Colligative properties: Properties of dilute solutions that are independent of
the 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 1
)
(freezing point), j
1
(solid)
| {z }
j
s
1
= j
1
(soln).
318
Using the Raoults law reference state (since we are interested in the behavior of
the dominant species), j
1
(soln) = j

1
+11 ln c
1
:
j
c
1
= j

1
+11 lnc
1
(46.35)
Rearranging this and taking the derivative with respect to 1 yields
J
J1

lnc
1
=
1
11
(j
c
1
j

1
) ==
J ln c
1
J1
=
1
11
2

Jj
c
1
J1

Jj

1
J1

(46.36)
Now, using
0j
0T
= H and integrating we get
Z

d lnc
1
=

1
11
2
(H
c
1
H

1
)

d1 =
4
)
H
11
2
d1 (46.37)
lnc
1
=
Z
T
]
T

]
4
)
H
11
2
d1
For small changes in the freezing point we may approximate 1 by 1

)
in the
integrand. So,
lnc
1
'
Z
T
]
T

]
4
)
H
11
2
)
d1 =
4
)
H
11
2
)
X. (46.38)
where X = 1

)
1
)
. The freezing point depression is
X =
11
2
)
ln c
1
4
)
H
46.4.2. Osmotic Pressure
We consider the osmotic pressure at a constant temperature, 1. (so, dG = \ d1).
319
In the above gure j
1
(left) = j
1
(right). hence
j

1
= j

1
+11 ln c
1
+

\
1
. (46.39)
where

\
1
is the partial molar volume of the solvent in solution (di!cult to measure)
and is the hydrostatic (osmotic) pressure.
From the above equation
ln c
1
=

\
1

11
(46.40)
Now we make the approximations

\
1
= \

n1
, c
1
= A
1
= 1 A
2
:
ln(1 A
2
) =
\

n1

11
(46.41)
For dilute solutions A
2
is small so ln(1 A
2
) may be expanded as
ln(1 A
2
) = A
2
+
A
2
2
2

A
3
2
3
' A
2
. (46.42)
but A
2
=
a
2
a
1
+a
2
'
a
2
a
1
for dilute solutions. Thus
:
2
:
1
'
\

n1

11
==:
2
'
\

1
z }| {
:
1
\

n1

11
. (46.43)
320
or,
=
:
2
\

1
|{z}
'c
11 = c11. (46.44)
where c is the concentration of the solute.
Note the similarity of this equation with the ideal gas equation: 1 = c11. Thus
the solute in a very dilute solution behaves as if it were an ideal gas.
321
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 timehow fast we are producing entropy.
47.1. Fundamentals
We know the dierence between reversible and irreversible processes from before.
However, we will state their respective denitions here in a manner best suited
for this chapter.
322
322
Reversible process: dynamical equations are invariant under time inversion
(t t).
e.g., the one dimensional wave equation,
1
c
J
2
n
Jt
2
=
J
2
n
Jr
2
tt
==
1
c
J
2
n
J(t)
2
=
J
2
n
Jr
2
==
1
c
J
2
n
Jt
2
=
J
2
n
Jr
2
. (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
i
J1
Jt
=
J
2
1
Jr
2
tt
==
1
i
J1
J(t)
=
J
2
1
Jr
2
==
1
i
J1
Jt
=
J
2
1
Jr
2
. (47.2)
is not invariant under time reversal.
We will be concerned with the change in entropy, do. which can be split into two
components do = d
c
o +d
i
o.
Denitions
d
c
o is the change in entropy due to interactions with the exterior environ-
ment.
d
i
o is the change in entropy due to internal changes of the system
The quantity d
i
o is called the entropy production.
323
Splitting up do into these two parts permits an easy discussion of both open and
isolated systemsthe dierence between the two appearing only in d
c
o.
General criteria for irreversibility:
d
i
o = 0 (reversible change)
d
i
o 0 (irreversible change)
For isolated systems have d
i
o = do and the principle of Clausius, d
i
o = do _ 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 su!cient
production 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
I
II
Considering the above gure of an isolated system, we write the principle of Clau-
sius as
do = do
1
+do
11
_ 0. (47.3)
The local formulation statement implies
d
i
o
1
_ 0 and d
i
o
11
_ 0 (47.4)
and the possibility of, for example, d
i
o
1
< 0 and d
i
o
11
0 such that d
i

o
1
+o
11


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 ow, chemical reactions, etc. As exam-
ple we now consider the last two of these: heat ow and chemical reactions.
325
47.3.1. Entropy Production due to Heat Flow
Recall from the lecture on transport phenomena that the heat ux Q
)
is given by
Q
)
= i
41
1
(47.5)
We are now interested in exposing the time dependence, so, using Q
)
=
q
4t
we get

