Skip to main content

Full text of "Introduction To Chemical Physics"

See other formats


THE  BOOK  WAS 
DRENCHED 


< 

=  CD 


OU  164180 


INTERNATIONAL  SERIES  IN  PHYSICS 
LEE  A.  DtrBRIDGE,  CONSULTING  EDITOR 


INTRODUCTION 

TO 

CHEMICAL  PHYSICS 


INTERNATIONAL    SERIES    IN 
PURE   AND   APPLIED    PHYSICS 

G.  P.  HARNWELL,  Consulting  Editor 


BRILLOUIN— WAVE  PROPAGATION  IN  PERIODIC  STRUCTURES 

CADY— PIEZOELECTRICITY 

CLARK— APPLIED  X-RAYS 

CURTIS— ELECTRICAL  MEASUREMENTS 

DAVEY— CRYSTAL  STRUCTURE  AND  ITS  APPLICATIONS 

EDWARDS— ANALYTIC  AND  VECTOR  MECHANICS 

GURNEY— INTRODUCTION  TO  STATISTICAL  MECHANICS 

HARDY  AND  PERRIN— THE  PRINCIPLES  OF  OPTICS 

HARNWELL— ELECTRICITY  AND  ELECTROMAGNETISM 

HARNWELL  AND  LIVINGOOD— EXPERIMENTAL  ATOMIC  PHYSICS 

HOUSTON— PRINCIPLES  OF  MATHEMATICAL  PHYSICS 

HUGHES  AND  DUBRIDGE— PHOTOELECTRIC  PHENOMENA 

HUND— HIGH-FREQUENCY  MEASUREMENTS 

INGERSOLL,  ZOBEL,  AND  INGERSOLL — HEAT  CONDUCTION 

KEMBLE— PRINCIPLES  OF  QUANTUM  MECHANICS 

KENNARD— KINETIC  THEORY  OF  GASES 

KOLLER— THE  PHYSICS  OF  ELECTRON  TUBES 

MORSE— VIBRATION  AND  SOUND 

PAULING  AND  GOUDSMIT— THE  STRUCTURE  OF  LINE  SPECTRA 

RICHTMYER  AND  KENNARD— INTRODUCTION  TO  MODERN  PHYSICS 

RUARK  AND  UREY— ATOMS,  MOLECULES,  AND  QUANTA 

SCHIFF— QUANTUM  MECHANICS 

SEITZ— THE  MODERN  THEORY  OF  SOLIDS 

SLATER— INTRODUCTION  TO  CHEMICAL  PHYSICS 

MICROWAVE  TRANSMISSION 
SLATER  AND  FRANK — ELECTROMAGNETISM 

INTRODUCTION  TO  THEORETICAL  PHYSICS 

MECHANICS 

SMYTHE— STATIC  AND  DYNAMIC  ELECTRICITY 
STRATTON— ELECTROMAGNETIC  THEORY 
WHITE— INTRODUCTION  TO  ATOMIC  SPECTRA 
WILLIAMS— MAGNETIC  PHENOMENA 


Dr.  Lee  A.  DuBridge  was  consulting  editor  of  the  series  from  1939  to  1946. 


INTRODUCTION 

TO 

CHEMICAL  PHYSICS 


BY 

J.  C.  SLATER 

Professor  of  Physics 
Massachusetts  Institute  of  Technology 


FIRST  EDITION" 
SEVENTH  IMPRESSION 


McGRAW-HILL  BOOK  COMPANY,  INC. 

NEW    YORK   AND    LONDON 
1939 


(\)PYRK8HT,   1939,  BY     TIIK 

i  BOOK  COMPANY,  INC. 


PRINTED    IN"    THE    UNITED    STATES    OP    AMERICA 

All  rights  reserved.   This  book,  or 

parts  thereof,  may  not  be  reproduced 

in  any  form  without  permission  of 

the  publishers. 


THE   MAPLE    PRESS    COMPANY,   YORK, 


PREFACE 

It  is  probably  unfortunate  that  physics  and  chemistry  over  were 
separated.  Chemistry  is  the  science  of  atoms  and  of  the  way  they  com- 
bine. Physics  deals  with  the  interatomic  forces  and  with  the  large-scale 
properties  of  matter  resulting  from  those  forces.  So  long  as  chemistry 
was  largely  empirical  and  nonmathematical,  and  physics  had  not  learned 
how  to  treat  small-scale  atomic  forces,  the  two  sciences  seemed  widely 
separated.  But  with  statistical  mechanics  and  the  kinetic  theory  on  the 
one  hand  and  physical  chemistry  on  the  other,  the  two  sciences  began  to 
come  together.  Now  that  statistical  mechanics  has  led  to  quantum  theory 
and  wave  mechanics,  with  its  explanations  of  atomic  interactions,  there  is 
really  nothing  separating  them  any  more.  A  few  years  ago,  though  their 
ideas  were  close  together,  their  experimental  methods  were  still  quite 
different :  chemists  dealt  with  things  in  test  tubes,  making  solutions,  pre- 
cipitating and  filtering  and  evaporating,  while  physicists  measured  every- 
thing with  galvanometers  and  spectroscopes.  But  even  this  distinction 
has  disappeared,  with  more  and  more  physical  apparatus  finding  its  way 
into  chemical  laboratories. 

A  wide  range  of  study  is  common  to  both  subjects.  The  sooner  we 
realize  this  the  better.  For  want  of  a  better  name,  since  Physical 
Chemistry  is  already  preempted,  we  may  call  this  common  field  Chemical 
Physics.  It  is  an  overlapping  field  in  which  both  physicists  and  chemists 
should  be  trained.  There4  seems  no  valid  reason  why  their  training  in  it 
should  differ.  This  book  is  an  attempt  to  incorporate  some  of  the 
material  of  this  common  field  in  a  unified  presentation. 

What  should  be  included  in  a  discussion  of  chemical  physics?  Logi- 
cally, we  should  start  with  fundamental  principles.  We  should  begin 
with  mechanics,  then  present  electromagnetic  theory,  and  should  work 
up  to  wave  mechanics  and  quantum  theory.  By  means  of  these  wre 
should  study  the  structure  of  atoms  and  molecules.  Then  we  should 
introduce  thermodynamics  and  statistical  mechanics,  so  as  to  handle 
large  collections  of  molecules.  With  all  this  fundamental  material  we 
could  proceed  to  a  discussion  of  different  types  of  matter,  in  the  solid, 
liquid,  and  gaseous  phases,  and  to  an  explanation  of  its  physical  and 
chemical  properties  in  terms  of  first  principles.  But  if  we  tried  to  do  all 
this,  we  should,  in  the  first  place,  be  writing  several  volumes  which  would 
include  almost  all  of  theoretical  physics  and  chemistry;  and  in  the 
second  place  no  one  but  an  experienced  mathematician  could  handle  the 


vi  PREFACE 

theory.  For  both  of  these  reasons  the  author  has  compromised  greatly 
in  the  present  volume,  so  as  to  bring  the  material  into  reasonable  com- 
pass and  to  make  it  intelligible  to  a  reader  with  a  knowledge  of  calculus 
and  differential  equations,  but  unfamiliar  with  the  more  difficult  branches 
of  mathematical  physics. 

In  the  matter  of  scope,  most  of  the  theoretical  physics  which  forms  a 
background  to  our  subject  has  been  omitted.  Much  of  this  is  considered 
in  the  companion  volume,  " Introduction  to  Theoretical  Physics,"  by 
Slater  and  Frank.  The  effort  has  been  made  in  the  present  work  to  pro- 
duce a  book  which  is  intelligible  without  studying  theoretical  physics 
first.  This  has  been  done  principally  for  the  benefit  of  chemists  and 
others  who  wish  to  obtain  the  maximum  knowledge  of  chemical  physics 
with  the  minimum  of  theory.  In  the  treatment  of  statistical  mechanics 
only  the  most  elementary  use  of  mechanics  is  involved.  For  that  reason 
it  has  not  boon  possible  to  give  a  complete  discussion,  although  the  parts 
used  in  the  calculations  have  been  considered.  Statistical  mechanics  has 
been  introduced  from  the  standpoint  more  of  the  quantum  theory  than  of 
classical  theory,  but  the  quantum  theory  that  is  used  is  of  a  very  elemen- 
tary sort.  It  has  seemed  desirable  to  omit  wave  mechanics,  which 
demands  more  advanced  mathematical  methods.  In  discussing  atomic 
and  molecular  structure  and  the  nature  of  interatomic  forces,  descriptive 
use  has  boon  made  of  the  quantum  theory,  but  again  no  detailed  use  of  it. 
Thus  it  is  hoped  that  the  reader  with  only  a  superficial  acquaintance  with 
modern  atomic  theory  will  be  able  to  read  the  book  without  great  diffi- 
culty, although,  of  course,  the  reader  with  a  knowledge  of  quantum 
theory  and  wave  mechanics  will  have  a  great  advantage. 

Finally  in  the  matter  of  arrangement  the  author  has  departed  from 
the  logical  order  in  the  interest  of  easy  presentation.  Logically  one 
should  probably  begin  with  the  structure  of  atoms  and  molecules,  crystals 
and  liquids  and  gases;  then  introduce  the  statistical  principles  that 
govern  molecules  in  large  numbers,  and  finally  thermodynamics,  which 
follows  logically  from  statistics.  Actually  almost  exactly  the  opposite 
order  has  been  chosen.  Thermodynamics  and  statistical  mechanics  come 
first.  Then  gases,  solids,  and  liquids  are  treated  on  the  basis  of  thermo- 
dynamics and  statistics,  with  a  minimum  amount  of  use  of  a  model. 
Finally  atomic  and  molecular  structure  are  introduced,  together  with  a 
discussion  of  different  types  of  substances,  explaining  their  interatomic 
forces  from  quantum  theory  and  their  thermal  and  elastic  behavior  from 
our  thermodynamic  and  statistical  methods.  In  this  way,  the  historical 
order  is  followed  roughly,  and,  at  least  for  chemists,  it  proceeds  from 
what  are  probably  the  more  familiar  to  the  less  familiar  methods. 

It  is  customary  to  write  books  either  on  thermodynamics  or  on 
statistical  mechanics;  this  one  combines  both.  It  seems  hardly  necessary 


PREFACE  Vll 

to  apologize  for  this.  Both  have  their  places,  and  both  are  necessary  in  a 
complete  presentation  of  chemical  physics.  An  effort  has  been  made  to 
keep  them  separate,  so  that  at  any  time  the  reader  will  be  clear  as  to 
which  method  is  being  used.  In  connection  with  thermodynamics,  the 
method  of  Bridgman,  which  seems  by  far  the  most  convenient  for  prac- 
tical application,  has  been  used. 

There  is  one  question  connected  with  thermodynamics,  that  of 
notation.  The  continental  notation  and  the  American  chemical  notation 
of  Lewis  and  Randall  are  quite  different.  Each  has  its  drawbacks.  The 
author  has  chosen  the  compromise  notation  of  the  Joint  Committee  of 
the  Chemical  Society,  the  Faraday  Society,  and  the  Physical  Society  (all 
of  England),  which  preserves  the  best  points  of  both.  It  is  hoped  that 
this  notation,  which  has  a  certain  amount  of  international  sanction,  may 
become  general  among  both  physicists  and  chemists,  whose  poblems  are 
similar  enough  so  that  they  surely  can  use  the  same  language. 

In  a  book  like  this,  containing  a  number  of  different  types  of  material, 
it  is  likely  that  some  readers  and  teachers  will  want  to  use  some  parts, 
others  to  use  other  parts.  An  attempt  has  been  made  to  facilitate  such 
use  by  making  chapters  and  sections  independent  of  each  other  as  far  as 
possible.  The  book  has  been  divided  into  three  parts:  Part  I,  Thermo- 
dynamics, Statistical  Mechanics,  and  Kinetic  Theory;  Part  II,  Gases, 
Liquids,  and  Solids;  Part  III,  Atoms,  Molecules,  and  the  Structure  of 
Matter.  The  first  part  alone  forms  an  adequate  treatment  of  thermo- 
dynamics and  statistical  theory,  and  could  be  used  by  itself.  Certain  of 
its  chapters,  as  Chap.  V  on  the  Fermi-Dirac  and  Einstein-Bose  Statistics, 
Chap.  VI  on  the  Kinetic  Method  and  the  Approach  to  Thermal  Equilib- 
rium, and  Chap.  VII  on  Fluctuations,  can  be  omitted  without  causing 
much  difficulty  in  reading  the  following  parts  of  the  book  (except  for  the 
chapters  on  metals,  which  depend  on  the  Fermi-Dirac  statistics).  In 
Part  II,  most  of  Chap.  IX  on  the  Molecular  Structure  and  Specific  Heat 
of  Polyatomic  Gases,  Chap.  X  on  Chemical  Equilibrium  in  Gases,  parts 
of  Chap.  XII  on  Van  der  Waals'  Equation  and  Chap.  XIII  on  the 
Equation  of  State  of  Solids,  Chap.  XV  on  The  Specific  Heat  of  Com- 
pounds, Chap.  XVII  on  Phase  Equilibrium  in  Binary  Systems,  and 
Chap.  XVIII  on  Phase  Changes  of  the  Second  Order  are  not  necessary 
for  what  follows.  In  Part  III,  Chap.  XIX  on  Radiation  and  Matter, 
Chap.  XX  on  lonization  and  Excitation  of  Atoms,  and  Chap.  XXI  on 
Atoms  and  the  Periodic  Table  will  be  familiar  to  many  readers.  Much 
of  the  rest  of  this  part  is  descriptive;  one  chapter  does  not  depend  on 
another,  so  that  many  readers  may  choose  to  omit  a  considerable  portion 
or  all,  of  this  material.  It  will  be  seen  from  this  brief  enumeration  that 
selections  from  the  book  may  be  used  in  a  variety  of  ways  to  serve  the 
needs  of  courses  less  extensive  than  the  whole  book. 


viii  PREFACE 

The  author  hopes  that  this  book  may  serve  in  a  minor  way  to  fill  the 
gap  that  has  grown  between  physics  and  chemistry.  This  gap  is  a 
result  of  tradition  and  training,  not  of  subject  matter.  Physicists  and 
chemists  are  given  quite  different  courses  of  instruction;  the  result  is 
that  almost  no  one  is  really  competent  in  all  the  branches  of  chemical 
physics.  If  the  coming  generation  of  chemists  or  physicists  could  receivo 
training,  in  the  first  place,  in  empirical  chemistry,  in  physical  chemistry, 
in  metallurgy,  and  in  crystal  structure,  and,  in  the  second  place,  in 
theoretical  physics,  including  mechanics  and  electromagnetic  theory,  and 
in  particular  in  quantum  theory,  wave  mechanics,  and  the  structure  of 
atoms  and  molecules,  and  finally  in  thermodynamics,  statistical 
mechanics,  and  what  we  have  called  chemical  physics,  they  would  be 
far  better  scientists  than  those  receiving  the  present  training  in  either 
chemistry  or  physics  alone. 

The  author  wishes  to  indicate  his  indebtedness  to  several  of  his 
colleagues,  particularly  Professors  B.  E.  Warren  and  W.  B.  Nottingham, 
who  have  read  parts  of  the  manuscript  and  made  valuable  comments. 
His  indebtedness  to  books  is  naturally  vory  great,  but  most  of  thorn  arc 
mentioned  in  the  list  of  suggested  references  at  the  end  of  this  volume. 

J.  C.  SLATER. 
CAMBRIDGE,  MASSACHUSETTS, 
September,  1939. 


CONTENTS 


PA  an 

PREFACE .        .  .  v 


PART  I 

THERMODYNAMICS,  STATISTICAL  MECHANICS, 
AND  KINETIC  THEORY 

CHAPTER  I 

HEAT  AS  A  MODE  OF  MOTION 

INTRODUCTION              ....  .3 

1.  THE  CONSERVATION  OF  ENERGY  ....                   .  .3 

2.  INTERNAL  ENERGY,  EXTERNAL  WORK,  AND  HEAT  FLOW  ...           6 

3.  THE  ENTROPY  AND  IRREVERSIBLE  PROCESSES  .        .           9 

4.  THE  SECOND  LAW  OF  THERMODYNAMICS  .    .    .  ...      12 

CHAPTER  II 

THERMODYNAMICS 

INTRODUCTION .    .  16 

1.  THE  EQUATION  OF  STATE  ...  .  16 

2.  THE  ELEMENTARY  PARTIAL  DERIVATIVES  ....  18 

3.  THE  ENTHALPY,  AND  HELMHOLTZ  AND  GIBBS  FREE  ENERGIES  20 

4.  METHODS  OF  DERIVING  THERMODYNAMIC  FORMULAS  23 

5.  GENERAL  CLASSIFICATION  OF  THERMODYNAMIC  FORMULAS  27 

6.  COMPARISON  OP  THERMODYNAMIC  AND  GAS  SCALES  OF  TEMPERATURE  30 

CHAPTER  III 

STATISTICAL  MECHANICS 

INTRODUCTION .                    .  .            .32 

1.  STATISTICAL  ASSEMBLIES  AND  THE  ENTROPY  ...                 .  32 

2.  COMPLEXIONS  AND  THE  PHASE  SPACE 36 

3.  CELLS  IN  THE  PHASE  SPACE  AND  THE  QUANTUM  THEORY  ....  38 

4.  IRREVERSIBLE  PROCESSES .    .               .        .  43 

5.  THE  CANONICAL  ASSEMBLY 46 

CHAPTER  IV 

THE  MAXWELL-BOLTZMANN  DISTRIBUTION  LAW 

INTRODUCTION 52 

1.  THE  CANONICAL  ASSEMBLY  AND  THE  MAXWELL-BOLTZMANN  DISTRIBUTION  .  52 

2.  MAXWELL'S  DISTRIBUTION  OF  VELOCITIES .  55 

3t  THE  EQUATION  OF  STATE  AND  SPECIFIC  HEAT  OF  PERFECT  MONATOMIC  GASES  58 

4.  THE  PERFECT  GAS  IN  A  FORCE  FIELD ....  62 

ix 


X  CONTENTS 

PAOB 
CHAPTER  V 

THE  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS 

INTRODUCTION  .          65 

1.  THE  MOLECULAR  PHASE  SPACE 65 

2.  ASSEMBLIES  IN  THE  MOLECULAR  PHASE  SPACE  .    .  68 

3.  THE  FERMI-DIRAC  DISTRIBUTION  FUNCTION 72 

4.  THERM ODYNAMIC  FUNCTIONS  IN  THE  FERMI  STATISTICS 76 

5.  THE  PERFECT  GAS  IN  THE  FERMI  STATISTICS.    ...  80 

6.  THE  EINSTEIN-ROSE  DISTRIBUTION  LAW 83 

CHAPTER  VI 
THE  KINETIC  METHOD  AND  THE  APPROACH  TO 

THERMAL  EQUILIBRIUM 
INTRODUCTION 86 

1.  THE  EFFECT  OF  MOLECULAR  COLLISIONS  ON  THE  DISTRIBUTION  FUNCTION  IN 

THE  BOLTZMANN  STATISTICS 86 

2.  THE  EFFECT  OF  COLLISIONS  ON  THE  ENTROPY  ...               ...  89 

3.  THE  CONSTANTS  IN  THE  DISTRIBUTION  FUNCTION 92 

4.  THE  KINETIC  METHOD  FOR  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS  .     96 

CHAPTER  VII 
FLUCTUATIONS 
INTRODUCTION .101 

1.  ENERGY  FLUCTUATIONS  IN  THE  CANONICAL  ASSEMBLY  .    .  101 

2.  DISTRIBUTION  FUNCTIONS  FOR  FLUCTUATIONS 104 

3.  FLUCTUATIONS  OF  ENERGY  \ND  DENSITY 107 


PART  II 
GASES,  LIQUIDS,  AND  SOLIDS 

CHAPTER  VIII 
THERMODYNAMIC  AND  STATISTICAL  TREATMENT  OF  THE  PERFECT 

GAS  AND  MIXTURES  OF  GASES 
INTRODUCTION 115 

1.  THERMODYNAMICS  OF  A  PERFECT  GAS 115 

2.  THERMODYNAMICS  OF  A  MIXTURE  OF  PERFECT  GASES 120 

3    STATISTICAL  MECHANICS  OF  A  PERFECT  GAS  IN  BOLTZMANN  STATISTICS   .    .    .124 

CHAPTER  IX 
THE  MOLECULAR  STRUCTURE  AND  SPECIFIC  HEAT  OF 

POLYATOMIC  GASES 

INTRODUCTION 130 

1.  THE  STRUCTURE  OF  DIATOMIC  MOLECULES 130 

2.  THE  ROTATIONS  OF  DIATOMIC  MOLECULES 134 

3.  THE  PARTITION  FUNCTION  FOR  ROTATION 138 

4.  THE  VIBRATION  or  DIATOMIC  MOLECULES 140 

5.  TH»  PARTITION  FUNCTION  FOR  VIBRATION 142 

6.  THE  SPECIFIC  HEATS  OF  POLYATOMIC  GASES 145 


CONTENTS  xi 

PAGB 
CHAPTER  X 

CHEMICAL  EQUILIBRIUM  IN  GASES 

INTRODUCTION 150 

1.  RATES  OP  REACTION  AND  THE  MASS  ACTION  LAW 151 

2.  THE  EQUILIBRIUM  CONSTANT,  AND  VAN'T  HOFF'S  EQUATION 154 

3.  ENERGIES  OF  ACTIVATION  AND  THE  KINETICS  OF  REACTIONS .158 

CHAPTER  XI 

THE  EQUILIBRIUM  OF  SOLIDS,  LIQUIDS,  AND  GASES 

INTRODUCTION  .       .              100 

1.  THE  COEXISTENCE  OF  PHASES 166 

2.  THE  EQUATION  OF  STATE          169 

3.  ENTROPY  AND  GIBBS  FREE  ENERGY 170 

4.  THE  LATENT  HEATS  AND  CLAPEYRON'S  EQUATION 174 

5.  THE  INTEGRATION  OF  CLAPEYRON'S  EQUATION  AND  THE  VAPOR  PRESSURE 

CURVE 176 

6.  STATISTICAL  MECHANICS  AND  THE  VAPOR  PRESSURE  CURVE  .  .178 

7.  POLYMORPHIC  PHASES  OF  SOLIDS 180 

CHAPTER  XII 

VAN  DER  WAALS'  EQUATION 

INTRODUCTION  ...       .           182 

1.  VAN  DER  WAALS'  EQUATION      182 

2.  ISOTHERMALS  OF  VAN  DER  WAALS' EQUATION .  184 

3.  GIBBS  FREE  ENERGY  AND  THE  EQUILIBRIUM  OF  PHASES  FOR  A  VAN  DER  WAALS 

GAS 187 

4.  STATISTICAL  MECHANICS  AND  THE  SECOND  VIRIAL  COEFFICIENT 190 

5.  THE  ASSUMPTIONS  OF  VAN  DER  WAALS'  EQUATION       .                   .  194 

6.  THE  JOULE-THOMSON  EFFECT  AND  DEVIATIONS  FROM  THE  PERFECT  GAS  LAW  196 

CHAPTER  XIII 

THE  EQUATION  OF  STATE  OF  SOLIDS 

INTRODUCTION .  199 

1.  THE  EQUATION  OF  STATE  AND  SPECIFIC  HEAT  OF  SOLIDS            199 

2.  THERMODYNAMIC  FUNCTIONS  FOR  SOLIDS.    .    .                         ...               .  205 

3.  THE  STATISTICAL  MECHANICS  OF  SOLIDS.   .                             ....  211 

4.  STATISTICAL  MECHANICS  OF  A  SYSTEM  OF  OSCILLATORS 215 

5.  POLYMORPHIC  TRANSITIONS 220 

CHAPTER  XIV 

DEBYE'S  THEORY  OF  SPECIFIC  HEATS 

INTRODUCTION .  222 

1.  ELASTIC  VIBRATIONS  OF  A  CONTINUOUS  SOLID .  222 

2.  VlBRATIONAL  FREQUENCY  SPECTRUM  OF  A  CONTINUOUS  SOLID 225 

3.  DEBYE'S  THEORY  OF  SPECIFIC  HEATS          234 

4.  DEBYE'S  THEORY  AND  THE  PARAMETER  y 238 

CHAPTER  XV 

THE  SPECIFIC  HEAT  OF  COMPOUNDS 

INTRODUCTION 241 

1.  WAVE  PROPAGATION  IN  A  ONE-DIMENSIONAL  CRYSTAL  LATTICE 241 

2.  WAVES  IN  A  ONE-DIMENSIONAL  DIATOMIC  CRYSTAL 247 


xii  CONTENTS 

PAGE 

3.  VIBRATION  SPECTRA  AND  SPECIFIC  HEATS  OF  POLYATOMIC  CRYSTALS  .  .    .       252 

4.  INFRARED  OPTICAL  SPECTRA  OP  CRYSTALS  ....          .  254 

CHAPTER  XVI 

THE  LIQUID  STATE  AND  FUSION 

INTRODUCTION  ...                                                               .   256 

1.  THE  LIQUID  PHASE  .  .  .  256 

2.  THE  LATENT  HEAT  OF  MELTING                                    258 

3.  THE  ENTROPY  OF  MELTING  260 

4.  STATISTICAL  MECHANICS  AND  MELTING  265 

CHAPTER  XVII 

PHA8K  EQUILIBRIUM  IN  BINARY  SYSTEMS 

INTRODUCTION .    .               .        .  270 

1.  TYPES  OF  PHASES  IN  BINARY  SYSTEMS.                   .            271 

2.  ENERGY  AND  ENTROPY  OF  PHASES  OF  VARIABLE  COMPOSITION              .  275 

3.  THE  CONDITION  FOR  EQUILIBRIUM  BETWEEN  PHASES  278 

4.  PHASE  EQUILIBRIUM  BETWEEN  MUTUALLY  INSOLUBLE  SOLIDS  283 

5.  LOWERING  OF  MELTING  POINTS  OF  SOLUTIONS  .           288 

CHAPTER  XVIII 

PHASE  CHANGES  OF  THE  SECOND  ORDER 

INTRODUCTION  .  291 

1.  ORDER-DISORDEU  TRANSITIONS  IN  ALLOYS  293 

2.  EQUILIBRIUM  IN  TRANSITIONS  OF  THE  Cu-ZN  TYPE  .  296 

3.  TRANSITIONS  OF  THE  Cu-ZN  TYPE  WITH  ARBITRARY  COMPOSITION   .  301 


PART  III 
ATOMS,  MOLECULES,  AND  THE  STRUCTURE  OF  MATTER 

CHAPTER  XIX 

RADIATION  AND  MATTER 

INTRODUCTION   .  .  .    .  307 

1.  BLACK  BODY  RADIATION  AND  THE  STEFAN-BOLTZMANN  LAW      .    .  307 

2.  THE  PLANCK  RADIATION  LAW 313 

3.  EINSTEIN'S  HYPOTHESIS  AND  THE  INTERACTION  OF  RADIATION  AND  MATTER  316 

CHAPTER  XX 

IONIZATION  AND  EXCITATION  OF  ATOMS 

INTRODUCTION  .    .                                 .  32 J 

1.  BOHR'S  FREQUENCY  CONDITION  .       ...               .  321 

2.  THE  KINETICS  OF  ABSORPTION  AND  EMISSION  OF  RADIATION  .  .                      324 

3.  THE  KINETICS  OF  COLLISION  AND  IONIZATION  ....              326 

4.  THE  EQUILIBRIUM  OF  ATOMS  AND  ELECTRONS  .    .               .   333 

CHAPTER  XXI 
ATOMS  AND  THE  PERIODIC  TABLE 

INTRODUCTION  .  .  336 

1.  THE  NUCLEAR  ATOM ....  336 


CONTENTS  xiii 

PAGE 

2.  ELECTRONIC  ENERGY  LEVELS  OF  AN  ATOM  ...  338 

3.  THE  PERIODIC  TABLE  OF  THE  ELEMENTS .  344 

CHAPTER  XXII 

INTERATOMIC  AND  INTERMOLECULAR  FORCES 
INTRODUCTION  ...          352 

1.  THE  ELECTROSTATIC*  INTERACTIONS  BETWEEN  RIGID  ATOMS  OR  MOLECULES  AT 

LARGE  DISTANCES 353 

2.  THE   ELECTROSTATIC:   OR   COULOMB   INTERACTIONS  BETWEEN   OVERLAPPING 

RIGID  ATOMS 361 

3.  POLARIZATION  AND  INTERATOMIC  FORCES  .  363 

4.  EXCHANGE  INTERACTIONS  BETWEEN  ATOMS  AND  MOLECULES 367 

5.  TYPES  OF  CHEMICAL  SUBSTANCES.    ...  ...  .  375 

CHAPTER  XXIII 
IONIC  CRYSTALS 
INTRODUCTION  .  .    .          377 

1.  STRUCTURE  OF  SIMPLE  BIN \RY  IONIC  COMPOUNDS 377 

2.  IONIC  RADII 382 

3.  ENERGY  AND  EQUATION  OF  STATE  OF  SIMPLE  IONIC  LATTICES  AT  THE  ABSOLUTE 

ZERO .  385 

4.  THE  EQUATION  OF  STATE  OF  THE  ALKALI  HALIDES  390 

5.  OTHER  TYPES  OF  IONIC  CRYSTALS .           396 

6.  POL  ARIZ  ABILITY  AND  UNSYMMETRICAL  STRUCTURES   .  .                       398 

CHAPTER  XXIV 

THE  HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS 

INTRODUCTION  400 

1.  THE  HOMOPOLAU  BOND  .                   400 

2.  THE  STRUCTURE  OF  TYPICAL  HOMOPOLAR  MOLECULES.  .       402 

3.  GASEOUS  AND  LIQUID  PHASES  OF  HOMOPOLAR  SUBSTANCES  407 

4.  MOLECULAR  CRYSTALS .  414 

CHAPTER  XXV 

ORGANIC  MOLECULES  AND  THEIR  CRYSTALS 

INTRODUCTION ....                                          420 

1.  CARBON  BONDING  IN  ALIPATIIIC  MOLECULES        .           420 

2.  ORGANIC  RADICALS 425 

3.  THE  DOUBLE  BOND,  AND  THE  AROMVTIC  COMPOUNDS 428 

4.  ORGANIC  CRYSTALS .   432 

CHAPTER  XXVI 

HOMOPOLAR  BONDS  IN  THE  SILICATES 

INTRODUCTION 435 

1.  THE  SILICON-OXYGEN  STRUCTURE .   435 

2.  SILICON-OXYGEN  CHAINS.    .    .           .               .                    .       436 

3.  SILICON-OXYGEN  SHEETS 439 

4.  THREE-DIMENSIONAL  SILICON-OXYGEN  STRUCTURES .  .    .  441 

CHAPTER  XXVII 
METALS 

INTRODUCTION .  ...  444 

1.  CRYSTAL  STRUCTURES  OF  METALS .    .  444 


xiv  CONTENTS 

PAOB 

2.  ENERGY  RELATIONS  IN  METALS ...  450 

3.  GENERAL  PROPERTIES  OF  METALS .  456 

CHAPTER  XXVIII 

THERMIONIC  EMISSION  AND  THE  VOLTA  EFFECT 

INTRODUCTION 460 

1.  THE  COLLISIONS  OP  ELECTRONS  AND  METALS.    .                     460 

2.  THE  EQUILIBRIUM  OP  A  METAL  \ND  AN  ELECTRON  GAS          463 

3.  KINETIC  DETERMINATION  OP  THERMIONIC  EMISSION     .               .    .            .  464 

4.  CONTACT  DIFFERENCE  OF  POTENTIAL 467 

CHAPTER  XXIX 

THE  ELECTRONIC  STRUCTURE  OF  METALS 

INTRODUCTION .    .                       .  472 

1.  THE  ELECTRIC  FIELD  WITHIN  A  MKTA.L  472 

2.  THE  FREE  ELECTRON  MODEL  OF  A  METAL  475 

3.  THE  FREE  ELECTRON  MODEL  AND  THERMIONIC  EMISSION  480 

4.  THE  FREE  ELECTRON  MODEL  AND  ELECTRICAL  CONDUCTIVITY           .  484 

.5.  ELECTRONS  IN  A  PERIODIC  FORCE  FIELD .        ...  489 

6.  ENERGY  BANDS,  CONDUCTORS,  AND  INSULATORS                                   .    .  495 

PROBABLE  VALUES  OF  THE  GENERAL  PHYSICAL  CONSTANTS 503 

SUGGESTED  REFERENCES 505 

INDEX 507 


PART  I 

THERMODYNAMICS,  STATISTICAL  MECHANICS, 
AND  KINETIC  THEORY 


CHAPTER  I 
HEAT  AS  A  MODE  OF  MOTION 

Most  of  modern  physics  and  chemistry  is  based  on  three  fundamental 
ideas:  first,  matter  is  made  of  atoms  and  molecules,  very  small  and  very 
numerous;  second,  it  is  impossible  in  principle  to  observe  details  of  atomic 
and  molecular  motions  below  a  certain  scale  of  smallncss;  and  third, 
heat  is  mechanical  motion  of  the  atoms  and  molecules,  on  such  a  small 
scale  that  it  cannot  be  completely  observed.  The  first  and  third  of  these* 
ideas  are  products  of  the  last  century,  but  the  second,  the  uncertainty 
principle,  the  most  characteristic  result  of  the  quantum  theory,  has  arisen 
since  1900.  By  combining  these  three  principles,  we  have  the  theoretical 
foundation  for  studying  the  branches  of  physics  dealing  with  matter  and 
chemical  problems. 

1.  The  Conservation  of  Energy.-  -From  Newton's  second  law  of 
motion-one  can  prove  immediately  that  the  work  done  by  an  external 
force  cm  a  system  during  any  nuitioiMHj^i^s  Uin  increase  of  kinetic  emvrgy 
of  the  system.  This  can  be  stated  in  the  form 

(1.1) 

where  KE  stands  for  the  kinetic  energy,  dW  the  infinitesimal  clement  of 
work  done  on  the  system.  Certain  forces  are  called  conservative;  they 
have  the  property  that  the  work  done  by  them  when  the  system  goes  from 
an  initial  to  a  final  state  depends  only  on  the  initial  and  final  state,  not  on 
the  details  of  the  motion  from  one  state  to  the  other.  Stated  technically, 
we  say  that  the  work  done  between  two  end  points  depends  only  on  the 
end  points,  not  on  the  path.  A  typical  example  of  a  conservative  force 
is  gravitation;  a  typical  nonconservative  force  is  friction,  in  which  the 
longer  the  path,  the  greater  the  work  done.  For  a  conservative  force, 
we  define  the  potential  energy  as 


PE,  =  ~*W.  (1.2) 

This  gives  the  potential  energy  at  point  1,  as  the  negative  of  the  work  done 
in  bringing  the  system  from  a  certain  state  0  where  the  potential  energy 
is  zero  to  the  state  1,  an  amount  of  work  which  depends  only  on  the  points 
1  and  0,  not  on  the  path.  Then  we  have 

(1.3) 


4  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  I 

and,  combining  with  Eq.  (1.1), 

+  PEl  =  KE*  +  PE*  =  E,  (1.4) 


where,  since  1  and  2  are  arbitrary  points  along  the  path  and  KE  +  PE 
is  the  same  at  both  these  points,  we  must  assume  that  KE  +  PE  remains 
constant,  and  may  set  it  equal  to  a  constant  E,  the  total  energy.  This  is 
the  law  of  conservation  of  energy. 

To  avoid  confusion,  it  is  worth  while  to  consider  two  points  connected 
with  the  potential  energy:  the  negative  sign  which  appears  in  the  defi- 
nition (1.2),  and  the  choice  of  the  point  where  the  potential  energy  is 
zero.  Both  points  can  bo  illustrated  simply  by  the  case  of  gravity  acting 
on  bodies  near  the  earth.  Gravity  acts  down.  We  may  balance  its 
action  on  a  given  body  by  an  equal  and  opposite  upward  force,  as  by 
supporting  the  body  by  the  hand.  We  may  then  define  the  potential 
energy  of  the  body  at  height  h  as  the  work  done  by  this  balancing  force 
in  raising  the  body  through  this  height.  Thus  if  the  mass  of  the  body  is 
m,  and  the  acceleration  of  gravity  g,  tho  force  of  gravity  is  —mg  (positive 
directions  being  upward),  the  balancing  force  is  +  nig,  and  the  work  dono 
by  the  hand  in  raising  the  mass  through  height  h  is  mgh,  which  we  define 
as  the  potential  energy.  Tho  negative  sign,  then,  comes  because  the 
potential  energy  is  defined,  not  as  the  work  done  by  the  force  we  are 
interested  in,  but  the  work  done  by  an  equal  and  opposite  balancing 
force.  As  for  the  arbitrary  position  whero  wo  choose  the  potential  energy 
to  bo  zero,  that  appears  in  this  example  because  \ve  nan  measure  our 
height  h  from  any  level  \\o  choose.  It  is  important  to  notice  that  tho 
same  arbitrary  constant  appears  essentially  in  the  energy  E.  Thus,  in 
Eq.  (1.4),  if  we  chose  to  redefine  our  zero  of  potential  energy,  we  should 
have  to  add  a  constant  to  the  total  energy  at  each  point  of  the  path. 
Another  way  of  stating  this  is  that  it  is  only  the  difference  E  —  PE  whose 
magnitude  is  determined,  neither  the  total  energy  nor  the  potential  energy 
separately.  For  E  —  PE  is  the  kinetic  energy,  which  alono  can  be  deter- 
mined by  direct  experiment,  from  a  measurement  of  velocities. 

Most  ftf*t.iifl.l  f(frf»pK  fiTn  Tint,  fionfiftf  vn.ii  vfi*  f^p  in  fl.ln^ost  fill  n|*Rot,ir*fl.l 
cases  there  is  friction  of  one  sort  or  another-  _  And  yo.t  thor  la^f  rontnrv 
has  seen  tho  conservation  of  'onofgy  -built.  np_fio  that,  H  iff  HOW  regarded  as 
the  most  important  principle  of  physics.  The  first  step  in  this  develop- 
merit  was  the  mechanical  theory  of  heat,  the  sciences  of  thermodynamics 
and  statistical  mechanics.  Heat  had  for  many  years  been  considered  as  a 
fluid,  sometimes  called  by  the  name  caloric,  which  was  abundant  in  hot 
bodies  and  lacking  in  cold  ones.  This  theory  is  adequate  to  explain 
calorimetry,  the  science  predicting  the  final  temperature  if  substances  of 
different  initial  temperatures  are  mixed.  Mixing  a  cold  body,  lacking  in 
caloric,  with  a  hot  one,  rich  in  it,  leaves  the  mixture  with  a  medium 


SEC.  1]  HEAT  AS  A  MODE  OF  MOTION  5 

amount  of  heat,  sufficient  to  raise  it  to  an  intermediate  temperature. 
But  early  in  the  nineteenth  century,  difficulties  with  the  theory  began  to 
appear.  As  we  look  back,  we  can  sec  that  these  troubles  came  from  tho 
implied  assumption  that  the  caloric,  or  heat,  was  conserved.  In  a 
calorimetric  problem,  some  of  the  caloric  from  the  hot  body  flows  to  the 
cold  one,  leaving  both  at  an  intermediate  temperature,  but  no  caloric  is 
lost.  It  was  naturally  supposed  that  this  conservation  was  universal. 
The  difficulty  with  this  assumption  may  be  soon  as  clearly  as  anywhere  in 
Rumford's  famous  observation  on  the  boring  of  cannon.  Rumford 
noticed  that  a  great  deal  of  heat  was  given  off  in  the  process  of  boring. 
The  current  explanation  of  this  was  that  the  chips  of  metal  had  their  heal, 
capacity  reduced  by  the  process  of  boring,  so  that  tho  heat  which  was 
originally  present  in  them  was  able  to  raise  them  to  a  higher  temperature. 
Rumford  doubted  this,  and  to  demonstrate  it  ho  used  a  very  blunl  tool, 
which  hardly  removed  any  chips  at  all  and  yet  produced  oven  moio  heat 
than  a  sharp  tool.  He  showed  by  his  experiments  beyond  any  doubt 
that  heat  could  be  produced  continuously  and  in  apparently  unlimited 
quantity,  by  the  friction.  Surely  this  was  impossible  if  heat,  or  caloric, 
were  a  fluid  which  was  conserved.  And  his  conclusion  stated  essentially 
our  modern  view,  that  heat  is  really  a  form  of  energy,  convertible  into 
energy.  In  his  words:1 

What  is  Heat?  Is  there  any  such  thing  as  an  igneous  fluid?  Is  there  any 
thing  that  can  with  propriety  be  called  caloric?  ...  In  reasoning  on  this  subject, 
we  must  not  forget  to  consider  that  most  remarkable  circumstance,  that  the  source 
of  Heat  generated  by  friction,  in  these  Experiments,  appeared  evidently  to  be 
inexhaustible. 

It  is  hardly  necessary  to  add,  that  any  thing  which  any  insulated  body,  or 
system  of  bodies,  can  continue  to  furnish  without  limitation,  cannot  possibly  be 
a  material  substance;  and  it  appears  to  me  to  be  extremely  difficult,  if  not  quite 
impossible,  to  form  any  distinct  idea  of  any  thing,  capable  of  being  excited  and 
communicated,  in  the  manner  the  Heat  was  excited  and  communicated  in  these 
experiments,  except  it  be  MOTION. 

From  this  example,  it  is  clear  that  both  conservation  laws  broke  down 
at  once.  In  a  process  involving  friction,  energy  is  not  conserved,  but 
rather  disappears  continually.  At  the  same  time,  however,  heat  is  not 
conserved,  but  appears  continually.  Rumford  essentially  suggested  that 
the  heat  which  appeared  was  really  simply  the  energy  which  had  dis- 
appeared, observable  in  a  different  form.  This  hypothesis  was  not  really 
proved  for  a  good  many  years,  however,  until  Joule  made  his  experiments 
on  the  mechanical  equivalent  of  heat,  showing  that  when  a  certain 
amount  of  work  or  mechanical  energy  disappears,  the  amount  of  heat 

1  Quoted  from  W.  F.  Magie,  "Source  Book  in  Physics,"  pp.  160-161,  McGraw- 
Hill  Book  Company,  Inc.,  1935. 


6  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  I 

appearing  is  always  the  same,  no  matter  what  the  process  of  trans- 
formation may  be.  The  calorie,  formerly  considered  as  a  unit  for 
measuring  the  amount  of  caloric  present,  was  seen  to  be  really  a  unit  of 
energy,  convertible  into  ergs,  the  ordinary  units  of  energy.  And  it 
became  plain  that  in  a  process  involving  friction,  there  really  was  no  loss 
of  energy.  The  mechanical  energy,  it  is  true,  decreased,  but  there  was 
an  equal  increase  in  what  we  might  call  thermal  energy,  or  heat  energy, 
so  that  the  total  energy,  if  properly  defined,  remained  constant.  This 
generalization  was  what  really  established  the  conservation  of  energy  as  a 
great  and  important  principle.  Having  identified  heat  as  a  form  of 
energy,  it  was  only  natural  for  the  dynamical  theory  of  heat  to  be 
developed,  in  which  heat  was  regarded  as  a  mode  of  motion  of  the  mole- 
cules, on  such  a  small  scale  that  it  could  not  be  observed  in  an  ordinary 
mechanical  way.  The  extra  kinetic  and  potential  energy  of  the  molcules 
on  account  of  this  thermal  motion  was  identified  with  the  energy  which 
had  disappeared  from  view,  but  had  reappeared  to  be  measured  as  heat. 
With  the  development  of  thermodynamics  and  kinetic  theory,  conser- 
vation of  energy  took  its  place  as  the  leading  principle  of  physics,  which 
it  has  held  ever  since. 

2.  Internal  Energy,  External  Work,  and  Heat  Flow. — We  have  seen 
that  the  theory  of  heat  is  based  on  the  idea  of  conservation  of  energy, 
on  the  assumption  that  the;  total  energy  of  the  universe  is  conserved,  if 
we  include  not  only  mechanical  energy  but  also  the  mechanical  equivalent 
of  the  heat  energy.  JLt  is  not  very  convenient  to  talk  about  the  whole 
universe  every  time  we  wish  to  work  a  problem,  however.  Ordinarily, 
thermodynamics  deals  with  a  finite  system,  isolated  from  its  neighbors 
by  an  imaginary  closed  surface.  Everything  within  the  surface  belongs 
to  the  system,  everything  outside  is  excluded.  Usually,  though  not 
always,  a  fixed  amount  of  matter  belongs  to  the  system  during  the 
thennodynamic  processes  we  consider,  no  matter  crossing  the  boundary. 
Very  often,  however,  we  assume  that  energy,  in  the  form  of  mechanical 
or  thermal  energy,  or  in  some  other  form,  crosses  the  boundary,  so  that 
the  energy  of  the  system  changes.  The  principle  of  conservation,  which 
then  becomes  equivalent  to  the  first  law  of  thermodynamics,  simply 
states  that  the  net  increase  of  energy  in  the  system,  in  any  process, 
equals  the  energy  which  has  flowed  in  over  the  boundary,  so  that  no 
energy  is  created  within  the  system.  To  make  this  a  precise  law,  we 
must  consider  the  energy  of  the  body  and  its  change  on  account  of  flow 
over  the  boundary  of  the  system. 

The  total  energy  of  all  sorts  contained  within  the  boundary  of  the 
system  is  called  the  internal  energy  of  the  system,  and  is  denoted  by  U. 
From  an  atomic  point  of  view,  the  internal  energy  consists  of  kinetic 
and  potential  energies  of  all  the  atoms  of  the  system,  or  carrying  it 


SEC.  2]  HEAT  AS  A  MODE  OF  MOTION  7 

further,  of  all  electrons  and  nuclei  constituting  the  system.  Since 
potential  energies  always  contain  arbitrary  additive  constants,  the 
internal  energy  U  is  not  determined  in  absolute  value,  only  differences  of 
internal  energy  having  a  significance,  unless  some  convention  is  made 
about  the  state  of  zero  internal  energy.  Macroscopically  (that  is,  viewing 
the  atomic  processes  on  a  large  scale,  so  that  we  cannot  see  what  indi- 
vidual  atoms  are  doing),  we  do  not  know  the  kinetic  and  potential  energies 
of  the  atoms,  and  we  can  only  find  tlm  p.lm.ny>  nf  internal  nnnrgy  by 
the  amounts  of  ClierV  added  to  tr>  {^stpm  n.nrn«K  t.hn 


and  by  making  use  of  the  law  of  conservation  of  energy.  Thermo- 
dynamics, which  is  a  macroscopic  science^  makes  no  attempt  to  analyze? 
internal  energy  into  its  parts,  as  for  Dimple  mechanical  energy  and  heat, 
onotgy.  It  simply  deals  with  t.hn  totnl  iiit.p.riml  onprgy  n.nH  wifji  i|[ti 
change^. 

Energy  can  (Miter  the  system  in  many  ways,  but  most  methods  can  bo 
classified  easily  and  in  an  obvious  way  into  mechanical  work  and  heat. 
Familiar  examples  of  external  mechanical  work  are  work  done  by  pistons, 
shafts,  belts  and  pulleys,  etc.,  and  work  done  by  external  forces  acting  at  a 
distance,  as  gravitational  work  dono  on  bodies  within  the  system  on 
account  of  gravitational  attraction  by  external  bodies.  A  familiar 
example  of  heat  flow  is  heat  conduction  across  the  surface.  Convection 
of  hrat  into  the  system  is  a  possible  form  of  energy  interchange  if  atoms 
and  molecules  are  allowed  to  cross  the  surface,  but  not  otherwise.  Elec- 
tric and  magnetic  work  done  by  forces  between  bodies  within  the  system 
and  bodies  outside  is  classified  as  external  work;  but  if  the  electro- 
magnetic energy  enters  in  the  form  of  radiation  from  a  hot  body,  it  is 
classified  as  heat.  There  are  casos  whore  the  distinction  between  the 
two  forms  of  transfer  of  energy  is  not  clear  and  obvious,  and  electro- 
magnetic radiation  is  one  of  them.  In  ambiguous  cases,  a  definite 
classification  can  be  obtained  from  the  atomic  point  of  view,  by  means  of 
statistical  mechanics. 