4t
=
i41
1
(47.6)
in dierential form this is
d
dt
= i
d1
dr
. (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 1
1
and
the other at 1
2
(take 1
1


1 1
2
) where

1 is the temperature at the interface.
326
Considering the whole system
do =
d
1
1
1
+
d
11
1
2
=
ocS
z }| {
d
c

1
1
1
+
d
c

11
1
2
+
o
.
S
z }| {
d
i

1
1
1
+
d
i

11
1
2
. (47.8)
The quantity d
c

)
is the amount of heat supplied by the environment to hold block
, at its xed temperature.
Furthermore the heat going out of I through the connecting wall is equal to the
heat coming into II through the connecting wall:
d
i

1
= d
i

11
. (47.9)
Using this we see that the entropy production is
d
i
o = d
i

1
1
1

1
1
2

. (47.10)
which we see is positive because d
i

1
< 0 when 1
1
1
2
.
We have still not made a connection to kinetics.
To do so we must consider the entropy production per unit time
o
.
S
ot
.
327
For this example
d
i
o
dt
=
d
i

1
dt

1
1
1

1
1
2

From chapter 24 we know


d
i

1
dt
=
i41
1
. (47.11)
So,
d
i
o
dt
=
i41
1

1
1
1

1
1
2

(47.12)
To determine

1 we use the fact that the heat ow out of I is equal to the heat
ow into II:
d
i

1
dt
=
d
i

11
dt
. (47.13)
Using the above expression for heat ow gives us

1 since,

i / /
1/

1
1

=
i / /
1/

1 1
2

=

1 =
1
1
+1
2
2
; (47.14)
a result we might have guessed.
47.3.2. Entropy Production due to Chemical Reactions
Denitions:
1. Chemical a!nity: a = (4
rxn
G)
T,1
=
P
i

i
j
i
and a = (4
rxn
)
T,\
=

P
i

i
j
i
2. Extent of reaction: is dened by d =
oa
.

.
, where :
i
is the number of moles
of the i
th
component and
i
the stoichiometric factor of the i
th
component.
328
e.g., for the reaction N
2
+ 3H
2
2NH
3
d =
d:
N
2
(1)
=
d:
H
2
(3)
=
d:
NH
3
(2)
(47.15)
and
a = 2j
NH
3
j
N
2
3j
H
2
(47.16)
The connection to kinetics: reaction rate v =
o
ot
The connection to thermodynamics:
(d)
T,\
=
X
i
j
i
d:
i
=
X
i

i
j
i
| {z }
a

d:
i
| {z }
o
= ad (47.17)
but
(d)
T,\
=
oq
z }| {
(dl)
T,\
1do =do =
d
1

ao
z }| {
(d)
T,\
1
(47.18)
so
do =
o
c
S
z}|{
d
1
+
o
.
S
z}|{
ad
1
(47.19)
The entropy production per unit time for a chemical reaction is a function of both
the chemical a!nity and of the reaction rate
d
i
o
dt
=
a
1
d
dt
=
a
1
v _ 0 (47.20)
We see that for a spontaneous process the entropy production per unit time is
positive. This is because a = (4
rxn
)
T,\
is positive as is v.
329
Simultaneous Reactions
For ` simultaneous chemical reactions, the entropy production per unit time
generalizes to
d
i
o
dt
=
1
1
.
X
)=1
a
)
v
)
_ 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 a
1
v
1
< 0. a
2
v
2
0
such that a
1
v
1
+a
2
v
2
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 conned to coupling between the same
types of processes.
That is, diusion is the ux of matter down a concentration gradient. The so-
called Soret eect is ux of matter down a temperature gradient. Conversely, the
so-called Dufour eect is heat ux down a concentration gradient
330
The following table lists a number of thermodynamically coupled phenomena
Flux : material Q (charge)
Gradient
1 Thermoconductivity
Thermomechanical
eect
Soret eect Seebeck eect
1
Mechanocaloric
eect
Hydrodynamic
ow
Reverse
osmosis
Potential of ow
C Dufour eect Osmosis Diusion Nernst Potential
Peltier eect 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 eld.
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 undoneit 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
A strong signal is seen because all the spinning nuclei cooperate.
Each nucleus is in a slightly dierent environment so each spin frequency is
slightly dierent.
The dierent environment (spin frequencies) cause the ensemble spinning
nuclei to dephase
Dephasing causes a decrease in the observed signal because now not all nuclei
are cooperating.
Now a radio pulse with the opposite phase is applied to make the nuclei spin
in 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
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
d1
d1
=
4