In  an  infinitesimal  change  of  the  system^  the  energy  which  has_ 
entered  the  system  as  heat  flow  is  called  dQ^  and  the  energy  which  has 
left  the  system  as  mechanical  work  is  called  dW  (so  that  the  energy  which 
has  entered  as  mechanical  work  is  called  —  rfTF).  The  reason  for  choosing 
this  sign  for  dw  is  simply  convention;  thermodynamics  is  very  often 
used  in  the  theory  of  heat  engines,  which  produce  work,  so  that  the 
important  case  is  that  in  which  energy  leaves  the  system  as  mechanical 
work,  or  when  dW  in  our  definition  is  positive.  We  see  then  that  the1 
total  energy  which  enters  the  system  in  an  infinitesimal  change  is 
dQ  —  dW.  By  the  law  of  conservation  of  energy,  the  increase  in  internal 
energy  in  a  process  equals  the  energy  which  has  entered  the  system: 

dU  =  dQ  -  dW.  _  (2.1) 


8  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  I 

Equation  (2.1)  is  the  mathematical  statement  of  the  first  law  of  thermo- 
dynamics. It  is  to  be  noted  that  both  sides  of  the  equation  should  be 
expressed  in  the  same  units.  Thus  if  internal  energy  and  mechanical 
work  are  expressed  in  ergs,  the  heat  absorbed  must  be  converted  to  ergs 
by  use  of  the  mechanical  equivalent  of  heat, 

1  calorie  =  4.185  X  107  ergs  =  4.185  joules. 

Or  if  the  hoat  absorbed  is  to  bo  measured  in  calories,  the  work  and 
internal  energy  should  be  converted  into  that  unit. 

It  is  of  tho  utmost  importance  to  realize  that  the  distinction  between 
heat  flow  n.nrf  mop.tin.ninn.1  work,  which  WQ  have  madn  in  talking  about 
pnnrgry  in  fmn^f  infr)  n  system,  does  not  apply  to  the  onergy  once  it  is  in 
the  system.  It  is  completely  fallacious  to  try  to  break  down  tho  stato- 
mont  of  Kq.  (2.1)  into  two  statements:  "Tho  iiicroaso  of  hoat  onorgy  of  a 
body  oquals  tho  hoat  which  has  fiowod  in,"  and  "Tho  doorcase  of  mechan- 
ical onorgy  of  a  body  oquals  tho  work  clono  by  tho  body  on  its  surround- 
ings. "  For  theso  statements  would  correspond  just  to  separate 
conservation  laws  for  hoat  and  moohanioal  onorgy,  and  we  havo  soon  in 
the  last  section  that  such  separate  laws  do  not  exist.  To  return  to  tho 
last  section,  Rumford  put  a  groat  deal  of  mechanical  work  into  his 
cannon,  produced  no  moohanioal  results  on  it,  but  succeeded  in  raising 
its  temperature  greatly.  As  we  havo  statod  before,  the  energy  of  a 
system  cannot  be  differentiated  or  separated  into  a  mechanical  and  a 
thermal  part,  by  any  method  of  thermodynamics.  Tho  distinction 
between  heat  and  work  is  made  in  discussing  energy  in  transit,  and 
only  there. 

The  internal  energy  of  a  system  depends  only  on  tho  state  of  the 
system;  that  is,  on  pressure,  volume,  temperature,  or  whatever  variables 
arc  used  to  describe  tho  system  uniquely.  Thus,  the  change1  in  internal 
energy  between  two  states  1  and  2  depends  only  on  these  states.  This 
change  of  internal  energy  is  an  integral, 

f/»  -  I/i  =  f*dU  =  f'dQ  -  f*dW.  (2.2) 

Since  this  integral  depends  only  on  the  end  points,  it  is  independent  of  the 
path  used  in  going  from  state  1  to  state  2.  But  the  separate  integrals 


J>        and        £dW, 


representing  the  total  heat  absorbed  and  the  total  work  done  in  going 
from  state  1  to  2,  are  not  independent  of  the  path,  but  may  be  entirely 
different  for  different  processes,  only  their  difference  being  independent 
of  path.  Since  these  integrals  are  not  independent  of  the  path,  they 
cannot  be  written  as  differences  of  functions  Q  and  W  at  the  end  points, 
as  /  dU  can  be  written  as  the  difference  of  the  C77s  at  the  end  points. 


SEC.  3]  HEAT  Ati  A  MODE  OF  MOT  JON  9 

Such  functions  Q  and  W  do  not  exist  in  any  unique  way,  and  we  are  not 
allowed  to  use  them.  W  would  correspond  essentially  to  the  negative  of 
the  potential  energy,  but  ordinarily  a  potential  energy  function  does  no1 
exist.  Similarly  Q  would  correspond  to  tho  amount  of  hoat  in  the  body, 
but  we  have  seen  that  this  function  also  dors  not  exist.  The  fact  that 
functions  Q  and  W  do  not  exist^  or  that  f  dQ  and  I  dW  are  not  independent 
of  patht  really  is  only  another  way  of  saying  that  mechanical  and  heat 
energy  are  interchangeable,  and  that  the  internal  energy  cannot  T? 
dividftfl  into  n.  mnnhanical  and  a  thermal  part  by  thermodynamics. 

At  first  sight,  it  seems  too  bad  that  /  dQ  is  not  independent  of  path, 
for  some  such  quantity  would  be  useful.  It  would  be  pleasant  to  be  able 
to  say,  in  a  given  state  of  the  system,  that  the  system  had  so  and  so  much 
heat  energy.  Starting  from  the  absolute  zero  of  temperature,  where  we 
could  say  that  the  heat  energy  was  zero,  we  could  heat  tho  body  up  to  the 
stato  we  wore  interested  in,  find  J  dQ  from  absolute  zero  up  to  this  state, 
and  call  that  the  heat  energy.  But  the  stubborn  fact  remains  that 
we  should  get  different  answers  if  we  heated  it  up  in  different  ways.  For 
instance,  we  might  heat  it  at  an  arbitrary  constant  pressure  until  we 
reached  the  desired  temperature*,  then  adjust  the  pressure  at  constant 
temperature  to  the  desired  value;  or  we  might  raise  it  first  to  the  desired 
pressure,  then  heat  it  at  that  pressure  to  the  final  temperature;  or  many 
other  equally  simple  processes.  Each  would  give  a  different  answer, 
as  we  can  easily  verify.  There  is  nothing  to  do  about  it. 

It  is  to  avoid  this  difficulty,  and  obtain  something  resembling  the 
"  amount  of  heat  in  a  body/'  which  yet  has  a  unique  meaning,  that  we 
introduce  the  entropy.  If  T  is  the  absolute  temperature,  and  if  the  heat 
dQ  is  absorbed  at  temperature  T  in  a  reversible  way,  then  /  dQ/T  proves 
tcTbe  an  integral  independent  of  path,  which  evidently  increases  as  the 
body  is  heated:  that  is.  as  heat  flows  into  it.  This  integral,  from  a  fixed 
zero  point  (usually  taken  to  be  the  absolute  zero  of  temperature),  is 
called  the  entropy.  Like  the  internal  energy,  it  is  determined  by  the 
state  of  the  system,  but  unlike  the  internal  energy  it  measures  in  a 
certain  way  only  heat  energy,  not  mechanical  energy.  We  next  take  up 
the  study  of  entropy,  and  of  the  related  second  law  of  thermodynamics. 

3.  The  Entropy  and  Irreversible  Processes. — Unlike  the  internal 
energy  and  the  first  law  of  thermodynamics,  the  entropy  and  the  second 
law  are  relatively  unfamiliar.  Like  them,  however,  their  best  inter- 
pretation comes  from  the  atomic  point  of  view,  as  carried  out  in  statistical 
mechanics.  For  this  reason,  we  shall  start  with  a  qualitative  description 
of  the  nature  of  the  entropy,  rather  than  with  quantitative  definitions 
and  methods  of  measurement. 

The  entropy  is  a  quantity  characteristic  of  thft  state  of  ft  system. 
Tnpfl.siiring  the  randomness  or  disorder  in  the  atomic  arrangement  of  that 


10  INTRODUCTION  TO  CHEMICAL  1'UYMCti  [CHAP.  1 

state.  It  increases  when  a  body  is  heated,  for  then  the  random  atomic 
motion  increases.  But  it  ftlso  innrnn^fia  wfrpn  a  regular,  orderly  motion  is 
f  npvnrfnH  info  n  fftpdom  motion.  Thus,  consider  an  enclosure  containing 
a  small  piece  of  crystalline  solid  at  the  absolute  zero  of  temperature,  in 
a  vacuum.  The  atoms  of  the  crystal  are  regularly  arranged  and  at  rest; 
its  entropy  is  zero.  Heat  the  crystal  until  it  vaporizes.  The  molecules 
are  now  located  in  random  positions  throughout  the  enclosure  and  have 
velocities  distributed  at  random.  Both  types  of  disorder,  in  the  coordi- 
nates and  in  the  velocities,  contribute  to  the  entropy,  which  is  now 
large.  But  we  could  have  reached  the  same  final  .state  in  a  different  way, 
not  involving  the  absorption  of  heat  by  the  system.  We  could  have 
accelerated  the  crystal  at  the  absolute  zero,  treating  it  as  a  projectile  and 
doing  mechanical  work,  but  without  heat  How.  We  could  arrange  a 
target,  so  that  the  projectile  would  automatically  strike  the  target, 
without  external  action.  If  the  mechanical  work  which  we  did  on  the 
system  were  equivalent  to  the  heat  absorbed  in  the  other  experiment, 
the  final  internal  energy  would  be  the  same  in  each  case.  In  our  second 
experiment,  then,  when  the  projectile  struck  the  target  it  would  be  heated 
so  hot  as  to  vaporize,  filling  the  enclosure  with  vapor,  and  the  final  state 
would  be  just  the  same  as  if  the  vaporization  were  produced  directly. 
The  increase  of  entropy  must  then  be  the  same,  for  by  hypothesis  the 
entropy  depends  only  on  the  state  of  the  system,  not  on  the  path  by 
which  it  has  reached  that  state.  In  the  second  case,  though  the  entropy 
has  increased,  no  heat  has  been  absorbed.  Rather,  ordered  mechanical 
energy  (the  kinetic  energy  of  the  projectile  as  a  whole,  in  which  each 
molecule  was  traveling  at  the  same  velocity  as  every  other)  has  been 
converted  by  the  collision  into  random,  disordered  energy.  Just  this 
change  results  in  an  increase  of  entropy.  It  is  plain  that  entropy  cannot 
be  conserved,  in  the  same  sense  that  matter,  energy,  and  momentum  are. 
For  here  entropy  has  been  produced  or  created,  just  by  a  process  of 
changing  ordered  motion  into  disorder. 

Many  other  examples  of  the  two  ways  of  changing  entropy  could 
be  given,  but  the  one  we  have  mentioned  illustrates  them  sufficiently. 
We  have  considered  the  increase  of  entropy  of  the  system;  let  us  now  ask 
if  the  processes  can  be  reversed,  and  if  the  entropy  can  be  decreased  again. 
Consider  the  first  process,  where  the  solid  was  heated  gradually.  Let  us 
be  more  precise,  and  assume  that  it  was  heated  by  conduction  from  a  hot 
body  outside;  and  further  that  the  hot  body  was  of  an  adjustable  temper- 
ature, and  was  always  kept  very  nearly  at  the  same  temperature  as  the 
system  we  were  interested  in.  If  it  were  just  at  the  same  temperature, 
heat  would  not  flow,  but  if  it  were  always  kept  a  small  fraction  of  a  degree 
warmer,  heat  would  flow  from  it  into  the  system.  But  that  process  can 
be  effectively  reversed.  Instead  of  having  the  outside  body  a  fraction 


SBC.  3]  HEAT  AS  A  MODE  OF  MOTION  11 

of  a  degree  warmer  than  the  system,  we  )et  it  be  a  fraction  of  a  degree 
cooler,  so  that  heat  will  flow  out  instead  of  in.  Then  things  will  cool 
down,  until  finally  the  system  will  return  to  the  absolute  zero,  and  every- 
thing will  be  as  before.  In  the  direct  process  heat  flows  into  the  system; 
in  the  inverse  process  it  flows  out,  an  equal  amount  is  returned,  and  when 
everything  is  finished  all  parts  of  the  system  and  the  exterior  are  in 
essentially  the  same  state  they  were  at  the  beginning.  But  now  try  to 
reverse  the  second  process,  in  which  the  solid  at  absolute  zero  was 
accelerated,  by  means  of  external  work,  then  collided  with  a  target,  and 
vaporized.  The  last  steps  were  taken  without  external  action.  To 
reverse  it,  we  should  have  the  molecules  of  the  vapor  condense  to  form  a 
projectile,  all  their  energy  going  into  ordered  kinetic  energy.  It  would 
have  to  be  as  shown  in  a  motion  picture  of  the  collision  run  backward, 
all  the  fragments  coalescing  into  an  unbroken  bullet.  Then  we  could 
apply  a  mechanical  brake  to  the  projectile  as  it  receded  from  the  target, 
and  get  our  mechanical  energy  out  again,  with  reversal  of  the  process. 
But  such  things  do  not  happen  in  nature.  The  collision  of  a  projectile 
with  a  target  is  essentially  an  irreversible  process,  which  never  happens 
backward,  and  a  reversed  motion  picture  of  such  an  event  is  inherently 
ridiculous  and  impossible.  The  statement  that  such  events  cannot  be 
reversed  is  one  of  the  essential  parts  of  the  second  law  of  thermodynamics. 
If  we  look  at  the  process  from  an  atomic  point  of  viow,  it  is  clear  why  it 
cannot  reverse.  The  change  from  ordered  to  disordered  motion  is  an 
inherently  likely  change,  which  can  be  brought  about  in  countless  ways; 
whereas  the  change  from  disorder  to  order  is  inherently  very  unlikely, 
almost  sure  not  to  happen  by  chance.  Consider  a  jigsaw  puzzle,  which 
can  be  put  together  correctly  in  only  one  way.  If  we  start  with  it  put 
together,  then  remove  each  picco  and  put  it  in  a  different  place  on  tho 
table,  we  shall  certainly  disarrange  it,  and  we  can  do  it  in  almost  countless 
ways;  while  if  we  start  with  it  taken  apart,  and  remove  each  pioce  and 
put  it  in  a  different  place  on  the  table,  it  is  true  that  we  may  happen  to 
put  it  together  in  the  process,  but  the  chances  arc  enormously  against  it. 
The  real  essence  of  irreversibility,  however,  is  not  merely  the  strong 
probability  against  the  occurrence  of  a  process.  It  is  something  deeper, 
coming  from  the  principle  of  uncertainty.  This  principle,  as  we  shall 
see  later,  puts  a  limit  on  the  accuracy  with  which  we  can  regulate  or 
prescribe  the  coordinates  and  velocities  of  a  system.  It  states  that  any 
attempt  to  regulate  them  with  more  than  a  certain  amount  of  precision 
defeats  its  own  purpose :  it  automatically  introduces  unpredictable  pertur- 
bations which  disturb  the  system,  and  prevent  the  coordinates  and 
velocities  from  taking  on  the  values  we  desire,  forcing  them  to  deviate 
from  these  values  in  an  unpredictable  way.  But  this  just  prevents  us 
from  being  able  experimentally  to  reverse  a  system,  once  the  randomness 


12  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  1 

has  reached  the  small  scale  at  which  the  principle  of  uncertainty  operates, 
To  make  a  complicated  process  like  a  collision  reverse,  the  molecules 
would  have  to  be  given  very  definitely  determined  positions  and  velocities 
so  that  they  would  just  cooperate  in  such  a  way  as  to  coalesce  and  become 
unbroken  again;  any  errors  in  determining  these  conditions  would  spoi 
the  whole  thing.  But  we  cannot  avoid  these  errors.  It  is  true  that  b> 
chance  they  may  happen  to  fall  into  lino,  though  the  chance  is  minute 
But  the  important  point  is  that  we  cannot  do  anything  about  it. 

From  the  preceding  examples,  it  is  clear  that  we  must  consider  twr 
types  of  processes:  reversible  and  irreversible.  Tjie  essential  feature  oi 
reversible  processes  is  that  things  are  almost  balanced,  almost  in  equilib- 
rium, at  every  stage,  so  that  an  infinitesimal  change  will  swing  the  motioi 
from  one  direction  to  the  other.  Tjrflveraihle  processes,  on  the  othei 
hand,  involve  complete  departure  from  equilibrium,  as  in  a  collision.  II 
will  be  worth  while  to  enumerate  a  few  other  common  examples  oi 
irreversible  processes.  Heat  flow  from  a  hot  body  to  a  cold  body  at  mon 
than  an  infinitesimal  difference  of  temperature  is  irreversible,  for  the  heal 
never  flows  from  the  cold  to  the  hot  body.  Another  example  is  viscosity 
in  which  regular  motion  of  a  fluid  is  converted  into  random  moleculai 
motion,  or  heat.  Still  another  is  diffusion,  in  which  originally  unmixec 
substances  mix  with  other,  so  that  they  cannot  bo  unmixed  again  withoul 
external  action.  In  all  these  cases,  it  is  possible  of  course  to  bring  tin 
system  itself  back  to  its  original  state.  Even  the  projectile  which  ha* 
been  vaporized  can  be  reconstructed,  by  cooling  and  condensing  the 
vapor  and  by  recasting  the  material  into  a  new  projectile.  But  th< 
surroundings  of  the  system  would  have  undergone  a  permanent  change 
the  energy  that  was  originally  given  the  system  as  mechanical  energy,  tc 
accelerate  the  bullet,  is  taken  out  again  as  heat,  in  cooling  the  vapor,  sc 
that  the  net  result  is  a  conversion  of  mechanical  energy  into  heat  in  the 
surroundings  of  the  system.  Such  a  conversion  of  mechanical  energy 
into  heat  is  often  called  degradation  of  energy,  and  it  is  characteristic  oi 
irreversible  processes.  A  reversible  process  is  one  which  can  be  reversed 
iu  such  a  way  that  the  system  itself  and  its  surroundings  both  return  tc 
their  original  condition;  while  an  irreversible  process  is  one  such  that  th( 
system  cannot  be  brought  back  to  its  original  condition  without  requir- 
ing a  conversion  or  degradation  of  some  external  mechanical  energy  intt 
heat. 

4.  The  Second  Law  of  Thermodynamics.  —  We  are  now  ready  to  give 
a  statement  of  the  second  law  of  ^hermnflyria^nica,  in  n^p  of  its  many 
forms:  The  entropy.  a  function  only  of  the  state  of  fl.  system,  increases  in  a 
reversible 


absorbed,   T  the  absolute  temperature  at  which  it  is  absorbed)  and 
increases  by  a  larger  amount  than  dQ/T  in  an  irreversible  process. 


SEC.  4]  HEAT  AS  A  MODE  OF  MOTION  13 

This  statement  involves  a  number  of  features.  First,  it  gives  a  way 
of  calculating  entropy.  By  sufficient  ingenuity,  it  is  always  possible  to 
find  reversible  ways  of  getting  from  any  initial  to  any  final  state,  pro- 
vided both  arc  equilibrium  states.  Then  we  can  calculate  /  dQ/T  for 
such  a  reversible  path,  and  the  result  will  bo  the  change  of  entropy 
between  the  two  states,  an  integral  independent  of  path.  We  can  then 
measure  entropy  in  a  unique  \vay.  If  we  now  go  from  the  same  initial  to 
the  same  final  stale  by  an  irreversible  path,  the  change  of  entropy  must, 
still  be  the  same,  though  now  /  dQ/T  must  necessarily  be  smaller  than 
before,  and  hence  smaller  than  the  change  in  entropy.  We  see  that, 
the  heat  absorbed  in  an  irreversible  path  must  be  less  than  in  a  reversible 
path  between  the  same  end  points.  Since  ^hfc  change  in  internal  energy 
must  be  the  same  in  either  ease,  the  first  law  then  tells  us  that  the  external 
work  done  by  the  system  is  less  for  the  irreversible  path  than  for  the 
reversible  one.  If  our  system  is  a  heat  engine,  whose  object  is  to  absorb 
heat  and  do  mechanical  work,  \\e  see  that  the  mechanical  work  accom- 
plished will  be  less  for  an  irreversible  engine  than  for  a  reversible  one, 
operating  between  the  same  end  points. 

It  is  interesting  to  consider  the  limiting  case  of  adiabatic  processes, 
processes  in  which  the  system  interchanges  no  heat  witkihc  surroundings, 
the  only  changes  in  internal  energy  coming  from  mechanical  work.  We 
see  that  in  a. reversible  adiabatic  process  the  entropy  does  not  change 
(a  convenient  way  of  describing  such  processes).  In  an  irreversibly 
fljin.hn.fip  prr>n^^  foe  entropy  increases.  In  particular,  for  a  system 
entirely  isolated  from  its  surroundings,  the;  entropy  increases  whenever 
irreversible  processes  occur  within  it.  An  isolated  system  in  which 
irreversible  processes  can  occur  is  surely  not  in  a  steady,  equilibrium  state; 
the  various  examples  which  we  have  considered  are  the  rapidly  moving 
projectile,  a  body  with  different  temperatures  at  different  parts  (to  allow 
heat  conduction),  a  fluid  with  mass  motion  (to  allow  viscous  friction), 
a  body  containing  two  different  materials  not  separated  by  an  impervious 
wall  (to  allow  diffusion).  All  these  systems  have  less  entropy  than  the 
state  of  thermal  equilibrium  corresponding  to  the  same  internal  energy, 
which  can  be  reached  from  the  original  state  by  irreversible  processes 
without  interaction  with  the  outside.  This  state  of  thermal  equilibrium 
is  one  in  which  the  temperature  is  everywhere  constant,  there  is  no  mass 
motion,  and  where  substances  are  mixed  in  such  a  way  that  there  is  no 
tendency  to  diffusion  or  flow  of  any  sort.  A  condition  for  thermal 
equilibrium,  which  is  often  applied  in  statistical  mechanics,  is  that  the 
equilibrium  state  is  that  of  highest  entropy  consistent  with  the  given 
internal  energy  and  volume. 

These  statements  concerning  adiabatic  changes,  in  which  the  entropy 
can  only  increase,  should  not  cause  one  to  forget  that  in  ordinary  changes, 


14  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  I 

in  which  heat  can  be  absorbed  or  rejected  by  the  system,  the  entropy 
can  either  increase  or  decrease.  In  most  thermodynamic  problems,  wre 
confine  ourselves  to  reversible  changes,  in  which  the  only  way  for  the 
entropy  to  change  is  by  heat  transfer. 

We  shall  now  state  the  second  law  in  a  mathematical  form  which  is 
very  commonly  used.  We  let  S  denote  the  entropy.  Our  previous 
statement  is  then  dS  ^  dQ/T,  or  T  dS  ^  dQ,  the  equality  sign  holding 
for  the  reversible,  the  inequality  for  irreversible,  processes.  But  now  we 
use  the  first  law,  Eq.  (2.1),  to  express  dQ  in  terms  of  dU  and  dW.  The 
inequality  becomes  at  once 

J^dS  ^  dU  +  dW,    '  (4.1) 


the  mathematical  formulation  of  the  second  law.  For  reversible .  pror- 
cssesf  which  we  ordinarily  consider,  the  ecmalitv  sign  is  to  be  used. 

The  second  law  may  be  considered  as  a  postulate.  We  shall  see  in 
Jhap.  II  that  definite  consequences  can  be  drawn  from  it,  and  they 
prove  to  be  always  in  agreement  with  experiment.  We  notice  that 
in  stating  it,  we  have  introduced  the  temperature  without  apology,  for 
the  first  time.  This  again  can  be  justified  by  its  consequences:  the 
temperature  so  defined  proves  to  agree  with  the  temperature  of  ordinary 
experience,  as  defined  for  example  by  the  gas  thermometer.  Thermo- 
dynamics is  the  science  that  simply  starts  by  assuming  the  first  and 
second  laws,  and  deriving  mathematical  results  from  them.  Both  laws 
are  simple  and  general,  applying  as  far  as  \vc  know  to  all  sorts  of  processes. 
As  a  result,  we  can  derive  simple,  general,  and  fundamental  results  from 
thermodynamics,  which  should  be  independent  of  any  particular  assump- 
tions about  atomic  and  molecular  structure,  or  such  things.  Thermo- 
dynamics has  its  drawbacks,  however,  in  spite  of  its  simplicity  and 
generality.  In  the  first  place,  there  are  many  problems  which  it  simply 
cannot  answer.  These  are  detailed  problems  relating,  for  insTanC^to 
the  equation  of  state  and  specjfi^  hfigf  nf  r^t.i/nilar  types  of  substances. 
Thfirnqftdynprnifta  must  assume JJiat_thp«p  qnanfjfog  «TP  determined  by 
experiment*  once  they  are  known,  it  can  predict  certain  relationships 
between  observed  quantities,  but  it  is  unable  to  say  what  values  the 
quantities  must  have.  IrT  addition  to  this,  thermodynamics  is  limited 
to  the  discussion  of  problems  in  equilibrium.  This  is  on  account  of  the 
form  of  the  second  law,  which  can  give  only  qualitative,  and  not  quanti- 
tative, information  about  processes  out  of  equilibrium. 

Statistical  mechanics  is  a  much  more  detailed  science  than  thermo- 
dynamics, and  for  that  reason  is  in  some  ways  more  complicated.  It 
undertakes  to  answer  the  questions,  how  is  each  atom  or  molecule  of  the 
substance  moving,  on  the  average,  and  how  do  these  motions  lead  to 
observable  large  scale  phenomena?  For  instance,  how  do  the  motions 


SEC.  4]  HEAT  AS  A  MODE  OF  MOTION  15 

of  the  molecules  of  a  gas  lead  to  collisions  with  a  wall  which  we  interpret 
as  pressure?  Fortunately  it  is  possible  to  derive  some  very  beautiful 
general  theorems  from  statistical  mechanics.  In  fact,  one  can  give 
proofs  of  the  first  and  second  laws  of  thermodynamics,  as  direct  conse- 
quences of  the  principles  of  statistical  mechanics,  so  that  all  the  results 
of  thermodynamics  can  be  considered  to  follow  from  its  methods.  But 
it  can  go  much  further.  It  can  start  with  detailed  models  of  matter 
and  work  through  from  them  to  predict  the  results  of  large  scale  experi- 
ments on  the  matter.  Statistical  mechanics  thus  is  much  more  powerful 
than  thermodynamics,  and  it  is  essentially  just  as  general.  It  is  some- 
what more  complicated,  however,  and  somewhat  more  dependent  on  the 
exact  model  of  the  structure  of  the  material  which  we  use.  Like  thermo- 
dynamics, it  is  limited  to  treating  problems  in  equilibrium. 

Kinetic  theory 'is  a  study  of  the  rates  of  atomic  and  molecular  proc- 
esses, treated  by  fairly  direct  methods,  without  much  benefit  of  general 
principles.  If  handled  properly,  it  is  an  enormously  complicated  subject, 
though  simple1  approximations  can  be  made  in  particular  cases.  It  is 
superior  to  statistical  mechanics  and  thermodynamics  in  just  two  respects. 
In  the  first  place,  it  makes  use  only  of  well-known  and  elementary 
methods,  arid  for  that  reason  is  somewhat  more  comprehensible  at  first 
sight  than  statistical  mechanics,  with  its  more  advanced  laws.  In  the 
second  place,  it  can  handle  problems  out  of  equilibrium,  such  as  the  rates 
of  chemical  reactions  and  other  processes,  which  cannot  be  treated  by 
thermodynamics  or  statistical  mechanics. 

We  see  that  eadi  of  our  three  sciences  of  heat  has  its  own  advantages. 
A  properly  trained  physicist  or  chemist  should  know  all  three,  to  be  able 
to  use  whichever  is  most  suitable  in  a  given  situation.  We  start  witli 
thermodynamics,  since  it  is  the  most  general  and  fundamental  method, 
taking  up  thermodynarnic  calculations  in  the  next  chapter.  Following 
that  we  treat  statistical  mechanics,  and  still  later  kinetic  theory.  Only 
then  shall  we  be  prepared  to  make  a  real  study  of  the  nature  of  matter. 


CHAPTER  II 
THERMODYNAMICS 

In  the  last  chapter,  we  became  acquainted  with  the  two  laws  of 
thermodynamics,  but  we  have  not  seen  how  to  use  them.  In  this 
chapter,  we  shall  learn  the  rules  of  operation  of  thermodynamics,  though 
we  shall  postpone  actual  applications  until  later.  It  has  already  been 
mentioned  that  thermodynamics  can  give  only  qualitative  information 
for  irreversible  processes.  Thus,  for  instance,  the  second  law  may  be 
stated 

dW  Z  TdS  -  dU,  (1) 

giving  an  upper  limit  to  the  work  done  in  an  irreversible  process,  but  not 
predicting  its  exact  amount.  Only  for  reversible  processes,  where  the 
equality  sign  may  bo  used,  can  thermodynamics  make  definite  predictions 
of  a  quantitative  sort.  Consequently  almost  all  our  work  in  this  chapter 
will  deal  with  reversible  systems.  We  shall  find  a  number  of  differential 
expressions  similar  to  Eq.  (1),  and  by  proper  treatment  we  can  convert 
these  into  equations  relating  one  or  more  partial  derivatives  of  one 
thcrmodynamic  variable  with  respect  to  another.  Such  equations, 
called  thermodynamic  formulas,  often  relate  different  quantities  all  of 
which  can  be  experimentally  measured,  and  hence  furnish  a  check  on  the 
accuracy  of  the  experiment.  In  cases  where1  one  of  the  quantities  is 
difficult  to  measure,  they  can  be  used  to  compute  one  of  the  quantities 
from  the  others,  avoiding  the  necessity  of  making  the  experiment  at  all. 
There  are  a  very  great  many  thermodynamic  formulas,  and  it  would  be 
hopeless  to  find  all  of  them.  But  we  shall  go  into  general  methods  of 
computing  them,  and  shall  set  up  a  convenient  scheme  for  obtaining  any 
one  which  we  may  wish,  with  a  minimum  of  computation. 

Before  starting  the  calculating  of  the  formulas,  we  shall  introduce 
several  new  variables,  combinations  of  other  quantities  which  prove  to  be 
useful  for  one  reason  or  another.  As  a  matter  of  fact,  we  shall  work  with 
quite  a  number  of  variables,  some  of  which  can  be  taken  to  be  inde- 
pendent, others  dependent,  and  it  is  necessary  to  recognize  at  the  outset 
the  nature  of  the  relations  between  them.  In  the  next  section  we  consider 
the  equation  of  state,  the  empirical  relation  connecting  certain  thermo- 
dynamic variables. 

1.  The  Equation  of  State. — In  considering  the  properties  of  matter, 
our  system  is  ordinarily  a  piece  of  material  enclosed  in  a  container  and 

16 


SEC.  1]  THERMODYNAMICS  17 

subject  to  a  certain  hydrostatic  pressure.  This  of  course  is  a  limited  type 
of  system,  for  it  is  not  unusual  to  have  other  types  of  stresses  acting,  such 
as  shearing  stresses,  unilateral  tensions,  and  so  on.  Thermodynamics 
applies  to  as  general  a  system  as  we  please,  but  for  simplicity  we  shall  limit 
our  treatment  to  the  conventional  case  where  the  only  external  work  is 
done  by  a  change  of  volume,  acting  against  a  hydrostatic  pressure.  That 
is,  if  P  is  the  pressure  and  V  the  volume  of  the  system,  we  shall  have 

dW  =  PdV.  (1.1) 

In  any  case,  even  with  much  more  complicated  systems,  the  work  done 
will  have  an  analogous  form;  for  Eq.  (1.1)  is  simply  a  force  (P)  times  a 
displacement  (dV),  and  we  know  that  work  can  always  be  put  in  such 
a  form.  If  there  is  occasion  to  set  up  the  thermodynamic  formulas  for  a 
more  general  type  of  force  than  a  pressure,  we  simply  set  up  dW  in  a  form 
corresponding  to  P]q.  (1.1),  and  proceed  by  analogy  with  the  derivations 
which  we  shall  give  here. 

We  now  have  a  number  of  variables :  P.  V}  rl\  U .  and  AS.  How  many 
of  these T  we  mav  ask,  are  independent.?  The_  answer  is,  any  two.  For 
example,  with  a  given  system,  we  may  fix  the  pressure  and  temperature 
Then  in  general  the  volume  is  determined,  as  we  can  find  experimentally 
The  experimental  relation  giving  volume  as  a  function  of  pressure  and 
temperature  is  called  the  equation  of  state.  Ordinarily,  of  course,  it  is 
not  a  simple  analytical  equation,  though  in  special  cases  like  a  perfect  gas 
it  may  be.  Instead  of  expressing  volume  as  a  function  of  pressure  and 
temperature,  we  may  simply  say  that  the  equation  of  state  expresses  a 
relation  between  these  three  variables,  which  may  equally  well  give,  pres- 
sure as  a  function  of  temperature  and  volume,  or  temperature  as  a 
function  of  volume  and  pressure.  Of  these  three  variables,  two  are  inde- 
pendent, one  dependent,  and  it  is  immaterial  which  is  chosen  as  the 
dependent  variable. 

The  equation  of  ata.t.e.  doos  not,  include  all  the  experimental  informa- 
tion which  wo  n^ist  have  about  a  system  or  substance.  We  need  'to  know 
also  its  hftftt  pfl.pu>it.y  or  specific  heat,  as  a.  function  of  temperature.  Sup- 
pose, for  instance,  that  we  know  the  specific  heat  at  constant  pressure 
CP  as  a  function  of  temperature  at  a  particular  pressure.  Then  we  can 
find  the  difference  of  internal  energy,  or  of  entropy,  between  any  two 
states.  From  the  first  state,  we  can  go  adiabatically  to  the  pressure  at 
which  we  know  Cp.  In  this  process,  since  no  heat  is  absorbed,  the  change 
of  internal  energy  equals  the  work  done,  which  we  can  compute  from  the 
equation  of  state.  Then  we  absorb  heat  at  constant  pressure,  until  we 
reach  the  point  from  which  another  adiabatic  process  will  carry  us  to  the 
desired  end  point.  The  change  of  internal  energy  can  be  found  for  the 
process  at  constant  pressure,  since  there  we  know  CP,  from  which  we  can 


18  INTRODUCTION  TO  CHEMICAL  PHYSICS  |CiiAi>.  II 

find  the  heat  absorbed,  and  since  the  equation  of  state  will  tell  us  the 
work  done;  for  the  final  adiabatic  process  we  can  likewise  find  the  work 
done  and  hence  the  change  of  internal  energy.  Similarly  we  can  find  the 
change  in  entropy  between  initial  and  final  stale.  In  our  particular  case, 
assuming  the  process  to  be  carried  out  reversibly,  the  entropy  will  not 
change  along  the  adiabatics,  but  the  change  of  entropy  will  be 

dQ       CrdT 
T  T 

in  the  process  at  constant  pressure.  We  see,  in  other  words,  that  the 
difference  of  internal  energy  or  of  entropy  bnt,ween  fl,Tiy  twn  «t"^«  can 
be  found  if  we  know  equation  of  state  and  specific  heat,  and  since  both 
these  quantities  have  arbitrary  additive  constants,  this  is  all  the  informa- 
tion which  we  can  expect  to  obtain  about  them  anyway. 

Given  the  equation  of  state  and  specific  hoatr  we  snn  that  we  can 
obtain  all  but  two  of  the  quantities  P,  Fy  T,  U,  S,  provided  those  two  are 
known.  We  have  shown  this  if  two  of  the  three  quantities  P,  F,  T  an* 
known;  but  if  U  and  S  are  determined  by  these  quantities,  that  means 
simply  that  two  out  of  the  five  quantities  are  independent,  the  rest 
dependent.  It  is  then  possible  to  use  any  two  as  independent  variables. 
For  instance,  in  thermodynamics  it  is  not  unusual  to  use  T  and  S,  or  V 
and  >S,  as  independent  variables,  expressing  everything  else  as  functions 
of  them. 

2.  The  Elementary  Partial  Derivatives. — We  can  set  up  a  number  of 
familiar  partial  derivatives  and  thermodynamic  formulas,  from  the 
information  which  we  already  have.  We  have  five  variables,  of  which 
any  two  are  independent,  the  rest  dependent.  We  can  then  set  up  the 
partial  derivative  of  any  dependent  variable  with  respect  to  any  inde- 
pendent variable,  keeping  the  other  independent  variable  constant.  A 
notation  is  necessary  showing  in  each  case  what  are  the  two  independent 
variables.  This  is  a  need  not  ordinarily  appreciated  in  mathematical 
treatments  of  partial  differentiation,  for  there  the  independent  variables 
are  usually  determined  in  advance  and  described  in  words,  so  that  there 
is  no  ambiguity  about  them.  Thus,  a  notation,  peculiar  to  thermody- 
namics, has  been  adopted.  In  any  partial  derivative,  it  is  obvious  that 
the  quantity  being  differentiated  is  one  of  the  dependent  variables,  and 
the  quantity  with  respect  to  which  it  is  differentiated  is  one  of  the  inde- 
pendent variables.  It  is  only  necessary  to  specify  the  other  independent, 
variable,  the  one  which  is  held  constant  in  the  differentiation,  and  the 
convention  is  to  indicate  this  by  a  subscript.  Thus  (dS/dT)Py  which  is 
ordinarily  read  as  the  partial  of  S  with  respect  to  T  at  constant  P,  is  the 
derivative  of  S  in  which  pressure  and  temperature  are  independent  vari- 
ables. This  derivative  would  mean  an  entirely  different  thing  from  the 
derivative  of  S  with  respect  to  T  at  constant  V,  for  instance. 


SEC.  2|  THERMODYNAMICS  19 

There  are  a  number  of  partial  derivatives  which  have  elementary 
meanings.  Thus,  consider  the  thermal  expansion.  This  is  the  fractional 
increase  of  volume  per  unit  rise  of  temperature,  at  constant  pressure: 


Thermal  expansion  —  TJ  -r-=  1  •  (2.1) 
..                             V\dl/ 

Similarly,   the  isothermal   compressibility  is  the  fractional  decrease  of 
volume  per  unit  increase  of  pressure,  at  constant  temperature: 


Isothermal  compressibility  =  ^ilv 


T 


This  is  the  compressibility  usually  employed;  sometimes,  as  in  considering 
sound  waves,  we  require  the  adiabatic  compressibility,  the  fractional 
decrease  of  volume  per  unit  increase  of  pressure,  when  no  heat  flows  in  or 
out.  If  there  is  no  heat  flow,  the  entropy  is  unchanged,  in  a  reversible 
process,  so  that  an  adiabatic  process  is  one  at  constant  entropy.  Then 
we  have 

Adiabatic  compressibility  =  —  TT(  ~7;  1  •  (2.3) 

v  V  \ol  Js 

The  specific  heats  have  simple  formulas.  At  constant  volume,  the  heat 
absorbed  equals  the  increase  of  internal  energy,  since  no  work  is  done. 
Since  the  heat  absorbed  also  equals  the  temperature  times  the  change  of 
entropy,  for  a  reversible  process,  and  since  the  heat  capacity  at  constant 
A^olume  CV  is  the  heat  absorbed  per  unit  change  of  temperature  at  constant 
volume,  we  have  the  alternative  formulas 

V 

To  find  the  heat  capacity  at  constant  pressure  Cp,  we  first  write  the  form- 
ula for  the  first  and  second  laws,  in  the  case  we  are  working  with,  where 
the  external  work  comes  from  hydrostatic  pressure  and  where  all  processes 
are  reversible: 

dU  =  TdS  -PdV, 
or 

TdS  =  dU  +  PdV.  (2.5) 

From  the  second  form  of  Kq.  (2.5),  we  can  find  the  heat  absorbed,  or 
T  dS.  Now  Cr  is  the  heat  absorbed,  divided  by  the  change  of  tempera- 
ture, at  constant  pressure.  To  find  this,  we  divide  Eq.  (2.5)  by  dTy 
indicate  that  the  process  is  at  constant  P,  and  we  have 


20  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  11 

Here,  and  throughout  the  book,  we  shall  ordinarily  mean  by  CV  and  CV 
not  the  specific  heats  (heat  capacities  per  gram),  but  the  heat  capacities 
of  the  mass  of  material  with  which  we  are  working;  though  often,  where 
no  confusion  will  arise,  wo  shall  refer  to  them  as  the  specific  heats. 

From  the  first  and  second  laws,  Eq.  (2.5),  we  can  obtain  a  number  of 
other  formulas  immediately.  Thus,  consider  the  first  form  of  the  equa- 
tion, dU  =  T  dS  —  PdV.  From  this  we  can  at  once  keep  the  volume 
constant  (set  dV  =  0),  and  divide  by  dS,  obtaining 

'dt/\-r.  (2.7) 

Similarly,  keeping  entropy  constant,  so  that  we  have  an  adiabatic  proc- 
ess, we  have 

/*TT\ 

=  -P.  (2.8) 

But  we  could  equally  well  have  used  the  second  form  of  Eq.  (2.5),  obtain- 
ing 


(dS\    _  1 

\w)v  ~  T' 


dS\    _P 

~  r  (2-9) 


From  these  examples,  it  will  be  clear  how  formulas  involving  partial 
derivatives  can  be  found  from  differential  expressions  like  Eq.  (2.5). 

3.  The  Enthalpy,  and  Helmholtz  and  Gibbs  Free  Energies.  —  Wo 
notice  that  Eq.  (2.6)  for  the  specific  heat  at  constant  pressure  is  rather 
complicated.  We  may,  however,  rewrite  it 


\a(u+PV)\    /  ,  n 

Cp  =  [ — dT—\p'(  (3>1) 

for  I  ^TJJT-  I    =  P[  -^TJ  )  )  since  P  is  held  constant  in  the  differentiation 

L    d'     >          wA 
The  quantity  U  +  PV  comes  in  sufficiently  often  so  that  it  is  worth  giving 
it  a  symbol  and  a  name.     We  shall  call  it  the  enthalpy,  and  denote  it  by 
H.     Thus  we  have 


= 


H  =  U  +  PV^ 
dH  =  dU  +PdV  +  VdP 

=  TdS+  VdP,  (3.2) 

using  Eq.  (2.5).  From  Eq.  (3.2),  we  see  that  if  dP  =  0,  or  if  the  process 
is  taking  place  at  constant  pressure,  the  change  of  the  enthalpy  equals  the 
heat  absorbed.  This  is  the  feature  that  makes  the  enthalpy  a  useful 
quantity.  Most  actual  processes  are  carried  on  experimentally  at  con- 


SBC.  3]  THERMODYNAMICS  21 

stant  pressure,  and  if  we  have  the  enthalpy  tabulated  or  otherwise  known, 
we  can  very  easily  find  the  heat  absorbed.     We  see  at  once  that 


Cp  =  \JT)P'  (3-3) 

a  simpler  formula  than  Eq.  (2.6).  As  a  matter  of  factT  the  enthalpy  fills, 
essentially  the  role  for  processes  at  constant  pressure  which  the  internal 
e tie rgy  docs  for  processes  at  constant  volume.  Thus  the  first  form  of 
Eq.  (2.5),  dU  =  T  dS  -  P  dV,  shows  that  the  heat  absorbed  at  constant 
volume  equals  the  increase  of  internal  energy,  just  as  Eq.  (3.2)  shows  that 
the  heat  absorbed  at  constant  pressure  equals  the  increase  of  the  enthalpy. 
In  introducing  the  entropy,  in  the  last  chapter,  we  stressed  the  idea 
that  it  measured  in  some  way  the  part  of  the  energy  of  the  body  bound  up 
in  heat,  though  that  statement  could  not  be  made  without  qualification. 
The  entropy  itself t  of  course,  has  not  the  dimcnsigns  of  energy,  hut,  the 
product  TS  has.  This  quantity  TS  is  sometimes  called  the  bound  energy, 
and  in  a  somewhat  closer  wav  it  represents  the  o.no.rgy  bound  RS  VIPHJ.  ln 
any  process,  the  change  in  TS  is  given  by  T  dS  +  SdT.  If  now  the 
process  is  reversible  and  isothermal  (as  for  instance  the  absorption  of 
heat  by  a  mixture  of  liquid  and  solid  at  the  melting  point,  whore  heat  can 
be  absorbed  without  change  of  temperature,  merely  melting  more  of  the 
solid),  dT  =  0,  so  that  d(TS)  =  T  dS  =  dQ.  Thus  the  increase  of  bound 
energy  for  a  reversible  isothermal  process  really  equals  the  heat  absorbed. 
This  is  as  far  as  the  bound  energy  can  be  taken  to  represent  the  energy 
hound  as  heat;  for  a  nonisothermal  process  the  change  of  hound  energy 
np  Inngpr  equals  the  hoat  absorbcdj  and  as  we  have  seenr  no  quantity 
which  is ji  function  of  the  state  alone  can  represent  the  total  heat  absorbed 


If  the  bound  energy  TS  represents  in  a  sense  the  energy  bound  fls 
heat,  the  remaining  part  of  the  internal  energy,  U  —  TS^  should  be  in 
the  same  sense  the  mechanical  part  of  the  energy,  which  is  available  to  do 
mechanical jwork.  We  sjiall  call  this  part  of  the  energy  the  Helmholtz 
free  energy,  and  denote  it  by  A.  Let  us  consider  the  change  of  the  Helm- 
holtz free  energy  in  any  process.  We  have 

A  =  U  -  TS, 
dA  =dU-  TdS-  SdT.  (3.4) 

By  Eq.  (1)  this  is 

dA  ^  -dW  -  SdTy 
or 

-dA  2      dW  +  SdT.  (3.5) 


22  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  II 

For  a  system  at  constant  temperature,  this  tells  us  that  the  work  done 
is  less  than  or  equal  to  the  decrease  in  the  Helmholtz  free  energy.  The 
Helmholtz  free  energy  then  measures  the  maximum  ^rk  which  can  be 
done  by  foe  system  in  nfn  isothcrma]  flhfl^g0  For  a  process  at  constant 
temperature,  in  which  at  the  same  time  no  mechanical  work  is  done,  the 
right  side  of  Eq.  (3.5)  is  zero,  and  wo  see  that  in  such  a  process  the  Holm- 
holtz  free  energy  is  constant  for  a  reversible  process,  but  decreases  for  an 
irreversible  process.  The  Holmholtz  free  energy  will  decrease  until  tho 
system  roaches  an  equilibrium  state,  whon  it  will  have  reached  tho  mini- 
mum value  consistent  with  the  temperature  and  with  the  fact  that  no 
external  work  can  be  done. 

For  a  system  in   equilibrium   under  hydrostatic   pressure,    wo   may 
rewrite  Kq.  (3.5)  as 

<IA  =  -PdV  -  Nrf7\  (:«)) 

.suggesting  that  the  convenient  variables  in  which  to  express  the4  Holm- 
holtz free  energy  are  the  volume}  and  the  temperature.  In  the1  case1  of 
equilibrium,  wo  find  from  Eq.  (3.6)  tho  important  relations 


(dA\ 

Ul  =  ^ 


Tho  first  of  those1  shows  that,  at  constant  temperature,  the  Holmholtz 
free  energy  has  some  of  the  properties  of  a  potential  energy,  in  that  its 
negative  derivative}  with  respect  to  it  coordinate  (tho  volume)  givos  tho 
force  (the)  pressure).  If  A  is  known  as  a  function  of  V  and  T,  the  first 
Eq.  (3.7)  givos  a  relation  between  P,  V,  and  ?T,  or  the  equation  of  state. 
From  the  second,  we>  know  entropy  in  terms  of  temperature  and  volume, 
and  differentiating  with  respect  to  temperature  at  constant  volume,  using 
Kq.  (2.4),  we  can  find  the  specific  heat.  Thus  a  knowledge  of  tho  Holm- 
holtz free  energy  as  a  function  of  volume  and  temperature  gives  both  the 
equation  of  state  and  specific  heat,  or  complete  information  about  the 
system. 