H
n
14

\
n
. (47.22)
The Clausius-Clapeyron equation is
d(ln1)
d(1,1)
=
4

H
n
1
(47.23)
Ficks rst law of diusion is
1

d:
dt
= 1
dC
dr
(47.24)
333
333
Ficks second law of diusion:
JC
Jt
= 1
J
2
C
Jr
2
. (47.25)
Relation between the viscosity and the diusion constant:
1 =
/1
,
)=6jv
=
/1
6:j:
. (47.26)
Fouriers law of heat conduction is
1

d
dt
= Q
)
= i
d1
dr
. (47.27)
Mixing
4
mix
= properties of soln
X
properties of pure. (47.28)
Chemical potential
j = j

+11 ln c (47.29)
Raoults Law:
1
i
= A
i
1

i
(47.30)
Raoults law reference
c
(RL)
i
=
(RL)
i
A
i
.
(RL)
i
1 as A
i
1 (47.31)
Henrys Law:
1
i
= /
A
.
A
i
. (47.32)
where /
A
.
is the Henrys law constant,
/
A
.
= lim
A
.
0

1
i
A
i

. (47.33)
Henrys law reference
c
(HL)
i
=
(HL)
i
A
i
.
(HL)
i
1 as A
i
0. (47.34)
334
Index
aLsoiplion spoclioscopy 241
aclivily 146, 811
malhomalical oonilion of 146
aclivily coo!cionl 146, 812
aoiaLalic oxpansion 280
ano hoal capacily 280
aoiaLalic wall 120
angulai momonlum
aooilion of 202
classical 102
oigonfunclions foi 100, 210
jj coupling 202
LS coupling 202
quanlum numLois 100, 210
spin 201
angulai momonlum quanlum num-
Loi 2
anliLonoing oiLilal 71
Aiihonious aclivalion onoigy 261
Aiihonious oqualion 261, 201
lompoialuio coiiocloo 262
alomic oiLilals 40
chomisls picluio 0
physicisls picluio 0
aufLau piinciplo 8
avoiago valuo lhooiom 20
Loilholol gas 18, 270
Linominal coo!cionl 00
Lluo sky 81
Lohi moool 18
Lohi iaoius 10
Lollzmann oisliiLulion 10, 06, 181
Lollzmann`s oqualion 00, 07, 124,
181
Lono oiooi 77
Lonoing oiLilal 71
Loin moool 170
coiioclions lo 17
onlhalpy of solvalion 174
onliopy of solvalion 174
fioo onoigy of solvalion 178, 178
paililion coo!cionl 174
LoinOpponhoimoi appioximalion
62, 00, 28, 240
ano lho FianckConoon piinci-
plo 248
Losons 6
Loylo lompoialuio 272
chain iulo
foi pailial ooiivalivos 107
chaiacloi laLlo
foi lho C2v gioup 22
chomical a!nily 828
chomical polonlial 144
foi a sall 161
iolalion lo aclivily 148
iolalion lo GiLLs fioo onoigy
14
iolalion lo Holmholz fioo on-
oigy 14
88
335
Clapoyion oqualion 208, 800, 888
Clausius-Clapoyion oqualion 200,
888
coo!cionl of lhoimal oxpansion 274
cooxislonco cuivo 208
colligalivo piopoilios 818
commulaloi 80, 180
complolonoss 101
complimonlaiy vaiiaLlos 80
compiossiLilily facloi
al lho ciilical poinl 20
compiossiLilly facloi 270, 201
conguialion 00
conuonl hypoigoomoliic funclions
6
coiiosponoonco piinciplo 41
ciilical poinl 800
cyclic iulo 14, 108
cylinoiical symmoliy 60
DoLyoHuckol limiling law164, 178
DoLyoHuckol lhooiy 168
DoLyoHuckolGuggonhoimoqualion
164
DoLyo`s law 120, 188
oogonoiacy 186
of lho onsomLlo 08
oialhoimic wall 120
oialomic moloculos
oloclion-oloclion polonlial on-
oigy opoialoi foi 61
oloclionic kinoliic onoigy opoi-
aloi foi 61
oloclionic wavofunclion foi 62
Hamillonian foi 61
nucloai kinolic onoigy opoialoi
foi 61
nucloai-oloclion polonlial onoigy
opoialoi foi 61
nucloai-nucloai polonlial onoigy
opoialoi foi 61
Schiooingoi oqualion foi 62
Dioloiici gas 270
oiusion 801
oiusion conslanl 802
oigonfunclion
oigonvaluo
oigonvaluo oqualion 100
olocliic oipolo appioximalion 70, 281
olocliolylos
sliong 161
olocliophoiolic oocl 167
olomonlaiy ioaclions 2
ano sloichiomoliy 26
moloculaiily 26
omission spoclioscopy 241
onomLlo 80
onsomLlo avoiago 108, 182
onlhalpy 186
onliopy 10
chango foi changos in lompoi-
aluio 286
chango foi isolhoimal oxpansion
286
chango foi mixing 287
of ioal gasos 288
onliopy pioouclion 822, 828
ouo lo chomical ioaclions 828
ouo lo hoal ow 826
oqualion of slalo 116
foi a Loilholol gas 118
foi a Dioloiici gas 118
foi a IoolichKwang gas 118
foi a van ooi Waals gas 117
foi an iooal gas 116
foi gasos 260
oquiliLiium conslanl 18
336
oquliLiium conslanl 18
Luloi`s ioonlily 4
oxpansion
of gasos 111
iovoisiLlo 114
oxlonl of ioaclion 828
Lyiing`s oqualion 26, 201
foimions 6
Fick`s isl law 802, 888
Fick`s socono law 802, 884
isl law of lhoimooynamics 121,
188
ipping coins 00
uclualion 02
uoiosconco 242
slokos shifl 242
Fouiioi`s lawof hoal conouclion 806,
884
FianckConoon inlogial 248
FianckConoon piinciplo 248
fioo onoigy
GiLLs 188
Holmhollz 187
fugacily 147
funoamonlal liansislions 66
gonoial oquliLiium 11
gonoializoo oisplacomonl 110
gonoializoo foico 110
goiaoo 60
GiLL`s fioo onoigy 106
GiLLs-Duhom oqualion 168
gooo lhooiy 16
gioup
malhomalical oonilion of 222
mulliplicalion laLlo 228
gioup lhooiy 221
Hamillonian opoialoi 27
Hamilonian
classical 27
haimonic oscillaloi 88
onoigy lovols foi 40, 44, 86
polonlial onoigy 80
Schiooingoi oqualion foi 80
hoal 100
sign convonlion 110
hoal capacily 11, 188
HoisonLoig uncoilainly piinciplo 80
ano lho haimonic oscillaloi 41
holium
oloclion-oloclion iopulsion loim