Instead  of  using  volume  and  tofnipfirfl.tiire  as  independent  variables, 
however^  wo  more  often  wish  to  use  pressure  and  temperature.  In  this 
case,  instead  of  using.  [fa*  Hnlmhnlt?  froo  energy,  it  is  more  convenient  to 
use  tho  Gibbs  free  .energy  G}  defined  bv  the  equations 

f  /  =  //  -  TS  =  U  +  PV  -  TS  =  A  +  PV-  (3.8) 

It  will  be  seen  that  this  function  stands  in  the  same  relation  to  the* 
enthalpy  that  the  Helmholtz  free  energy  does  to  the  internal  energy.  We 
can  now  find  the  change  of  the  Gibbs  free  energy  G  in  any  process.  By 
definition,  we  have  dG  =  dH  -  T  dS  -  S  dT.  Using  Eq.  (3.2),  this  is 
dO  =  dU  +PdV  +  VdP  -  TdS-SdT,  and  by  Eq.  (1)  this  is 

dG  :£  V  dP  -  S  dT.  (3.9) 


SEC.  4]  THERMODYNAMICS  23 

For  a  system  at  constant  pressure  anc}  temperature,  we  see  that  the  Gibbs 
free  energy  is  constant  for  a  reversible  process  but  decreases  for  an  irrever- 
sible process,  reaching  a  minimum  value  consistent  with  the  pressure  and 
temperature  for  the  equilibrium  state;  just  as  for  a  system  at  constant 
volume  the  Holmholtz  free  energy  is  constant  for  a  reversible  process  but 
decreases  for  an  irreversible  process.  As  with  A,  we  can  get  the  equation 
of  state  and  specific  heat  from  the  derivatives  of  (7,  in  equilibrium.  We 
have 


the  first  of  these  giving  the  volume  as  a  function  of  pressure  and  tempera- 
ture, the  second  the  entropy  as  a  function  of  pressure  and  temperature, 
from  which  we  can  find  CP  B^mFSFrts  of  Eq.  (2.6). 

Thp  (^rilifys  jjffi^  nnnrgy  O  is  pfl.yt(jf?ularlv  iniDOrfaftUti  nn  gM^-fmntj  nf  f]rtiMM  1 

physical  processes  that  occur  at  constant  pressure  and  temperature  The 
most  important  of  these  processes^  a  change  of  phase,  as  the  melting  of  a 
solid  or  the  vaporization  of  a  liquid  If  unit  mass  of  a  substance  chan&es 
phase  reycrsiblv.tit  constant  pro^n™  *"<\  ^^"p'^+urffj  thr  tnf.n.1  (li^lm 
frecMMiergy  must  be  unchanged.  That  is.  in  equilibrium,  the  Gibbs  free 
energy  per  unit  mass  must  hr>  thp.  sn.mr>  For  both  plm.sps.  On  tlie  other 
hand,  at  a  temperature  and  pressure  which  do  not  correspond  to  equilib- 
rium between  two  phases,  the  Gibbs  free  energies  per  unit  mass  will  be 
different  for  the  two  phases.  Then  the  st.n.hlf»  plm«o  nmW  t.h<^»  rttpfo- 
t  in|]«N|  fpijaf  \\r>  ^.Vinf  wliif'h  fr^.s  thp.  low<*r  Gibbs  free  energy.  If  the  svstem 
is  actually  found  in  the  phase  of  higher  Gibbs  free  energy,  it  will  be 
unstable  and  will  irreversibly  change  to  the  other  phase.  Thus  for 
instance,  the  Gibbs  free  energies  of  liquid  and  solid  as  functions  of  the 
temperature  at  atmospheric  pressure  are  represented  by  curves  whici  i 
cross  at  the  melting  point!  Below  the  melting  point  the  solid  has  the 
lower  Gibbs  free  energy.  It  is  possible  to  have  the  liquid  boln^y  jLlif> 
melting  point:  it  is  in  the  condition  known  as  supercooling.  But  any 
Alight  disturbance  is  enough  to  produce  a  sudden  and  irreversible  solidi- 
fication, with  reduction  of  Gibbs  free  energy,  the  final  stable  state  being 
the  solid.  It  is  evident  from  these  examples  that  the  Gibbs  free  energy  is 
uf  gyeat  importance  in  discussing  physical  and  chemical  processes  Thp 
Helmholtz  free  energy  ^f>ers  nnt.  WVP  *my  annh  ^p0rf,jinnf>  Wo  tt|ml|  see 
later,  however,  that  the  methods  of  fltatfctjf»«i.l  mpnlmnW  WH  pa.rf.if»i|1«.rly 
simply  to  a  calculation  of  the  Helmholtz  free  energy,  and  its  principal 
valueTc'omeg  about  IhTKiTway.  "  "  k 

4.  Methods  of  Deriving  Thermodynamic  Formulas.  —  We  have  now 
introduced  all  the  thermodynamic  variables  that  we  shall  meet:  P,  V, 
T,  S,  U,  H,  A,  G.  The  number  of  partial  derivatives  which  can  be  formed 


24  INTRODUCTION  TO  CHEMICAL  PHY  SIC  $  [CHAP.  II 

from  these  is  8  X  7  X  6  =  336,  since  each  partial  derivative  involves  one 
dependent  and  two  independent  variables,  which  must  all  be  different.  A 
few  of  these  are  familiar  quantities,  as  we  have  seen  in  Sec.  2,  but  the  great 
majority  are  unfamiliar.  It  can  be  shown,1  however,  that  a  relation  can 
be  found  between  any  four  of  those  derivatives,  and  certain  of  the  thermo- 
dynamie  variables.  Those  rotations  are  the  thormodynarnic  formulas. 
Since  there  are  336  first  derivatives,  there  aro  336  X  335  X  334  X  333 
ways  of  picking  out  four  of  these,  so  that  tho  numbor  of  independent  rela- 
tions is  this  number  divided  by  4!,  or  521,631,180  separate  formulas.  No 
other  branch  ot  physics  is  so  rich  in  mathematical  formulas,  and  some 
systematic  method  must  be  used  to  bring  order  into  the  situation.  No 
one  can  be  expected  to  derive  any  considerable  number  of  the  formulas  or 
to  keep  them  in  mind.  There  are  four  principal  methods  of  mathematical 
procedure  used  to  derive  those  formulas,  and  in  the  present  section  wo 
shall  discuss  them.  Then  in  the  next  section  we  shall  describe  a  system- 
atic procedure  for  finding  any  particular  formula  that  we  friay  wish.  Tho 
four  mathematical  methods  of  finding  formulas  are 

1.  We  have  already  seen  that  there  are  a  number  of  differential  rela- 
tions of  the  form 

dx  =  Kdy  +Ldz,  (4.1) 

where  K  and  L  are  functions  of  the  variables.  The  most  important  rela- 
tions of  this  sort  which  we  have  met  aro  found  in  Kqs.  (2.5),  (3.2),  (3.6), 
and  (3.9),  and  aro 

dU  =  -PdV  +  TdS, 
dll  =      VdP  +  T  dflf, 

dA  =  -PdV  -  S  dT, 

dG  =       VdP  -  SdT.  (4.2) 

We  have  already  seen  in  Eq.  (2.6)  how  wo  can  obtain  formulas  from  such 
an  expression.  Wo  can  divide  by  tho  differential  of  one  variable,  say 
du,  and  indicate  that  the  process  is  at  constant  value  of  another,  say  w. 
Thus  wo  have 


In  rlninpr  fjijg     WP  mnwf  Ko  giirp  fh^  fa  is  tho  diffnrnntin.1   of  a  function 

of  tho-Statfi_Qf  the  &v^tojii_-£Qii-£iiil  v  in  that  case  is  it  proper  to  write  R  DRT- 
tial  derivative  like  (Ax/Au^^    Thus  in  particular  we  cannot  proceed  in 

snporfu'.mllv  t.hf>y  look 


1  For  the  method  of  classifying  thermodynamic  formulas  presented  in  Sees.  4  and 
5,  see  P.  W.  Bridgman,  "A  Condensed  Collection  of  Thermodynamic  Formulas," 
Harvard  University  Press. 


SEC.  4] 


THERMODYNAMICS 


25 


like  Eq.  (4.1).  Using  the  method  of  Eq.  (4.3),  a  very  large  number  of 
formulas  can  be  formed.  A  special  case  has  been  seen,  for  instance,  in  the 
Eqs.  (2.7)  and  (2.8).  This  is  the  case  in  which  u  is  one  of  the  variables 
y  or  z,  and  w  is  the  other.  Thus,  suppose  u  =  y,  w  =  z.  Then  we  have 


(!),=«- 


and  similarly 


=  L. 


(4.4) 


It  is  to  be  noted  that,  using  Eq.  (4.4),  we  may  rewrite  Kq.  (4.1)  in  the 
form 


? 


(4.5) 


a  form  in  which  it  becomes  simply  the  familiar  mathematical  equation 
expressing  a  total  differential  in  terms  of  partial  derivatives. 

2.  Suppose  we  have  two  derivatives  such  as  (dx/du)t,  (dy/du)s,  taken 
with  respect  to  t^e  same  variable  and  with  the  same  variable  held  con- 
stant. Since  z  is  held  fixed  in  both  cases,  they  act  like  ordinary  deriva- 
tives with  respect  to  the  variable  u.  But  for  ordinary  derivatives  we 

should  have  T   , =  --,—     Thus,  in  this  case  we  have 

dy/du      ay 


(**\ 

w. 


?V 


A 


(4.6) 


We  shall  find  that  the  relation  in  Eq.  (4.6)  is  of  great  service  in  our  sys- 
tematic tabulation  of  formulas  in  the  next  section.  For  to  find  all  partial 
derivatives  holding  a  particular  z  constant,  we  need  merely  tabulate  the 
six  derivatives  of  the  variables  with  respect  to  a  particular  u,  holding  this 
z  constant.  Then  we  can  find  any  derivative  of  this  type  by  Eq.  (4.6). 
3.  Let  us  consider  Eq.  (4.5),  and  set  x  constant,  or  dx  =  0.  Then 
we  may  solve  for  dy/dz,  and  since  x  is  constant,  this  will  be  (dy/dz)x. 
Doing  this,  we  have 


x  (-} 

dy\    _      \dz/ 

dZ/x  /dx\ 


(4.7) 


26  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  II 

Using  Eq.  (4.6),  we  can  rewrite  Eq.  (4.7)  in  either  of  the  two  forms 


(4.8) 
(4.9) 


, 


I     (?*\  (**}(*}    __! 
\\dyj  \te/\dxL 


The  reader  should  note  carefully  the  difference  between  Eq.  (4.8)  and 
Eq.  (4.6).  At  first  glance  they  resemble  each  other,  except  for  the  differ- 
ence of  sign;  but  it  will  be  noted  that  in  Eq.  (4.6)  each  of  the  throe  deriva- 
tives has  the  same  variable  held  constant,  while  in  Eq.  (4.8)  each  one  has  a 
different  variable  held  constant. 

4.  We  Start  with  Eq.  (4.4).     Then  we  use  the  fundamental  theorem 
regarding  second  partial  derivatives: 


Substituting  from  Eq.  (4.4),  this  gives  us 

/a*\      /a/A 

\dz)y       \dy)z 


(4.10) 


(4.11) 


In  Eq.  (4.10),  it  is  essential  that  .r  be  a  function  of  the  state  of  the  system, 
or  of  y  and  z.     Four  important  relations  result  from  applying  Eq.  (4.11) 

to  the  differential?  expressions  (4.2).     These  are 

1    t 

~  jtSJv 


dP/.« 


dTv  \dV 


dT 


dP 


(4.12) 


The  Eqs.  (4.12)  are  pjcnerallv  called  Maxwell's  relations. 

We  have  now  considered  the  four  processes  used  in  deriving  thermo- 
dynamic  formulas.  By  combinations  of  them,  any  desired  relation 
connecting  first  derivatives  can  be  obtained.  In  the  next  section  we 
consider  the  classification  of  these  formulas. 


SEC.  5]  THERMODYNAMICS  27 

5.  General  Classification  of  Thermodynamic  Formulas. — Bridgman1 
has  suggested  a  very  convenient  method  of  classifying  all  the  thermody- 
namic  formulas  involving  first  derivatives.  As  we  have  pointed  out,  a 
relation  can  be  found  between  any  four  of  the  derivatives.  Bridpman's 
method  is  then  to  write  each  derivative  in  terms  of  three  standard  deriva- 
tives,  for  which  he  chooses  (dV/dT)p_,  (dV/dP)T,  and  CP  =  (dH/dT)P 
These  are  chosen  because  they  can  DC  found  immediately  from  expori- 
ment,  the  first  two  being  closely  related  to  thermal  expansion  and 
compressibility  [see  Eqs.  (2.1)  and  (2.2)].  If  now  we  wish  a  relation 
between  two  derivatives,  we  can  write  each  in  terms  of  these  standard 
derivatives,  and  the  relations  will  immediately  become  plain.  Our  task, 
then,  is  to  find  all  but  these  three  of  the  336  first  partial  derivatives,  in 
terms  of  these  three.  As  a  matter  of  fact,  we  do  not  have  to  tabulate 
nearly  all  of  these,  on  account  of  the  usefulness  of  Eq.  (4.6).  We  shall 
tabulate  all  derivatives  of  the  form  (dx/dT)^  (dx/BP)r.  (dx/dT)v.  and 
(dx/dT)s^  Then  by  application  of  Eq.  (4.6),  we  can  at  once  find  any 
derivative  whatever  at  constant  P,  constant  77,  constant  V,  or  constant 
S.  We  could  continue  the  same  thing  for  finding  derivatives  holding  the 
other  quantities  fixed;  but  we  shall  not  need  such  derivatives  very  often, 
and  they  are  very  easily  found  by  application  of  methods  (2)  and  (3)  of 
the  preceding  section,  and  by  the  use  of  our  table.  Wo  shall  now  tabulate 
these  derivatives,  later  indicating  the  derivations  of  the  only  ones  that 
are  at  all  involved,  and  giving  examples  of  their  application.  We  shall 
be  slightly  more  general  than  Bridgman,  in  that  alternative  forms  are 
given  for  some  of  the  equations  in  terms  either  of  CP  or  (7, . 

TABLE  1-1. — TABLE  OF  TIIBUMODYNAMH;  RELATIONS 


(dS\     =CV 
\dT/P        T 


=  -P  -  8 

r  \dTjp 


1  See  reference  under  Sec.  4. 


28  INTRODUCTION  TO  CHEMICAL  PHYSICS  [€HAP.  II 

TABLE  1-1.  —  TABLE  OF  THERMODYNAMIC  RELATIONS  (Continued] 


dI>/ 


AP 


OPT  dP 


(ov\ 

=  _  \dT/P 
/QV\ 

\dP/T 

(dS\    =  Cv  =  CP 
\d'r)v       T        T  + 


dTP 


dP/T 

("} 
aH 


r 
=    ' 


SEC.  5]  THERMODYNAMICS  29 

TABLE  1-1. — TABLE  OF  THERMODYNAMIC  RELATIONS  (Continued) 


fav\~ 
\af)P 


<±  +  -< 

rii     ~T~     > 


ov 


a 


("'-V 

•c,  ,  \#rj, 

W) 

\dPjr 


-T  + 


/dH\    __  VCP 

W-  =  JV 


~ 


2-1 


-s 


The  formulas  of  Table  1-1  all  follow  in  very  obvious  ways  from  the 
methods  of  See.  4,  except  perhaps  the  relation  between  Cv  and  CV,  used 
in  the  derivatives  at  constant  V  and  constant  *S.  To  prove  the  relation 
between  these  two  specific  heats,  we  proceed  as  follows.  We  have 


(5.1) 


Tds  =  r(^ 


+  T\M)TdV  =  CvdT  +  T(  ^ 


=  T(^    dT  +  I 


,(dS 
\dp 


dP  =  CPdT-  T(~ 


We  subtract  the  first  of  Eqs.  (5.1)  from  the  second,  and  sot  dV  =  0, 
obtaining 


(CP-Cv)dT=  T(^-r)dP. 


30  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  II 

Dividing  by  dT7,  this  is 


dp 


T 

the  result  used  in  the  formulas  of  Table  1-1.  The  result  of  Eq.  (5.2)  is 
an  important  formula,  and  is  the  generalization  holding  for  all  substances 
of  the  familiar  formula  CP  —  Cv  =  nR  holding  for  perfect  gases.  It 
serves  in  any  case  to  find  Cp  —  Cv  from  the  equation  of  state.  Since1 
(dV/dP)r  is  always  negative,  we  see  that  CV  is  always  greater  than  C  v  . 
"^  o.  Comparison  of  Thermodynamic  and  Gas  Scales  of  Temperature.- 
In  Chap.  I,  Sec.  4,  we  have  discussed  the  therrnodynamic  temperature, 
introduced  in  the  statement  of  the  second  law  of  thermodynamics,  and  tt  e 
have  mentioned  that  it  can  be  proved  that  this  is  the  same  temperature 
that  is  measured  by  a  gas  thermometer.  We  arc  now  in  position  to  prove 
this  fact.  First,  we  must  define  a  perfect  gas  in  a  way  that  could  be 
applied  experimentally  without  knowing  any  temperature  scale  except 
that  furnished  by  the  gas  itself.  Wo  can  define  it  as  a  gas  which  in  the 
first  place?  obeys  Boyle's  law:  PV  —  constant  at  constant  temperature,  or 
PV  =  f(T),  where  T  is  the  thermodynamic  temperature,  and  /  is  a  func- 
tion as  yet  unknown.  Secondly,  it  obeys  what  is  called  Joule's  la\v:  the 
internal  energy  1T  is  independent  of  the  volume  at  constant  temperature. 
These  assumptions  can  both  be  proved  by  direct  experiment.  We  can 
certainly  observe1  constancy  of  temperature  without  a  temperature  scale, 
so  that  we  can  verify  Boyle's  law.  To  chock  Joule's  law,  we  may  consider 
the  free  expansion  of  the*  gas.  We  let  the  gas  expand  irreversibly  into 
a  vacuum.  It  is  assumed  that  the  process  is  carried  out  aeliabatically, 
and  since  there  is  ne>  external  work  done,  the  internal  energy  is  unchanged 
in  the  process.  We  allow  the  gas  to  come  te)  equilibrium  at  its  new 
volume,  and  observe  whether  the  temperature  is  the?  same  that  it  was 
originally,  or  different.  If  it  is  the  same,  then  the  gas  is  said  to  obey 
Joule's  law.  To  che»ck  the  mathematical  formulation  of  this  law,  we  note 
that  the  experiment  of  free  expansion  tells  us  directly  that  the  tempera- 
ture is  independent  erf  volume,  at  constant  internal  energy:  (3T/dV)u  —  0. 
But  by  Eq.  (4.9),  we  have 

/d[A  (dlf\(dT\  (dT\  ,    n 

\av)r  =  ~\dT)\dv)v  -  ~CvWA  =  °*  ((U) 

Equation  (6.1)  states  that  the  internal  energy  is  independent  of  volume 
at  constant  temperature,  the  usual  statement  of  Joule's  law. 

Without  further  assumption  about  the  gas,  we  can  now  prove  that 
f(T)  =  constant  X  T,  so  that  the  pressure  of  a  perfect  gas  at  constant 


SEC.  G]  THERMODYNAMICS  31 

volume  is  proportional  to  the  thermodynamic  temperature,  and  if  we  uae 
proper  units,  the  gas  scale  of  temperature  is  identical  with  the  thermo- 
dynamic scale.  Using  Table  1-1  we  have  the  important  general  relation 


.         _  _  __     /dP 

av-  - 


/dV\ 
(dp)T 


giving  the  change  of  internal  energy  with  volume  of  any  substance  at 
constant  temperature.  We  set  this  equal  to  zero,  on  account  of  Joule's 
law.  From  the  equation  of  state, 

f(T)        J'(T) 

Substituting  Eq.  (6.3)  in  Kq.  (6.2),  and  canceling  out  a  factor  P,  we  have 

Tf(T)  = 


or 

d  In/  =  d  In  r,         ln/(T)  =  In  T  +  const., 

f(T)  -  const.  X  T,  (6.4) 

which  was  to  be  proved. 

Instead  of  defining  a  perfect  gas  as  we  have  done,  by  Boyle's  law  and 
Joule's  law,  we  may  prefer  to  assume  that  a  thermodynamic  temperature 
scale  is  known,  and  that  the  perfect  gas  satisfies  the  general  gas  law 
PV  =  const.  X  T.  Then  we  can  at  once  use  the  relation  (6.2)  to  calcu- 
late the  change  of  internal  energy  with  volume  at  constant  temperature, 
and  find  it  to  be  zero.  That  is,  we  show  directly  by  thermodynamics  that 
Joule's  law  follows  from  the  gas  law,  if  that  is  stated  in  terms  of  the 
thermodynamic  temperature. 


CHAPTER  III 
STATISTICAL  MECHANICS 

Thermodynamics  is  a  simple,  general,  logical  science,  based  on  two 
postulates,  the  first  and  second  laws  of  thermodynamics.  We  have  seen 
in  the  last  chapter  how  to  derive  results  from  these  laws,  though  we  have 
not  used  them  yet  in  our  applications.  But  we  have  seen  that  they  are 
limited.  Typical  results  are  like  Kq.  (5.2)  in  Chap.  II,  giving  the  differ- 
ence of  specific  heats  of  any  substance,  Cp  —  CV,  in  terms  of  derivatives 
which  can  be  found  from  the  equation  of  state.  Thermodynamics  can 
give  relations,  but  it  cannot  derive  the  specific  heat  or  equation  of  state 
directly.  To  do  that,  we  must  go  to  the  statistical  or  kinetic  methods. 
Even  the  second  law  is  simply  a  postulate,  verified  because  it  leads  to 
correct  results,  but  not  derived  from  simpler  mechanical  principles  as  far 
as  thermodynamics  is  concerned.  We  shall  now  take  up  the  statistical 
method,  showing  how  it  can  lead  not  only  to  the  equation  of  state  and 
specific  heat,  but  to  an  understanding  of  the  second  law  as  well. 

1.  Statistical  Assemblies  and  the  Entropy. — To  apply  statistics  to 
any  problem,  we  must  have  a  great  many  individuals  whose  average 
properties  we  are  interested  in.  We  may  ask,  what  are  the  individuals  to 
which  we  apply  statistics,  in  statistical  mechanics?  The  answer  is,  they 
are  a  great  many  repetitions  of  the  same  experiment,  or  replicas  of  the 
same  system,  identical  as  far  as  all  large-scale,  or  macroscopic,  properties 
are  concerned,  but  differing  in  the  small-scale,  or  microscopic,  properties 
which  we  cannot  directly  observe.  A  collection  of  such  replicas  of  the 
same  system  is  called  a  statistical  assembly  (or,  following  Gibbs,  an 
ensemble).  Our  guiding  principle  in  setting  up  an  assembly  is  to  arrange 
it  so  that  the  fluctuation  of  microscopic  properties  from  one  system  to 
another  of  the  assembly  agrees  with  the  amount  of  such  fluctuation  which 
would  actually  occur  from  one  repetition  to  another  of  the  same 
experiment. 

Let  us  ask  what  the  randomness  that  we  associated  with  entropy  in 
Chap.  I  means  in  terms  of  the  assembly.  A  random  system,  or  one  of 
large  entropy,  is  one  in  which  the  microscopic  properties  may  be  arranged 
in  a  great  many  different  ways,  all  consistent  with  the  same  large-scale 
behavior.  Many  different  assignments  of  velocity  to  individual  mole- 
cules, for  instance,  can  be  consistent  with  the  picture  of  a  gas  at  high 
temperatures,  while  in  contrast  the  assignment  of  velocity  to  molecules 
at  the  absolute  zero  is  definitely  fixed:  all  the  molecules  are  at  rest.  Then 

32 


SEC.  1]  STATISTICAL  MECHANICS  33 

to  represent  a  random  state  we  must  have  an  assembly  which  is  dis- 
tributed over  many  microscopic  states,  the  randomness  being  measured 
by  the  wideness  of  the  distribution.  We  can  make  this  idea  more  precise. 
Following  Planck,  we  may  refer  to  a  particular  miprosnopir.  stn.t.o.  nf  thu 
system  as  a  complexion.  We  may  describe  an  assembly  by  stating  what 
fraction  of  the  systems  of  the  assembly  is  found  in  each  possible  com- 
plexion. We  shall  call  this  fraction,  for  the  it\\  complexion.  /V.  and  shall 
refer  to  the  sot  of  /t's  as  the  distribution  function  describing  tho  assembly. 
Plainly,  since  all  systems  must  be  in  one  complexion  or  another, 


Then  in  a  random  assembly,  describing  a  system  of  large  entropy,  there 
will  be  systems  of  the  assembly  distributed  over  a  great  many  complex- 
ions, so  that  many  fl's  will  be  different  from  zero,  each  ono  of  these  frac- 
tions being  necessarily  small.  On  the  other  hand,  in  an  assembly  of  low 
entropy,  systems  will  be  distributed  over  only  a  small  number  of  com- 
plexions, so  that  only  a  few  /t's  will  be  different  from  zero,  and  these  will 
bo  comparatively  large. 