Hamillonian
Holmhollz fioo onoigy 106
Honiy`s law 811, 816, 884
Honiy`s law conslanl 816, 884
Hoimilo polynominals 40
hol Lanos 66
Huno`s iulo 20
hyoiogon alom
ionizlion onoigy of 10
hyoiogon moloculo 74
hyoiogonic sysloms 46
onoigy lovols foi 40, 86
Hamillonian 47
noimalizalion conslanl 40, 8
polonlial onoigy foi 47
Schiooingoi oqualion foi 47
wavofunclion (no spin) 40
wavofunclion (wilh spin) 2
iooal solulion
Iaoull`s law 814
immisciLlo solulions 18
infiaioo spoclioscopy 66
inloinal onoigy 108, 121
inliamoloculai viLialional iolaxalion
(IVI) 242
337
invoision symmoliy 60
opoialoi 60
ion moLilily 166
ano cuiionl 168
ion liansfoi 174
II spoclioscopy 281
ano lho chaiacloi laLlo 282
isolhoimal compiossiLilily 274
isolhoimal oxpansion 270
Joulo oxpansion 282
Joulo-1homson oxpansion 288
kinolic lhooiy of gasos 20
Lagiango mullipliois 0
Laguoiio polynominals 40
laminai ow 804
law of coiiosponoing slalos 206
law of ioclilinoai oiamolois 208
Logonoia polynomials 200
linoai comLinalions of alomic oi-
Lilals (LCAO) 72
Loionz numLoi 807
many oloclion alom
Hamlilonian foi 0
maximal woik 118
Maxwoll iolalions 140
Maxwoll`s oisliiLulion of spooos 22,
200
moan fioo palh 28, 200
moan ionic aclivily 162
moan ionic aclivily coo!cionl 162
molhoo of inilial volocilios 20
molhoo of isolalion 20
micioslalo 00
Mio scalloiing 84
miiioi plano symmoliy 70
molai hoal capacily 11
moloculai collisions
simplo moool foi 22
moloculai hyoiogon ion 67
Hamillonian foi 67
moloculai oiLilal oiagiam 76
moloculai oiLilals 68
moloculai iolalions 28
asymmoliic lops 280
conliifugal sliolching 286
linoai lops 288
polyalomic moloculos 287
sphoiical lops 280
symmoliic lops 288
viLialional slalo ooponoonco of
286
moloculai viLialions 228
moloculo
Sciooingoi oqualion foi 78
momonlum opoialoi
Moiso oscillaloi 64
onoigy lovols foi 6, 86
Schiooingoi oqualion foi 6
wavofunclion foi 6
Moiso polonlial 64, 86, 240
foico conslanl associaloo wilh
0
1ayloi soiios oxpansion of 8
noimal mooos 220
opoialoi
Hoimilian 180
laoooi 10
linoai 180
symmoliy 222
opoialoi algoLia 187
oiionlalion quanlum numLoi 8
oilhogonalily 101
ovoilono liansilions 66
paiamolois
oxlonsivo 100
inlonsivo 100
338
pailiclo in a Lox 81, 181
onoigy lovols 188
onoigy lovols foi 84, 44, 218
foaluios of lho onoigy lovols 8
noimalizalion conslanl foi 88
polonial onoigy 81
Schiooingoi oqualion foi 82
lhioo oimonsional 188
lhioo oimonsional onoigy lovols
18
lhioo oimonsional wavofunclion
18
wavofunclion foi 188