We  shall  now  postulate  a  mathematical  definition  of  entropy,  in  terms 
of  the/?s,  whicTTlij  lafiafm  the  case  of  a  random  distribution,  small  other- 
wise! This  definition  is 

(1.2) 


Hero  k  is  a  constant,  called  Boltzmann's  constant,  which  will  appear  fre- 
quently in  our  statistical  work.  It  has  the  same  dimensions  as  entropy. 
or  specific  heat,  that  is.  energy  divided  by  tnmpnratiim.  ^Its  value  in 
absolute  units  is  1.379  X  1Q~16  er  er  degree.  This  value  is  derived 


indirectly;  using  Eq.  (1.2)?  for  thn  pnt.rnpyj  <™<*  ^«"  flfTJv^  tl 

gas  law  and  the  gas  constant,  in  torms  of  fc,  thereby  determining  k  from 

It  is  easy  to  see  that  Eq.  (1.2)  has  the  required  property  of  being 
large  for  a  randomly  arranged  system,  small  for  one  with  no  randomness. 
If  there  is  no  randomness  at  all,  all  values  of  fl  will  be  zero,  except  one, 
which  will  be  unity.  But  the  function  /  In  /  is  zero  when  /  is  either  zero  or 
unity,  so  that  the  entropy  in  this  case  will  be  zero,  its  lowest  possible 
value.  On  the  other  hand,  if  the  system  is  a  random  one,  many  com- 
plexions will  have  /t  different  from  zero,  equal  to  small  fractions,  so  that 
their  logarithms  will  be  large  negative  quantities,  and  the  entropy  will  be 
large  and  positive.  We  can  see  this  more  clearly  if  we  take  a  simple 
case:  suppose  the  assembly  is  distributed  through  W  complexions,  with 


34  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 


equal  fractions  in  each.  The  value  rf  *y>.nh  f-  in  titlfififi  finTnplpr;/™°  is 
1/W,  while  for  other  complexions  /»•  is  zero.  Then  we  have 

Q  —  _J&.T/F_  in  — 

o  —        /t  rr  jrr  "I  ™- 
L  =  fc  In  W  (1.3) 

Tho^p  j^py,  in  «nr»h  a  r»flfflj  j<g  nrflpprtional  to  the  logarithm  of  the  number 
nf  nnrnplpviflna  in  whipli  ay^ff  TUB  of  thfi  assembly  can  be  found.  AsTtnis 
number  of  complexions  increases,  the  distribution  becomes  more  random 
or  diffuse,  and  the  entropy  increases. 

Boltzmann1  based  his  theory  of  t.hn  relation  of  probability  to  entropy 
on  Eq.  (1.3),  rather  than  using  the  more  general  relation  M  .2V  HP  pnllod 
)V  the  thermodynamic  probability  "f  *  «^tf  r  P-rtPI^  much  as  wo  have 
that  a  random  state,  which  is  inherently  likely  to  be  realized,  will  have  a 
large  value  of  W.  Planck2  has  shown  by  the  following  simple  argument 
that  the  logarithmic  form  of  Eq.  (1.3)  is  reasonable.  Suppose  the  system 
consists  of  two  parts,  as  for  instance  two  different  masses  of  gas,  riot  con- 
nectod  with  each  other.  In  a  given  state,  represented  by  a  given  assem- 
bly, let  there  be  Wi  complexions  of  the  first  part  of  the  system  consistent 
with  the  macroscopic  description  of  the  state,  and  Wo  complexions  of  the 
second  part.  TV™  ,  «j"™>  *h*  *™"  p?rf  s  nf  f1"*  *v*tf  ™  f^re  independonTof 
each  otherTthere  must  be  W^W^  comnlexioiis  of  the  combined  system, 
since  each  complexion  of  the  first  part  can  be  joined  to  any  one  of  the 
complexions  of  the  second  part  to  give  a  complexion  of  the  combined 
system.  We  shall  then  find  for  the  entropy  of  the  combined  system 

S  *=  k  In  WiWz 

=  k\nWi  +  k\n  W2.  (1.4) 

But  if  we  considered  the  first  part  of  the  system  by  itself,  it  would  have 
an  entropy  Si-»  k  In  Wi,  and  the  second  part  by  itself  would  have  an 
entropy  /S>2  =  Arin  TF2.  Thus,  on  account  of  the  relation  (1,3),  we  have 

S  =  S!  +  S2.  (1.5) 

But  surely  this  relation  must  bo  true;  in  thermodynamics,  the  entropy 
of  two  separated  systems  is  tho  sum  of  the  entropies  of  the  parts,  as  wo 
can  see  directly  from  the  second  law,  since  the  changes  of  entropy,  dQ/T, 
in  a  reversible  process,  are  additive.  Then  we  can  reverse  the  argument 
fl-bovp.  Fjfliiation  (\.&\  must  be  trno.  &iicl  if_tliii-£iiiiti*Qny  is  a,  function  of 
W}  it  can  be  shown  that  the  only  possible  function  consistent  with  the 
additivity  of  the  entropy  is  tho  logarithmic  function  of  ECL  fl.SV 

1See  for  example,  L.  Boltzmarm,  "Vorlesungen  uber  Gastheorie,"  Sec.  6,  J.  A. 
Barth. 

2  See  for  example,  M.  Planck,  "Heat  Radiation,"  Sec.  119,  P.  Blakiston's  Sons  & 
Company. 


SEC.  1]  STATISTICAL  MECHANICS  35 

Going  back  to  our  more  general  formula  (1.2),  we  can  show  that  if 
the  assembly  is  distributed  through  W  complexions,  the  entropy  will  have 
its  maximum  value  when  the  /»'s  are  of  equal  magnitude,  and  is  reduced 
by  any  fluctuation  in  /t  from  cell  to  cell,  verifying  that  any  concentration 
of  systems  in  particular  complexions  reduces  the  entropy.  Taking  the 
formula  (1.2)  for  entropy,  wo  find  how  it  changes  when  the/t's  aro  varied. 
Differentiating,  wo  havo  at  onco 


dS  =  -*](l  +  In  /.MA-  (1-6) 

i 

But  we  know  from  Eq.  (1.1)  that  X/t  =  1,  from  which  at.  onco 

i 

,  =  0.  (1.7) 


Thus  the  first  term  of  Kq.  (1.6)  vanishes;  and  if  we  assume  that  the 
density  is  uniform,  so  that  ln/»  is  really  independent  of  /„  we  can  take  it- 
out  of  the  summation  in  Eq.  (1.6)  as  a  common  factor,  and  the  remaining 
term  will  vanish  too,  giving  dS  =  0.  That  is.  fof  uniform  ffan^yj  fl™ 
variation  of  the  entropy  for  small  variations  of  the  asscn^^fv  vfl-fllfihfifti  **- 
necessary  condition  for  a  maximum  of  the  entropy.  A  little  further 
investigation  would  convince  us  that  this  really  gives  a  maximum,  not  a 
minimum,  of  entropy,  and  that  in  fact  Eq.  (1.3)  gives  the  absolute  maxi- 
mum, the  highest  value  of  which  S  is  capable,  so  long  as  only  W  complex- 
ions are  represented  in  the  assembly.  The  only  way  to  get  a  still  greater 
value  of  S  would  be  to  havo  more  terms  in  the  summation,  so  that  each 
individual  /»  could  bo  even  less. 

We  have  postulated  a  formula  for  the  entrypy.  How  can  we  expeot, 
to  prove  that  it  is  correct?  We  can  do  this  onlv  bv  goiny  bank  to  thn_ 
second  law  of  thermodynamics^  showing  that  our  entropy  has  the  prop- 
erties ciemanciea  Dy  ttiat  law,  and  that  in  forma  of  it  thn  Uw  i«  «a.t.i«finfj. 
We  have  already  shown  that,  our  formula  for  the  entropy  has  one  of  the 
properties  demanded  ot  the  entropy:  it  is  determined  bv  the  Ht.fl.te  of  thn 
system.  In  statistical  mechanics,  the  only  thing  we  can  mean  by  the 
state  of  the  system  is  the  statistical  assembly,  since  this  determines 
average  or  observable  properties  of  all  sorts,  and  our  formula  (1.2)  for 
entropy  is  determined  by  the  statistical  assembly.  Next  we  must  show 
that  our  formula,  represents  a  quantity  that  increases  in  an  irreversible* 
process.  This  will  be  done  by  qualitative  but  valid  reasoning  in  a  later 
section.  It  will  then  remain  to  consider  thermal  equilibrium  and  reversi- 
ble processes,  and  to  show  that  in  such  processes  the  change  of  entropy  is 
dQ/T. 


36  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 

2.  Complexions  and  the  Phase  Space. — We  wish  to  find  how  our 
formula  for  the  entropy  changes  in  an  irreversible  process.  To  do  this, 
we  must  find  how  the/t's  change  with  time,  or  how  systems  of  the  assem- 
bly, as  time  goes  on,  change  from  one  complexion  to  another.  This  is  a 
problem  in  kinetics,  and  we  shall  not  take  it  up  quantitatively  until  the 
chapter  on  kinetic  methods.  For  the  present  we  shall  be  content  with 
qualitative  discussions.  The  first  thing  that  we  must  do  is  to  get  a  more 
precise  definition  of  a  complexion.  We  have  a  certain  amount  of  informa- 
tion to  guide  us  in  making  this  definition.  We  are  trying  to  make  our 
definition  of  entropy  agree  with  experience,  and  in  particular  we  want  the 
state  of  maximum  entropy  to  be  the;  stable,  equilibrium  state.  But  we 
have  just  seen  that  for  an  assembly  distributed  through  W  complexions, 
the  state  of  maximum  entropy  is  that  in  which  equal  numbers  of  systems 
are  found  in  each  complexion.  This  is  commonly  expressed  by  saying 
that  complexions  have  equal  a  priori  probability;  that  is,  if  we  have  no 
specific  information  to  the  contrary,  we  are  as  likely  to  find  a  system  of  an 
assembly  in  one  complexion  as  in  another,  in  equilibrium.  Our  definition 
of  a  complexion,  then,  must  be  consistent  with  this  situation. 

The  method  of  defining  complexions  depends  on  whether  we  are  treat- 
ing our  systems  by  classical,  Newtonian  mechanics  or  by  quantum  theory. 
First  we  shall  take  up  classical  mechanics,  for  that  is  more  familiar.  But 
later,  when  we  describe  the  methods  of  quantum  theory,  we  shall  observe 
that  that  theory  is  more  correct  and  more  fundamental  for  statistical 
purposes.  In  el^ssien.1  ^elianics.  a  system  is  f]esr*rihpfl  by  giving  th^ 
coordinates  and  velocities  of  all  its  particles.  -Instead  of  the  velocities, 
it  proves  to  be  more  desirable  to  use  the  momenta.  With  rectangular 
coordinates,  the  momentum  associated  with  each  coordinate  is  simply  the 
mass  of  the  particle  times  the  corresponding  component  of  velocity;  with 
angular  coordinates  a  momentum  is  an  angular  momentum ;  and  so  on.  If 
there  are  N  coordinates  and  N  momenta  (as  for  instance  the  rectangular 
coordinates  of  JV/3  particles,  with  their  momenta),  we  can  then  visualize 
the  situation  by  setting  up  a  2N  dimensional  space,  called  a  phase  space, 
in  which  the  coordinates  and  momenta  are  plotted  as  variables,  and  a 
single  point,  called  a  representative  point,  gives  complete  information 
about  the  system.  An  assembly  of  systems  corresponds  to  y|  p.nllop.finn  nf 
representative  points,  and  we  shall  generally  assume  that  ty™  QTO  san 
many  systems  in  the  assembly  that  the  distribution  of  representative 
points  is  practically  continuous  in  tliejjhase  space.  Now  a  complexion, 

Or  microSCOPic  State,  of  thp  system  mn«f.  nopiypnnH  t.n  n 

or  small  region,  of  the  phase  space;  to  bemore  precise, 
spond  to  a  small  volume  of  the  phase  space.  We  subdivide  the  whole 
phase  space  into  small  volume  elements  and  call  each  volume  element  a 
complexion,  saying  that  ftt  the  fraction  of  systems  of  the  assembly  in  a 


SBC.  2]  STATISTICAL  MECHANICS  37 

particular  complexion,  simply  equals  the  fraction  of  all  representative 
points  in  the  corresponding  volume  clement.  The  only  question  that 
'q  fllp  ghqpg  and  SJEP  of  Y^^mc  elements  representing 


complexions. 
"""To  answer  this  question,  we  must  consider  how  points  move  in  the 


phase  space.  We  must  know  the  time  rates  of  chuufiQ  of  all  coordinates 
and  momenta,  in  terms  of  the  coordinates  and  momenta  themselves. 
Newton's i  second  law  gives  us  the  time  rate  of  change  of  each  momentum, 
stating  that  it  equals  the  corresponding  component  of  force,  which  is  a 
function  of  the  coordinates  in  a  conservative  system.  The  time  rate  of 
change  of  each  coordinate  is  simply  the  corresponding  velocity  com- 
ponent, which  can  be  found  at  onee  from  the  momentum.  Thus  we  can 
find  what-  is  essentially  the  2N  dimensional  velocity  vector  of  each  repre- 
sentative point.  This  velocity  vector  is  determined  at  each  point  of 
phase  space  and  defines  a  rate  of  flow,  the  representative  points  streaming 
through  the  phase  space  as  a  fluid  would  stream  through  ordinary  space. 
We  are  thus  in  a  position  to  find  how  many  points  enter  or  leave  each 
element  of  volume,  or  each  complexion,  per  unit  time,  and  therefore  to 
find  the  rate  at  which  the  fraction  of  systems  in  that  complexion  changes 
with  time.  It  is  now  easy  to  prove,  from  the  equations  of  motion,  a 
general  theorem  callec}  LiouvihVs  theorem. }  This  theorem  states,  in 
mathematical  language,  the  following  fact :  the  swarm  of  points  moves  in 
such  a  wfl.y  f.Vm|,  the  H?*ic"fy  "f  p™***^  «*•*  «™  fallow  along  with  the  swarm, 
never  changes^  The  flow  is  like  a  streamline  flow  of  an  incompressible 
fluid,  each  particle  of  fl*^]  nlwnya  proMf»rvingr  it.M  r>Wn  density.  This  does 
not  mean  that  the  rlnnsitv  ut,  a.  given  point  of  spn.o.n  dons  not  rhginge  with 
time;  in  general  it  does,  for  in  the  course  uf  the  flow,  first  a.  f|on«o  pnrf  W 
the  swarm,  then  a  less  dense  onn,  m,«i.y  woll  be  swept,  by  f-.hn  point,  jjj_ 
question,  as  if  we  had  an  incompressible  fluid,  but  one  whose  density 
changed  from  point  to  point.  It  does  mean,  however,  that  we  can  find  a 
very  simple  condition  which  is  necessary  and  sufficient  for  the  density 
at  a  given  point  of  space  to  be  independent  of  time:  the  density  of  points 
must  be  constant  all  along  each  streamline,  or  tube  of  flow,  of  the  points. 
For  then,  no  matter  how  long  the  flow  continues,  the  portions  of  the 
swarm  successively  brought  up  to  the  point  in  question  all  have  the  same 
density,  so  that  the  density  there  can  never  change. 

To  find  the  condition  for  equilibrium,  then,  we  must  investigate 
the  nature  of  the  streamlines.  For  a  periodic  motion,  a  streamline  will  be 
closed,  the  system  returning  to  its  original  state  after  a  single  period. 
This  is  a  very  special  case,  however;  most  motions  of  many  particles  are 
not  periodic  and  their  streamlines  never  close.  Rather,  they  wind  around 

1  For  proof,  see  for  example,  Slater  and  Frank,  "  Introduction  to  Theoretical 
Physics,"  pp.  365-366,  McGraw-Hill  Book  Company,  Inc.,  1933. 


38  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 

in  a  very  complicated  way,  coming  in  the  course  of  time  arbitrarily  close 
bo  every  point  of  phase  space  corresponding  to  the  same  total  energy  (of 
course  the  energy  cannot  change  with  time,  so  that  the  representative 
point  must  stay  in  a  region  of  constant  energy  in  the  phase  space).  Such 
a  motion  is  called  quasi-ergodic?  and  it  can  be  shown  to  be  the  general 
hypfi  of  rnofirm  pnrinHif*.  motions  being  a  rare  pv^^n|(jpT^  Then,  from  the 
statement  in  the  last  paragraph,  we  see  that  to  have  a  distribution  inde- 
pendent of  time,  we  must  have  a  density  of  DOII^S  in  phase  apftf.e  whinh 
is  constant  for  all  regions  of  the  same  energy,  _  But  on  the  0ther  hand 
thermal  equilibrium  must  correspond  to  a  distribution  inHp.pondnnt.  nf 
time,  and  we  have  seen  that  the  state  of  maximum  entropy  is  one  in  which 
all  complexions  have  the  same  number  of  avstgiflfi,  Tfrpsp  lwn  «*«-fg- 
menta  are  only  compatible  if  each  complexion  corresponds  to  the  same 
volume  of  phase  space.  For  then  a  constant  volume  density  of  points, 
which  by  Liouville's  theorem  corresponds  to  a  distribution  independent 
of  time,  will  at  the  same  time  correspond  to  a  maximum  entropy.  We 
thus  draw  the  important  conclusion  that  regions  of  equal  volume  in  phase 
space  have  equal  a  priori  probability,  pr  that.  ^  nomple^on  norr^ponds 
to  a  quite  definite  volume  of  phase  space.  Classical  mechanics,  however, 
does  not  lead  to  any  way  of  sayi_ri^hQw  1Qrp?°  thiil  Arolume  is.  Thi^g  it 
cannot  lead  to  any  unique  definition  of  the  entropy;  fr>r  tJhfi  Ajfl  dfiPftnr*  on 
how  large  a  volume  each  complexion  corrggpQiifo  t.n;.  «.nH  they  jp  turn 
determine  the  entropy. 

3.  Cells  in  the  Phase  Space  and  the  Quantum  Theory.  —  Quantum 
mechanics  starts  out  quite  differently  from  classical  mechanics.  It  does 
not  undertake  to  say  how  the  coordinates  and  momenta  of  the  particles 
change  as  time  goes  on.  Rather,  it  is  a  statistical  theory  from  the  begin- 
ning: it  sets  up  a  statistical  assembly,  and  tells  us  directly  how  that 
assembly  changes  with  time,  without  the  intermediate  step  of  solving  for 
the  motion  of  individual  systems  by  Newton's  laws  of  motion.  And  it 
describes  the  assembly,  from  the  outset,  in  terms  of  definite  complexions, 
so  that  the  problem  of  defining  the  complexions  is  answered  as  one  of  the 
postulates  of  the  theory.  It  sets  up  quantum  states,  of  equal  a  priori 
probability,  and  describes  an  assembly  by  giving  the  fraction  of  all  sys- 
tems in  each  quantum  state.  Instead  of  giving  laws  of  motion,  like 
Newton's  second  law,  its  fundamental  equation  is  one  telling  how  many 
systems  enter  or  leave  each  quantum  state  per  second.  In  particular,  if 
?qual  fractions  of  the  systems  are  found  in  all  quantum  states  associated 
with  the  same  energy,  we  learn  that  these  fractions  will  not  change  with 
time;  that  is,  in  a  steady  or  equilibrium  state  all  the  quantum  states  are 
equally  occupied,  or  have  equal  a  priori  probabilities,  ye  are  then 

jiiat.ififtfl  in  iHpnt.ifvinp;  t.hpgP  qimniiiinj  af[ft.f.ofl 


innc  ^yV.^V.  Wg  ViQir*>  moyif^™/^      When  we  deal  witL  quantum  statistics, 


SBC.  3)  STATISTICAL  MECHANICS  39 


A  will  refer  to  the  fraction  of  all  svstmns  in  tfif  tt.h  niianfnm  g+Qfo  This 
gives  a  definite  meaning  |.n  t.hn  pnmpWinnfi  BT1fJ  i™^c  ^  o  ^»finjt.P  nni^n^ 
ical  value  for  the  entropy. 

Quantum  tfteory  provides  no  unique  way  of  setting  up  the  quantum 
states,  or  the  complexions.  We  can  understand  this  much  better  by 
considering  the  phase  space.  iyrn.nv  feature^  of  the  Quantum  theory  can 
be  described  by  dividing  the  phase  space  into  cells  of  equal  volume,  and 
associating  each  cell  with  a  quantum  state.  The  volume  of  these  cells  is 
uniquely  fixed  by  the  quantum  theory,  but  not  their  shape.  We  can,  for 
example,  take  simply  rectangular  cells,  of  dimensions  A#i  along  the  axis 
representing  the  first  coordinate,  A<?2  for  the  second  coordinate,  and  so  on 
up  to  A#.v  for  the  Nth  coordinate,  and  Api  to  ApN  for  the  corresponding 
momenta.  Then  there  is  a  very  simple  rule  giving  the  volume  of  such  a 
cell:  we  have 

A?tAp,  =  A,  (3.1) 

where  h  is  Planck's  constant,  equal  numerically  to  6.61  X  10~~7  absolute 
units.  Thus,  with  N  coordinates,  the  2Ar-dimensional  volume  of  a  cell 
is  hN. 

We  can  equally  well  take  other  shapes  of  cells.  A  method  which  is 
often  useful  can  be  illustrated  with  a  problem  having  but  one  coordinate  q 
and  one  momentum  p.  Then  in  our  two-dimensional  phase  space  we  can 
draw  a  curve  of  constant  energy.  Thus  for  instance  consider  a  particle* 
of  mass  m  held  to  a  position  of  equilibrium  by  a  restoring  force  propor- 
tional to  tho  displacement,  so  that  its  energy  is 

E  -  SL  +  27r2w*>y,  (3.2) 


where  v  is  the  frequency  with  which  it  would  oscillate  in  classical 
mechanics.  The  curves  of  constant  energy  are  then  ellipses  in  the  p-q 
space,  as  we  see  by  writing  the  equation  in  the  form 


02 


the  standard  form  for  the  equation  of  an  ellipse  of  semiaxes  -\/2mE  and 
\/E/2ir2mv'2.  Such  an  ellipse  is  showrn  in  Fig.  III-l .  Then  we  can  choose 
cells  bounded  by  such  curves  of  constant  energy,  such  as  those  indicated 
in  Fig.  III-l.  Since  the  area  between  curves  must  be  A,  it  is  plain  that  the 
nth  ellipse  must  have  an  area  nA,  where  n  is  an  integer.  The  area  of  an 
ellipse  of  semiaxes  a  and  b  is  irdb\  thus  in  this  case  we  have  an  area  of 
ir\/2mE\/E/2'ir2mvz  =  E/v,  so  that  the  energy  of  the  ellipse  connected 


40 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  Ill 


with  a  given  integer  n  is  given  by 

En  =  nhv. 