wavofunclions foi 84, 44, 218
pailiclo on a iing 104
Lounoaiy conoilions 104
onoigy lovols foi 10, 218
Hamilonian foi 104
wavofunclions foi 10, 218
paililion coo!cionl 14
ano oiug oolivoiy 1
foi lho Loin moool 174
paililion funclion
canonical 06, 181
oloclionic 101
giano canonical 07
isolhoimalisoLaiic 07
miciocanonical 06
moloculai 100
iolalional 101, 182
lianslalional 101, 182
viLialional 101, 182
Iauli oxclusion piinciplo 6
consquoncos of 8
poiluiLalion lhooiy 207
oxamplo of lho quailic oscilla-
loi 208
phaso oiagiam 800
Ioisouillo`s foimula 804
polaiizaLilily 70
poslulalo I (of quanlum mochan-
ics) 22
poslulalo II (of quanlum mochan-
ics) 24
poslulalo III (of quanlum mochan-
ics) 2
piossuio 104
piinciplo of Clausius 12, 824
piinciplo quanlum numLoi 2
pioLaLilily ampliluoo 22
pioLaLilily oisliiLulion 22
IV woik 111, 188
Iaman scalloiing 80
Iaman spoclioscopy 66, 288
ano lho chaiacloi laLlo 284
Iaoull`s law 811, 812, 814, 884
oovialions fiom 81
iofoionco slalo 81
ialo law 2
ialo laws 24
ooloiminalion of 28
inlogialoo 20
Iayloigh scalloiing 80
Iayloigh scalloiing law 81, 82, 87
ioaclion volocily 2, 201
iocipiocal iulo 108
ioo sunsols 82
Ioolich-Kwang gas 270
iofoionco slalos 147
iolalionship Lolwoon CI ano CV
180, 276
iolaxalion oocls 167
iigio ioloi 200
oogonoiacy of 28, 248
onoigy 28, 247
iolalional onoigy lovols 200, 210
339
oogonoiacy of 200
iolalional Hamillonian 200
iulo of mulual oxclusion 284
IyoLoig conslanl 20
SA1I 120
Schiooingoi oqualion
limo ooponoonl 214
limo inooponoonl 27
socono law
local" foimulalion 824
socono law of lhoimooynamics 126,
188
slalomonls of 127
simplo collision lhooiy 262
Slaloi ooloiminanl 8
foi lilhium 0
solai syslom moool 17
solvalion 160
solvophoLic oocl 176
spocic hoal 11
sphoiical haimonic funclions 48, 200
spin 201
quanlum numLoi 1, 8
wavofunclion 1
spin oiionlalion
quanlum numLoi 1, 8
spin-oiLil
coupling 20
Hamillonian 20
inloiaclion onoigy 20
sponlanoous piocoss 142
slalo funclion 121
laLlo of impoilanl onos 186
Sloilings appioximalion 02
Sloko`s law 167, 804
S1I 120
supoiposilion 101
sysloms
lypos of 108
lompoialuio 11
loim symLols 204
lhoimal conouclivily 801
of gasos 806
of liquios 806
lhoimal oquiliLiium 120
lhiio law of lhoimooynamics 128,
188
lips foi solving pioLloms 2
lolal ooiivalivo 107
liansfoi maliix 11
liiplo poinl 800
lwo lovol syslom 211
lofl` ano iighl` slalos 218, 210
Hamillonian foi 212
1ynoall scalloiing 84
ungoiaoo 60
van ooi Waals oqualion
340

Вам также может понравиться