(3.4) 


Another  illustration  of  this  method  is  provided  by  a  freely  rotating  wheel 
of  moment  of  inertia  7.  The  natural  coordinate  to  use  to  describe  it  is 
the  angle  0,  and  the  corresponding  momentum  pQ  is  the  angular  momen- 

P  

'TrrTT 


Kiu.   1II-1       Cells  in  phase  spare,   for  the  linear  oscillator       The  shaded  area,   between 
two  ellipses  of  constant  energy,  has  an  area  h  in  the  qinntum  theory 

turn,  /co,  \\licre  co  =  dO/dt  is  the  angular  velocity.     If  no  torques  act,  the 
energy  is  wholly  kinetic,  equal  to 


E  =  ^ 


(3.5) 


Then,  as  shown  in  Fig.  III-2,  lines  of  constant  energy  are  straight  lines  at 
constant  value  of  po.     Since  0  goes  from  zero  to  2?r,  and  then  the  motion 


P0 


3H/27T 


repeats,  we*  use  only  values  of  the  coordinate 
in  this  range.  Then,  if  the  cells  an*  set  up  so 
that  the  area  of  each  is  h,  we  must  have  them 
bounded  by  the  lines 


nh 


(3.6) 


so  that  the  energy  associated  with  the  nth 
line  is 

E.  =  Sr  (3-7) 


Km.  III-2 —Cells  m  phase         j[u  forms  of  these  cells,  we  can  now  under- 

space,    for    the    rotator.     The  xr       j  xix±  x 

shaded  area  has  an  area  of  h  stand  one  of  the  most  fundamental  statements 
in  the  quantum  theory.  of  ^ jie  quantum  theory,  the  principle  of  uncer- 

tainty :  it  is  impossible  to  regulate  the  coordinates  and  momenta  of  a  system 
more  accurately  than  to  require  that  they  lie  somewhere  within  a  given  cell. 
Any  attempt  to  be  more  precise,  on  account  of  the  necessary  clumsiness 
of  nature,  will  result  in  a  disturbance  of  the  system  just  great  enough  to 


SEC.  3]  STATISTICAL  MECHANICS  41 

shift  the  representative  points  in  an  unpredictable  way  from  one  part  of 
the  cell  to  another.  The  best  we  can  do  in  setting  up  an  assembly,  in 
other  words,  is  to  specify  what  fraction  of  the  systems  will  be  found  in 
each  quantum  state  or  complexion,  or  to  give  the  /,'s.  This  does  not. 
imply  by  any  means,  however,  that  it  does  not  make  sense  to  talk  about 
the  coordinates  and  momenta  of  particles  with  more  accuracy  than  to 
locate  the  representative  point  in  a  given  cell.  Xh^re  is  nothing  inher- 
ently impossible  in  knowing  the  coordinates  and  momenta  of  a  system  as 
accurately  as  wo  please;  the  restriction  is  only  that  we  cannot  prepare  a 
system,  or  an  assembly  of  systems,  with  as  precisely  determined  coordi- 
nates  and  momenta  as  we  might  please. 

Since  we  may  be  interested  in  precise  values  of  the  momenta  and 
coordinates  of  a  system,  there  must  bo  something  in  the  mathematical 
framework  of  the  theory  to  describe  them.  We  must  be  able  to  answer 
questions  of  this  sort:  given,  that  an  assembly  has  a  given  fraction  of  its 
systems  in  each  coll  of  phase  space,  what  is  the  probability  that  a  certain 
quantity,  such  as  ono  of  the  coordinates,  lies  within  a  certain  infinitesimal 
range  of  valuos?  Put  in  another  way,  if  we  know  that  a  system  is  in  a 
given  coll,  what  is  the  probability  that  its  coordinates  and  momenta  lie 
in  definite  ranges?  The  quantum  theory,  and  specifically  the  wave 
ID'  chanics,  can  answer  such  questions;  and  because  it  can,  wo  aro  justified 
ia  regarding  it  as  an  essentially  statistical  theory.  By  experimental 
methods,  we  can  insure  that  a  system  lies  in  a  given  cell  of  phase  space. 
That  is,  we  can  prepare  an  assembly  all  of  whose  representative  points  lie 
in  this  single  cell,  but  this  is  the  nearest  we  can  come  to  setting  up  a  sys- 
tem of  quite  definite  coordinates  and  momenta.  Having  prepared  such 
an  assembly,  however,  quantum  theory  says  that  the  coordinates  and 
momenta  will  be  distributed  in  phase  space  in  a  definite  way,  quite  inde- 
pendent of  the  way  we  prepared  the  assembly,  and  therefore  quite  unpre- 
dictable from  the  previous  history  of  the  system.  In  other  words,  all 
that  the  theory  can  do  is  t.o  p;ivo.  ijs  statistical  information  about  a  system. 
not  jrlp^ily/*  krmwlodge  of  exactly  what  it  will  do.  This  is  in  striking 
contrast  to.  the  classical  mechanics,  which  allows  precise  prediction  of  the 
future  of  a  system  if  we  know  its  past  history. 

The  cells  of  the  type  described  in  Figs.  III-l  and  III-2  have  a  special 
property:  all  the  systems  in  such  a  quantum  state  have  the  same  energy. 
The  momenta  and  coordinates  vary  from  system  to  system,  roughly  as  if 
systems  were  distributed  uniformly  through  the  cell,  as  for  example 
through  the  shaded  area  of  either  figure,  though  as  a  matter  of  fact  the 
real  distribution  is  much  more  complicated  than  this.  But  the  energy  is 
fixed,  the  same  for  all  systems,  and  is  referred  to  as  an  energy  level.  It  is 
equal  to  some  intermediate  energy  value  within  the  cell  in  phase  space,  as 
computed  classically.  Thus  for  the  oscillator,  as  a  matter  of  fact,  the 


42  INTRODUCTION  TO  CHEMICAL  PHYMCti  [CHAP.  Ill 

energy  levels  arc 

E»  =  (*  +  JV'  (3.8) 

\         */ 

which,  as  we  see  from  Eq.  (3.4),  is  the  energy  value  in  the  middle  of  the 
cell,  and  for  a  rotator  the  energy  value  is 

Kn  =  n(n  +  IJg^j*  (3.9) 

approximately  the  mean  value  through  the  cell.  The  integer  n  is  called 
the  quantum  number.  The  distribution  of  points  in  a  quantum  state  of 
fixed  energy  is  independent  of  time,  and  for  that  reason  the  state  is  called 
a  stationary  state.  This  is  in  contrast  to  other  ways  of  sotting  up  cells. 
For  instance,  with  rectangular  cells,  we  find  in  general  that  tho  systems  in 
one  state  have  a  distribution  of  energies,  and  as  time  goes  on  systems  jump 
at  a  certain  rate  from  one  state  to  another,  having  what  are  called  quan- 
tum transitions,  so  that  the*  number  of  systems  in  each  state  changes  with 
time.  One  can  draw  a  certain  parallel,  or  correspondence,  between  the 
jumping  of  systems  from  one  quantum  state  to  another,  and  the  uniform 
flow  of  representative  points  in  the  phase  space  in  classical  mechanics. 
Suppose  we  have  a  classical  assembly  whose  density  in  the  phase  space 
changes  very  slowly  from  point  to  point,  changing  by  only  a  small  amount 
in  going  from  what  would  be  one  quantum  cell  to  another.  Then  we  can 
set  up  a  quantum  assembly,  the  fraction  of  systems  in  each  quantum  state 
being  given  by  tho  fraction  of  the  classical  systems  in  the  corresponding 
cell  of  phase  space.  And  the  time  rate  of  change  of  the  fraction  of  sys- 
tems in  each  quantum  state  will  be  given,  to  a  good  approximation,  by 
the  corresponding  classical  value.  This  correspondence  breaks  down, 
however,  as  soon  as  the  density  of  the  classical  assembly  changes  greatly 
from  cell  to  cell.  In  that  case,  if  we  set  up  a  quantum  assembly  as  before, 
we  shall  find  that  its  time  variation  does  not  agree  at  all  accurately  with 
what  we  should  get  by  use  of  our  classical  analogy. 

Antp.1  ftfomif  systems  pbev  thf  fljjQ-nfiim  *Vnry  n/rt  r»]flssif»fll 
mechanicsTso  that  we  shall  be  concerned  with  auaptum  statistics.  The 
only  cases  in  whinfi  we  can  use  classical  theory  as  an  app^oxinmf.inii  art* 
tHbse.in  which  the  density  in  phase  varies  only  a,  lif^lr  fr-nm  gfn.fn  ^  state, 
—the  case  we  have  mentioned  in  the  last  paragraph  As  a  matter  of  fact, 
as  we  shall  see  later,  this  corresponds  roughly  to  the  limit  of  high  tempera- 
ture. Thus,  we  shall  often  find  that  classical  results  are  correct  at  high 
temperatures  but  break  down  at  low  temperature.  \  f-v™nnil  jfflftTnnlf-  of 
this  is  the  theory  of  specific  heat;  we  shall  find  others  as  we  go  on.  We 
now  understand  the  qualitative  features  of  quantum  statistics  well  enough 
so  that  in  the  next  section  we  can  go  on  to  our  task  of  understanding  the 


SBC.  4]  STATISTICAL  MECHANICS  43 

nature  of  irreversible  processes  and  the  way  in  which  the  entropy  increases 
with  time  in  such  processes. 

4.  Irreversible  Processes.  —  We  shall  start  our  discussion  of  irreversi- 
ble processes  using  classical  mechanics  and  Liouville's  theorem.  Let  us 
try  to  form  a  picture  of  what  happens  when  we  start  with  a  system  out  of 
equilibrium,  with  constant  energy  and  volume,  follow  its  irreversible 
change  into  equilibrium,  and  examine  its  final  steady  state.  To  have  a 
specific  example,  consider  tho  approach  to  equilibrium  of  a  perfect  gas 
having  a  distribution  of  velocities  which  originally  does  not  correspond 
to  thermal  equilibrium.  Assume  that  at  the  start  of  an  experiment,  a 
mass  of  gas  is  rushing  in  one  direction  with  a  large  velocity,  us  if  it  had  just 
been  shot  into  a  container  from  a  jet.  This  is  far  from  an  equilibrium 
distribution.  The  random  kinetic  energy  of  the  molecules,  which  we 
should  interpret  as  heat  motion,  may  be  very  small  and  the  temperature 
low,  and  yet  they  have  a  lot  of  kinetic  energy  on  account  of  their  motion 
in  the  jet.  In  the  phase  space,  the  density  function  will  be  large  only  in 
the  very  restricted  region  where  all  molecules  have  almost  the  same 
velocity,  the  velocity  of  the  jot  (that  is,  the  equations 

Pi!  _  Prf   .  _   y     H(, 

-    —  ~      -  —      r   j  ,    I  I  C  .  . 

m  i        w2 

where  Vx  is  the  x  component  of  velocity  of  the  jot,  are  almost  satisfied  by 
all  points  in  the  assembly),  and  all  have  coordinates  near  tho  coordinate 
of  the  center  of  gravity  of  the  rushing  mass  of  gas  (that  is,  tho  equations 
Xi  =  x2  =  -  -  -  =  A',  where  X  is  the  x  coordinate  of  the  center  of  gravity 
of  the  gas,  are  also  approximately  satisfied).  We  see,  then,  that  the 

entropy,  as  defined  by  —  kf,  In/,,  will  be  small  under  these  conditions. 


But  as  time  goes  on,  the  distribution  \\ill  change.  The  jet  of  molecules 
will  strike  the  opposite  wall  of  tho  container,  and  after  bouncing  back 
and  forth  a  few  times,  will  become  moro  and  more  dissipated,  with  irregu- 
lar currents  and  turbulence  sotting  in.  At  first  we  shall  describe  those 
things  by  hydrodynamics  or  aerodynamics,  but  wo  shall  find  that  tho 
description  of  the  flow  gets  more  and  more  complicated  with  irregularities 
on  a  smaller  and  smaller  scale.  Finally,  with  the  molecules  colliding 
with  the  walls  and  with  each  other,  things  will  become  extremely  involved, 
some  molecules  being  slowed  down,  some  speeded  up,  the  directions 
changed,  so  that  instead  of  having  most  of  the  molecules  moving  with 
almost  the  same  velocity  and  located  at  almost  the  same  point  of  space, 
there  will  be  a  whole  distribution  of  momentum,  both  in  direction  and 
magnitude,  and  the  mass  will  cease  its  concentration  in  space  and  will  be 
uniformly  distributed  over  the  container.  There  will  now  be  a  great 


44  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 

many  points  of  phase  space  representing  states  of  the  system  which  could 
equally  well  be  this  final  state,  so  that  the  entropy  will  be  large.  And  the 
increase  of  entropy  has  come  about  at  the  stage  of  the  process  where  we 
cease  to  regard  the  complication  in  the  motion  as  large-scale  turbulence, 
and  begin  to  classify  it  as  randomness  on  a  microscopic  or  atomic  scale. 
Finally  the  gas  will  come  to  an  equilibrium  state,  in  which  it  no  longer 
changes  appreciably  with  time,  and  in  this  state  it  will  have  reached  its 
maximum  entropy  consistent  with  its  total  energy. 

This  qualitative  argument  shows  what  we  understand  by  an  irreversi- 
ble process  and  an  increase  of  entropy:  an  assembly,  originally  concen- 
trated in  phase  space,  changes  on  account  of  the  motion  of  the  system  in 
such  a  way  that  the  points  of  the  assembly  gradually  move  apart,  filling 
up  larger  and  larger  regions  of  phase  space.  This  is  likely,  for  there  are 
many  ways  in  which  it  can  happen ;  while  the  reverse  process,  a  concentra- 
tion of  points,  is  very  unlikely,  and  we  can  for  practical  purposes  say  that 
it  does  not  happen. 

The  statement  we  have  just  made  seems  at  first  to  be  directly  con- 
trary to  Liouville's  theorem,  for  we  have  just  said  that  points  originally 
concentrated  become  dispersed,  while  Liouville's  theorem  states  that  as 
we  follow  along  with  a  point,  tho  density  never  changes  at  all.  We  can 
give  an  example  used  by  Gibbs1  in  discussing  this  point.  Suppose  we 
have  a  bottle  of  fluid  consisting  of  two  different  liquids,  one  black  and 
one  white,  which  do  not  mix  with  each  other.  We  start  with  one  black 
drop  in  the  midst  of  the  white  liquid,  corresponding  to  our  concentrated 
assembly.  Now  we;  shake  or  stir  the  liquid.  Tho  black  drop  will  become 
shaken  into  smaller  drops,  or  be  drawn  out  into  thin  filaments,  which  will 
become  dispersed  through  the  white  liquid,  finally  forming  something  like 
an  emulsion.  Each  microscopic  black  drop  or  filament  is  as  black  as 
ever,  corresponding  to  the  fact  that  the  density  of  points  cannot  change  in 
the  assembly.  But  eventually  the  drops  will  become  small  enough  and 
uniformly  enough  dispersed  so  that  each  volume  element  within  the  bottle 
will  seem  uniformly  gray.  This  is  something  like  what  happens  in  the 
irreversible  mixing  of  the  points  of  an  assembly.  Just  as  a  droplet  of 
black  fluid  can  break  up  into  two  smaller  droplets,  its  parts  traveling  in 
different  directions,  so  it  can  happen  that  two  systems  represented  by 
adjacent  representative  points  can  separate  and  have  quite  different 
histories;  one  may  be  in  position  for  certain  molecules  to  collide,  while  the 
other  may  be  just  different  enough  so  that  these  molecules  do  not  collide 
at  all,  for  example.  Such  chance  events  will  result  in  very  different, 
detailed  histories  for  the  various  systems  of  an  assembly,  even  if  the 
original  systems  of  the  assembly  were  quite  similar.  That  is,  they  will 

!J.  W.  Gibbs,  "Elementary  Principles  in  Statistical  Mechanics,"  Chap.  XII, 
Longmans,  Green  &  Company. 


SEC.  4]  STATISTICAL  MECHANICS  4/'» 

result  in  representative  points  which  were  originally  close  together  in 
phase  space  moving  far  apart  from  each  other. 

From  the  example  and  the  analogy  we  have  used,  we  see  that  in  an 
irreversible  process  the  points  of  the  original  compact  and  orderly  assem- 
bly gradually  get  dissipated  and  mixed  up,  with  consequent  increase  of 
entropy.  Now  let  us  see  how  the  situation  is  affected  when  we  consider 
the  quantum  theory  and  the  finite  size  of  cells  in  phase  space.  Our 
description  of  the  process  will  depend  a  good  deal  on  the  scale  of  the 
mixing  involved  in  the  irreversible  process.  So  long  as  the  mixing  is  on  a 
large  scale,  by  Liouville's  theorem,  the  points  that  originally  wore  in  one 
roll  will  simply  be  moved  bodily  to  another  cell,  so  that  the  contribution 
of  these  points  to  —  fc2/t  ln/t  will  be  the  same  as  in  the  original  distribu- 
tion, and  the  entropy  will  be  unchanged.  The  situation  is  very  different, 
however,  when  the  distribution  as  we  should  describe  it  by  classical 
mechanics  involves  a  set  of  filaments,  of  different  densities,  on  a  scale 
small  compared  to  a  cell.  Then  the  quantum /t,  rather  than  equaling  the 
classical  value,  will  be  more  nearly  the  average  of  the  classical  values 
through  the  cell,  leading  to  an  increase  of  entropy,  at  the  samo  time  that 
the  average  or  quantum  density  begins  to  disobey  Liouville's  theorem. 

It  is  at  this  same  stage  of  the  process  that  it  becomes  really  impossible 
to  reverse  the  motion.  It  is  a  well-known  result  of  Newton's  laws  that 
if,  at  a  given  instant,  all  the  positions  of  all  particles  are  left  unchanged, 
but  all  velocities  are  reversed  in  direction,  the  whole  motion  will  reverse1, 
and  go  back  over  its  past  history.  Thus  every  motion  is,  in  theory,  rever- 
sible. What  is  it  that  in  practice  makes  some  motions  reversible,  others 
irreversible?  It  is  simply  the  practicability  of  setting  up  the  system  with 
reversed  velocities.  If  the  distribution  of  velocities  is  on  a  scale  large 
enough  to  see  and  work  with,  there  is  nothing  making  a  reversal  of  the 
velocities  particularly  hard  to  set  up.  With  our  gas,  we  could  suddenly 
interpose  perfectly  reflecting  surfaces  normal  to  the  various  parts  of  the 
jet  of  gas,  reversing  the  velocities  on  collision,  or  could  adopt  some  such 
device.  But  if  the  distribution  of  velocities  is  on  too  small  n  scale  to  see 
and  work  with,  we  have  no  hope  of  reversing  the  velocities  experimentally. 
Considering  our  emulsion  of  black  and  white  fluid,  which  we  have  pro- 
duced by  shaking,  there,  is  no  mechanical  reason  why  the  fluid  could  not 
be  unshaken,  by  exactly  reversing  all  the  motions  that  occurred  in  shaking 
it.  But  nobody  would  be  advised  to  try  the  experiment. 

It  used  to  be  considered  possible  to  imagine  a  being  of  finer  and  more 
detailed  powers  of  observation  than  ours,  who  could  regulate  systems  on  a 
smaller  scale  than  we  could.  Such  a  being  could  reverse  processes  that 
we  could  not;  to  him,  the  definition  of  a  reversible  process  would  be  differ- 
ent from  what  it  is  to  us.  Such  a  being  was  discussed  by  Maxwell  and  is 
often  called  "Maxwell's  Demon."  Is  it  possible,  we  may  well  ask,  to 


46  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 

imagine  demons  of  any  desired  degree  of  refinement?  If  it  is,  we  can 
make  any  arbitrary  process  reversible,  keep  its  entropy  from  increasing, 
and  the  second  law  of  thermodynamics  will  cease  to  have  any  signifi- 
cance. The  answer  to  this  question  given  by  the  quantum  theory  is 
No.  An  improvement  in  technique  can  carry  us  only  a  certain  distance, 
a  distance  practically  reached  in  plenty  of  modern  experiments  with 
single  atoms  and  electrons,  and  no  conceivable  demon,  operating  accord- 
ing to  the  laws  of  nature,  could  carry  us  further.  The  quantum  theory 
gives  us  a  fundamental  size  of  cell  in  the  phase  space,  such  that  we  cannot 
regulate  the  initial  conditions  of  an  assembly  on  any  smaller  scale.  And 
this  fundamental  cell  furnishes  us  with  a  unique  way  of  defining  entropy 
and  of  judging  whether  a  given  process  is  reversible  or  irreversible. 

5.  The  Canonical  Assembly. — In  the  preceding  section,  we  have* 
shown  that  our  entropy,  as_de_fined  in  Kg.  (1.2),  has  one  f}f  fdlfi  p^p^ti^ 
of  the  physical  entropy:  it  increases  in  an  irreversible  process T  for  it 
incrcascTwhonovor  the  assembly  becomes  ftffused  or  sc^ttero^  and  this 
happens  in  irreversible  processes.  We  must  next  take  up  thermal  eauilib- 
ium.  finding  tirst  the  correct  assembly  to  describe  the  fiensity  fungtion  in 
thermal  equilibrium,  and  then  proving  from  this  fonaitv  fnnct.ionr  thnf, 
our  entropy  satisfies  the  condition  dS  =  dO IT  for  a  reversible,  proress. 
From  Liouville's  foeorem,  we  have  one  piece  of  information  aboqt  the, 
assembly:  in  or^ler  that  it  ffiav  be  independent  of  time,  the  quantity /, 
must^be  a  function  only  of  the  energy  of  the  system.  We^et  A\  be  the 
energy  of  a  system  in  the  ii\\  ccll?  choosing  for  this  purpose  the;  type  of 
quantum  cells  representing  stationary  states  oy  ^"p-rp[y  H™lg 
wish  to  haye/t  a  function  of  J?,,  but  we  do  not,  vot,  xr  how-to 
this r  function. 

The  essential  method  which  we  use  is  the  following:  We  have  seen  that 
in  an  irreversible  process,  the  entropy  tends  to  increase  to  a  maximum,  for 
an  assembly  of  isolated  systems.  If  all  systems  of  the  assembly  have  the 
same  energy,  then  the  only  cells  of  phase  space  to  which  systems  can 
travel  in  the  course  of  the  irreversible  process  are  cells  of  this  same  energy, 
— a  finite  number.  The  distribution  of  largest  entropy  in  such  a  case,  as 
we  have  seen  in  Sec.  1,  is  that  in  which  systems  are  distributed  with 
uniform  density  through  all  the  available  cells.  This  assembly  is  called 
the  microcanonical  assembly,  and  it  satisfies  our  condition  that  the 
density  be  a  function  of  the  energy  only:  all  the  /t's  of  the  particular 
energy  represented  in  the  assembly  are  equal,  and  all  other  /t's  are  zero. 
But  it  is  too  specialized  for  our  purposes.  For  thermal  equilibrium,  we 
do  not  demand  that  the  energy  be  pr^p|y  "dp.tenninpd  WP  dpnmnd 
rather  that  the  tempcrnturfl  nf  ^1]  yretems  of  th*  a^mblv  be  flie  same. 
This  can  be  interpreted  most  properly  in  the  .following  way.  We  allow 
eacfr  system  of  the  assembly  to  be  in  contact  with  a  temperature  bath  of 


SBC.  51  STATISTICAL  MECHANICS  4? 

the  required  temperature,  a  body  of  very  large  heat  capacity  held  at  the 
desired  temperature.  The  systems  of  the  assembly  are  then  not  isolated. 
Rather,  they  can  change  their  energy  by  interaction  with  the  temperature 
bath.  Thus,  even  if  we  started  out  with  an  assembly  of  systems  all  of  the 
same  energy,  some  would  have  their  energies  increased,  some  decreased, 
by  interaction  with  the  bath,  and  the  final  stable  assembly  would  havo  a 
whole  distribution  of  energies.  There  would  certainly  be  a  definite  aver- 
age energy  of  the  assembly,  however;  with  a  bath  of  a  given  temperature, 
it  is  obvious  that  systems  of  abnormally  low  energy  will  tend  to  gain 
energy,  those  of  abnormally  high  energy  to  lose  energy,  by  the  interaction. 
To  find  the  final  equilibrium  statcf  then,  wo  may  ask  this  question:  what 
is  the  assembly  of  systems  wrhich  has  the  maximum  entropy,  subject  onT7 
to  the  condition  that  its  moan  energy  havo  a  given  value?..  It  seems  most 
reasonable  that  this  will  he  the  assembly  which  will..  bo  tho  final  result,  of 
the  irreversible  contact  of  any  group  of  systems  with  a  largo  temperature 


The1  assembly  that  results  from  these  conditions  is  called  tho  canonical 
assembly!     Lot  us  formulate  tho  conditions  which  it  must  satisfy.     It 

must  bo  the  assembly  for  which  S  =  ~k2\ft  In  /t  is  a  maximum,  subject 


(o  a  constant  mean  energy.  But  we  can  find  the  mean  energy  immedi- 
ately in  terms  of  our  distribution  function/,.  In  tho  ?th  coll,  u  system  has 
energy  E%.  The  fraction  fv  of  all  systems  will  be  found  in  this  coll.  Hence 
the  weighted  mean  of  the  energies  of  all  systems  is 


\ 


(5.1) 


This  quantity  must  be  held  constant  in  varying  the  /t's.     Also,  as  we  saw- 
in  Eq.  (1.1),  the  quantity  ]^/t  equals  unity.     This  must  always  be  satis- 


fled,  no  matter  how  the  /t's  vary.     We  can  restate  the  conditions,  by 
finding  dS  and  dU:  we  must  have 


,  (In/.  +  1),  (5.2) 

for  all  sets  of  d//s  for  which  simultaneously 

dU  =  0  =  2///,A\,  (5.3) 

» 

and 


48  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 

On  account  of  Eq.  (5.4),  we  can  rewrite  Eq.  (5.2)  in  the  form 


dS  =  0  -  -kdfi  ln/i.  (5.5) 

i 

The  set  of  simultaneous  equations  (5.3),  (5.4),  (5.5)  can  be  handled  by  the 
method  called  undetermined  multipliers:  the  most  general  value  which 
In/*  can  have,  in  order  that  dS  should  be  zero  for  any  set  of  d/Vs  for  which 
Eqs.  (5.3)  and  (5.4)  are  satisfied,  is  a  linear  combination  of  the  coeffi- 
cients of  dfi  in  Eqs.  (5.3)  and  (5.4),  with  arbitrary  coefficients: 

In  fi  =  a  +  bE%.  (5.6) 

For  if  Eq.  (5.6)  is  satisfied,  Eq.  (5.5)  becomes 


=  -ka^dfi  -  kb^dfrft,  ,  (5.7) 

%  % 

which  is  zero  for  any  values  of  cZ/t  for  which  Eqs.  (5.3)  and  (5.4)  are  satis- 
fied. 

The  values  of  /t  for  the  canonical  assembly  are  determined  by  Eq. 
(5.6).     It  may  be  rewritten 


/,  =  fe6*..  (5.8) 

Clearly  b  must  bo  negative;  for  ordinary  systems  have  possible  states  of 
infinite  energy,  though  not  of  negatively  infinite  energy,  and  if  b  wore 
positive,  fi  would  become  infinite  for  the  states  of  infinite  energy,  an 
impossible  situation.  We  may  easily  evaluate  the  constant  a  in  torms  of 

6,  from  the  condition  5/»  =  1.     This  gives  at  once 


so  that 

J>K. 

(5.9) 


If  the  assembly  (5.9)  represents  thermal  equilibrium,  the  change  of 
entropy  when  a  certain  amount  of  heat  is  absorbed  in  a  reversible  process 


SEC.  5]  STATISTICAL  MECHANICS  49 

should  be  dQ/T.     The  change  of  entropy  in  any  process  in  thermal 
equilibrium,  by  Eqs.  (5.5)  and  (5.9),  is 

dS  =  -k^df>  In  /,  =  -k^dMbEt  -  In 

i  l  J 

(5.10) 


using  Eq.  (5.4).     Now  consider  the  change  of  internal  energy.     This  is 

E^  (5.11) 


The  first  term  in  Ea.  (5.11^  arises  when  the  external  forces  stay  constant, 
resulting  in  constant  values  of  Et)  but  there  is  a  change  m  tne  assembly, 
meaning  a  shift  of  molecules  from  one  position  and  velocity  to  another. 
This  change  of  course  is  different  from  that  considered  in  Eq.  (5.3),  for 
that  referred  to  an  irreversible  approach  to  equilibrium,  while  this  refers, 
to  a  change  from  one  equilibrium  state  to  another  of  different  energy.! 
Such  a  change  of  molecules  on  a  molfif.nlfl.r  sp.fl.lft  is  t.q  \}&  interpreted  as  ail 
absorption  of  heat.  The  second  term,  however,  comes  about  when  the 
/Vs  and  the  entropy  do  not  change,  but  the  energies  of  the  cells  themselves 
change,  on  account  of  changes  in  external  forces  and  in  the  potential 
energy.  This  is  to  be  interpreted  as  external  work  donft  on  thn  s 
the  negative  of  the  work  done  by  the  system.  Thus  we^have 


I 


dQ  =       Etdf          dW  =  - 


(5.12) 


Combining  Eq.  (5.12)  with  Ea.  (5.11^  yivns  us  the  first  law, 

dU  =  dO  -  dW. 
Combining  with  Eq.  (5.10)T  we  have 

dS  =  -kbdQ.^  (5.13) 

Equation  (5.13),  stating  the  proportionality  of  dS  and  dQ  for  a  reversible 
process,  is  a  statement  of  the  second  law  of  thermodynamics  for  a  reversi- 
ble process,  if  we  have 


-  (5.14) 


Using  Eq.  (5.14),  we  can  identify  the  constants  in  Eq.  (5.9),  obtain- 
ing as  the  representation  of  the  canonical  assembly 

~"—  •  (5.15) 


50  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  Ill 

It  is  now  interesting  to  compute  the  Helmholtz  free  energy  A  =  (7  —  TS. 
This,  using  Eqs.  (5.1)  and  (5.15),  is 


A  =  2/t(  #t  -  E\  -  kT  In 
t      \ 

=  -ArTMn  JV**,    -  (5.16) 


or 

^  t 

(5.17) 


j» 


Using  Eq.  (5.17),  we  may  rewrite  the  formula  (5.15)  as 


I 


=  6    *y   •  (5.18) 


The  result  of  Eq.  (5.17)  is,  for  practical  purposes,  the  most  importan  t 
result  of  statistical  mechanics.  For  it  gives  a  perfectly  direct  and 
strajghtforwftff]  way  nf  (jpfiving  the  Helmholtz  fyee  energy,  find  hence 
the  equation  of  state  and  specific  heat,  of  any  system,  if  we  know  its 
energy  as  a  function  of  coordinates  and  Hininmtft  —  The  sum  of  Eq. 
(5.17),  which  we  have  denoted  bv  Z.  is  often  called  the  partition  function. 

Oftnn  it  fc  flfiflfill  ffft  fro  flh^  **^a™*  l^o  pn|rnnv[  ^ptP.i^^pnP.rpv,   and 

specific  heat  directly  from  the  partition  function,  without  separately 
computing  the  Helmholtz  free  energy.     For  the  entropjt,  using 


we  have 

S  =  \ — (kT  In  Z)  /    =•  k  In  Z  H ( —  I  (5.19) 

For  the  internal  energy,  U  =  A  +  TS,  we  have 

U  =  -*r  In  Z  +  kT]nZ  +  ^-fe 


(5.20) 
where  the  last  form  is  often  useful.     For  the  specific  heat  at  constant 


SBC.  5|  STATISTICAL  MECHANICS  51 

volume,  we  may  use  either  CV  =  (dV/dT)v  or  Cv  =  T(dS/dT)r.     From 
the  latter,  we  have 

Jd*(kTluZ)\ 
Cv  =  1—-      -- 


. 

(5.21) 

We  have  stated  our  definitions  of  entropy,  partition  function,  and 
other  quantities  entirely  in  terms  of  summations.  Often,  however,  the 
quantity  /t  changes  only  slowly  from  cell  to  cell;  in  this  case  it  is  con- 
venient to  replace  the  summations  bv  integrations  over  the  ohaso  space. 
We  recall  that  all  cells  are  of  the  same  volume,  An,  if  there  are  n  coordi- 
nates and  n  momenta  in  the  phase  space.  Thus  the  number  of  cells  in  a 
volume  element  dcji  .  .  .  da*.  dv\  .  .  .  dvn  of  phase  space  is 


Then  the  partition  function  becomes — 

e    kT  dqi  .  .  .  dpn,      \  (5.22) 


\ 


a  very  convenient  form  for  such  problems  as  finding  the  partition  function 
of  a  perfect  gas. 


CHAPTER  IV 
THE  MAXWELL-BOLTZMANN  DISTRIBUTION  LAW 

In  most  physical  applications  of  statistical  mechanics,  we  deal  with  a 
system  composed  of  a  great  number  of  identical  atoms  or  molecules,  and 
are  interested  in  the  distribution  of  energy  between  these  molecules.  The 
simplest  case,  which  we  shall  take  up  in  this  chapter,  is  that  of  the  perfect 
gas,  in  which  the  molecules  exert  no  forces  on  each  other.  We  shall  be 
led  to  the  Maxwell-Boltzmann  distribution  law,  and  later  to  the  two  forms 
of  quantum  statistics  of  perfect  gases,  the  Fermi-Dirac  and  Einstein-Bose 
statistics. 

1.  The  Canonical  Assembly  and  the  Maxwell-Boltzmann  Distribu- 
tion.— Let  us  assume  a  gas  of  N  identical  molecules,  and  let  each  molecule 
have  n  degrees  of  freedom.  That  is,  n  quantities  are  necessary  to  specify 
the  configuration  completely.  Ordinarily,  three  coordinates  are  needed 
to  locate  each  atom  of  the  molecule,  so  that  n  is  three  times  the;  number 
of  atoms  in  a  molecule.  In  all,  then,  Nn  coordinates  are  necessary  to 
describe  the  system,  so  that  the  classical  phase  space  has  2Nn  dimen- 
sions. It  is  convenient  to  think  of  this  phase  space  as  consisting  of  N 
subspaces  each  of  2n  dimensions,  a  subspace  giving  just  the  variables 
required  to  describe  a  particular  molecule  completely.  Using  the  quan- 
tum theory,  each  subspace  can  be  divided  into  cells  of  volume  hn,  and  a 
state  of  the  whole  system  is  described  by  specifying  which  quantum  state 
each  molecule  is  in,  in  its  own  subspace.  The  energy  of  the  whole  system 
is  the  sum  of  the  energies  of  the  N  molecules,  since  for  a  perfect  gas  there 
are  no  forces  between  molecules,  or  terms  in  the  energy  depending  on 
more  than  a  single  molecule.  Thus  we  have 

N 
#=2<(f),  (1.1) 


where  6(t)  is  the  energy  of  the  zth  molecule.  If  the  ith  molecule  is  in  the 
Atth  cell  of  its  subspace,  let  its  energy  be  €j£.  Then  we  can  describe 
the  energy  of  the  whole  system  by  the  set  of  kjs.  Now,  in  the  canonical 
assembly,  the  fraction  of  all  systems  for  which  each  particular  molecule,  as 
the  fth,  is  in  a  particular  state,  as  the  fctth,  is 

52 


SBC.  1]         THE  MAXWELL-BOLTZMANN  DISTRIBUTION  LAW  53 


kl      kN 

,—t^ki/kT 

(1.2) 


ki  kN 

It  is  now  interesting  to  find  the  fraction  of  all  systems  in  which  a  particu- 
lar molecule,  say  the  ith,  is  in  the  fctth  state,  independent  of  what  other 
molecules  may  be  doing.  To  find  this,  we  merely  sum  the  quantity  (1.2) 
over  all  possible  values  of  the  fc's  of  other  molecules.  The  numerator  of 
each  separate  fraction  in  Eq.  (1.2),  when  summed,  will  then  equal  the 
denominator  and  will  cancel,  leaving  only 

(1.3) 


as  the  fraction  of  all  systems  of  the  assembly  in  which  the  zth  molecule  is 
in  the  /btth  state,  or  as  the  probability  of  finding  the  ith  molecule  in  the 
A;tth  state.  Since  all  molecules  are  alike,  we  may  drop  the  subscript  i 
in  Eq.  (1.3),  saying  that  the  probability  of  finding  any  particular  molecule 
in  the  kih  state  is 


2e~Ff 


II 


(1.4) 


Equation  ^1.4)  expresses  what  is  called  the  Maxwell-Boltzmann  distribu- 
tion law.  If  Eq.  (1.4)  gives  the  probability  of  finding  any  particular 
molecule  in  the  fcth  state,  it  is  clear  that  it  also  gives  the  fraction  of  ^11 
molecules  to  be  found  in  that  state,  averaged  through  the  assembly. 

The  Maxwell-fioltzmann  distribution  law  can  be  used  for  many 
calculations  regarding  gases ;  in  a  later  chapter  we  shall  take  up  its  appli- 
cation to  the  rotational  and  vibrational  levels  of  molecules.  For  the 
present,  we  shall  describe  only  its  use  for  monatomic  gases,  in  which  there 
is  only  translational  energy  of  the  molecules,  no  rotation  or  vibration.  In 
this  case,  as  we  shall  show  in  the  next  paragraph,  the  energy  levels  «*  of  a 
single  molecule  are  so  closely  spaced  that  we  can  regard  them  as  continu- 
ous and  can  replace  our  summations  by  integrals.  We  shall  have  three 
coordinates  of  space  describing  the  position  of  the  molecule,  and  three 
momenta,  px,  py,  pt,  equal  to  the  mass  ra  times  the  components  of  velocity, 
vx,  vv,  vg.  The  energy  will  be  the  sum  of  the  kinetic  energy,  %mv2  =  p */2m, 


54  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IV 

and  the  potential  energy,  which  we  shall  denote  by  <£(z,  y,  z)  and  which 
may  come  from  external  gravitational,  electrostatic,  or  other  types  of 
force  field.  Then  the  fraction  of  molecules  in  the  range  of  coordinates 
and  momenta  dx  dy  dz  dpx  dpv  dp,  will  be 


_?_        kT     ^dx  dy  dz  dpx  dpy  dpz 

'"  (pV2m+*)  *  '  U-O; 

J7JJ  J  J  e        kT       dx  dy  dz  dpx  dpv  dpz 

In  the  next  section  we  shall  derive  some  simple  consequences  from  this 
form  of  the  Maxwcll-Boltzmann  distribution  for  perfect  moriatomic 
gases. 

In  the  last  paragraph  we  have  used  the  fact  that  the  translational 
energy  levels  of  a  perfect  monatomic  gas  are  very  closely  spaced,  accord- 
ing to  the  quantum  theory.  We  can  sec  this  as  follows,  limiting  ourselves 
to  a  gas  in  the  absence  of  an  external  force  field.  Each  molecule  will  have 
a  six-dimensional  phase  space.  Consider  one  pair  of  variables,  as  x  and 
px.  Since  no  forces  act  on  a  molecule,  the  momentum  px  stays  constant 
during  its  motion,  which  must  take  placo  in  a  range  X  along  the  x  axis, 
if  the  gas  is  confined  to  a  box  of  sides  X,  F,  X  along  the  three  coordinates. 
Thus  the  area  enclosed  by  the  path  of  the  particle  in  the  x  —  px  section  of 
phase  space  is  pxX,  which  must  be  equal  to  an  integer  nx  times  h.  Then 
we  have 

nxh  nvh  nji  ,,  „, 

~ 


where  the  n's  are  integers,  which  in  this  case  can  be  positive  or  negative 
(since  momenta  can  be  positive  or  negative).  The  energy  of  a  molecule 
is  then 


.- 

2m  ~  2m\X*       F2      Z* 

To  get  an  idea  of  the  spacing  of  energy  levels,  let  us  see  how  many  levels 
are  found  below  a  given  energy  e.  We  may  set  up  a  momentum  space,  in 
which  px,  pvi  pz  are  plotted  as  variables.  Then  Eq.  (1.6)  states  that  a 
lattice  of  points  can  be  set  up  in  this  space,  one  to  a  volume 


where  V  =  XYZ  is  the  volume  of  the  container,  each  point  corresponding 
to  an  energy  level.     The  equation 


2m 


SEC.  2]         THE  M  AX  W  ELL-BOLT  Z  MANN  DISTRIBUTION  LAW  55 

is  the  equation  of  a  sphere  of  radius  p  =  \/2we  in  this  space,  and  the 
number  of  states  with  energy  less  than  6  equals  the  number  of  points 
within  the  sphere,  or  its  volume,  ^7r(2mc)^,  times  the  number  of  points 
per  unit  volume,  or  V/h*.  Thus  the  number  of  states  with  energy  less 
than  e  is 


,  (1.9) 

and  the  number  of  states  between  e  and  e  +  rfe,  differentiating,  is 

(/c.  (1.10) 


The  average  energy  between  successive  states  is  the  reciprocal  of  Eq. 
(1.10),  or 

1 


Let  us  see  what  this  is  numerically,  in  a  reasonable  case.  We  take  a 
helium  atom,  with  mass  6.63  X  10~24  gm.,  in  a  volume  of  1  cc.,  with  an 
energy  of  k  =  1.379  X  10~~16  erg,  which  it  would  have  at  a  fraction  of  a 
degree  absolute.  Using  h  =  6.61  X  10~27,  this  gives  for  the  energy 
difference  between  successive  states  the  quantity  8. 1  X  10~38  erg,  a  com- 
pletely negligible  energy  difference.  Thus  we  have  justified  our  state- 
ment that  the  energy  levels  for  translational  energy  of  a  perfect  gas  are  so 
closely  spaced  as  to  be  essentially  continuous. 

2.  Maxwell's  Distribution  of  Velocities. — Returning  to  our  distribu- 
tion law  (1.5),  let  us  first  consider  the  case  where  there  is  no  potential 
energy,  so  that  the  distribution  is  independent  of  position  in  space.  Then 
the  fraction  of  all  molecules  for  which  the  momenta  lie  in  dpx  dpv  dpx  is 

e   2mhTdpx  dpy  dp,  (e>  -  v 

_ ,  \^'  * ) 

fffe~*"*Tdpxdpvdpz 

where  p2  stands  for  pi  +'  p2  +  p2.  The;  integral  in  the  denominator  can 
be  factored,  and  written  in  the  form 

/•oo  P'*  /•«       ?"*-  /•  QO  P«* 

I      g   2mkT  ^px  i      g   %mkT  dpyl      e   %mkT  dpg.  (2.2) 

a/  •• •  oo  J  —  w  •/  ™  oo 

/OO  1 

e~au*  du,  where  a  =       *„,•     We  shall 

meet  many  integrals  of  this  type  before  we  are  through,  and  we  may  as 


56                        INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IV 

well  give  the  formulas  for  them  here.  We  have 

I     une~QH*  du  = 
Jo 

Wa  f°r  n  =  °  5?  f°r  W  =  * 

1  [?r '   =  2  L for ;/  s  3 

=  4  4-}  for  ?i  =  5,  etc.             (2.3) 


3    It 
sVo1 


Starting  with  fe~au*  du,  for  the  even  powers  of  n,  or  with  fue~au*  du  for  the 
odd  powers,  each  integral  can  be  found  from  the  one  above  it  by  differen- 
tiation with  respect  to  —  a,  by  means  of  which  the  table  can  be  extended. 
To  get  the  integral  from  —  oo  to  <x> ,  the  result  with  the  even  powers  is 
twice  the  integral  from  0  to  oo  t  and  for  the  odd  powers  of  course  it  is 
zero. 

Using  the  integrals  (2.3),  the  quantity  (2.2)  becomes  (2wmkT)^'. 
Thus  we  may  rewrite  Eq.  (2.1)  as  follows:  the  fraction  of  molecules  with 
momentum  in  the  range  dpx  dpy  dpg  is 


(ZirmkT)       e   2mkT  dpt  dpu  dp,. 


(2.4) 


Equation   (2.4)   is   one  form   of   the  famous   Maxwell  distribution   of 
velocities. 

Often  it  is  useful  to  know,  not  the  fraction  of  molecules  whose  vector 
velocity  is  within  certain  limits,  but  the  fraction  for  which  the  magnitude 
of  the  velocity  is  within  certain  limits.  Thus  let  v  be  the  magnitude  of 
the  velocity: 


m 

Then  we  may  ask,  what  fraction  of  the  molecules  have  a  speed  between 
i)  and  v  +  dv,  independent  of  direction?  To  answer  this  question  we 
consider  the  distribution  of  points  in  momentum  space.  The  volume  of 
momentum  space  corresponding  to  velocities  between  v  and  v  +  dv  is 
the  volume  of  a  spherical  shell,  of  radii  mv  and  rn(v  +  dv).  Thus  it  is 
4w(mvy  d(mv).  We  must  substitute  this  volume  for  the  volume  dps  dpv 
dpz  of  Eq.  (2.4).  Then  we  find  that  the  fraction  of  molecules  for  which 
the  magnitude  of  the  velocity  is  between  v  and  v  +  dv,  is 


This  is  the  more  familiar  form  of  Maxwell's  distribution  law.     We  give  a 


SEC.  2]         THE  M AXW ELL-BOLT Z MANN  DISTRIBUTION  LAW 


57 


graph  of  the  function  (2.6)  in  Fig.  IV-1.  On  account  of  the  factor  y'2,  tho 
function  is  zero  for  zero  speed;  and  on  account  of  the  exponential  it  is 
zero  for  infinite  speed.  In  between,  there  is  a  maximum,  which  is  easily 
found  by  differentiation  and  comes  at  v  —  \/2kT/m.  That  is,  the  maxi- 
mum, and  in  fact  the  whole  curve,  shifts  outward  to  larger  velocities  as 
the  temperature  increases. 

From  Maxwell's  distribution  of  velocities,  either  in  the  form  (2.4)  or 
(2.6),  we  can  easily  find  the  moan  kinetic  energy  of  a  molecule  at  tem- 
perature T.  To  find  this,  we  multiply  the  kinetic  energy  p2/2m  by  tho 
fraction  of  molecules  in  a  given  range  dpx  dpy  dpg,  and  integrate  over  all 
values  of  momenta,  to  get  the  weighted  mean.  Thus  we  have 

-54  C  C  Cv* — 

Mean  kinetic  energy  =  (2irmkT)        III  s—e   2mkT  dp*  dpu  dpz 

-  v  (  C  "°      2   -  PT* 

\  £—e~2"'^'dpx(2TrmkT) 

(J-  »  2m 

+  similar  terms  in  p*,  pi  / 

*($kT)(2TrmkT)** 
=  tfcZ7.  (2.7) 

The  formula  (2.7)  for  the  kinetic  energy  of  a  molecule  of  a  perfect  gas 
leads  to  a  result  called  the  equipartition  of  onorgy.  Each  molecule  has 
three  coordinates,  or  three  degrees  of 
freedom.  On  the  average,  each  of 
those  will  have  one-third  of  the  total 
kinetic  energy,  as  we  can  see  if  we  find 
the  average,  not  of  (pi  +  p%  +  pi) /2m, 
but  of  the  part  pl/2m  associated  with 
the  x  coordinate.  Thus  eqg|i  of  thcao 
degrees  of  frccdom^haa  nn  the  average* 
tJ^gj^gjgyA-fcT7.  Thy  energy,  in  other 
words,  is  equally  distributed  between 
each  one  having  the 


1.0 


2.0 


and  this  is  called 


the  eauipartition  of  energy.     Mathe-  .,      ...  ,     *,v  ^7  *.     »    . 

.   i      "                    Yft        ..    n^            -—  FIG.   IV-1. —Maxwell's  distribution  of 

matlCally,    as   an   examination   Of   Our  velocities,  giving  the  fraction  of  molecules 

proof  shows    ftniiinfl.rtit.inn   is  fl.  result  whose  velocity  is  between  v  and  v  +  dv,  in 

«— — • •••iM^^rftaBrfbBi^ « fc^Mfc^M— ii—M^  a  gas  a^  temperature  T. 
of  th(^fact^thaJ^th£1J^BJ|2^^ 

with  each  degree  of  freedom  is  nronprtional  to  the  square 


of 


^rro< 


rifii 


or  coordinate. 


which  is  found  in  the  energy  only  as  a  square,  will  be  found  to  have  a 
mean   energy   of   ^feTy  provided   the   energy  levels   are   continuously 


58  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IV 


that  anmmg.f.inng  fian  be  replaced  bv  integrations.  This 
applies  not  only  to  the  momenta  associated  with  translation  of  a  single 
atom,  but  to  such  quantities  as  the  angular  momentum  of  a  rotating 
molecule  (whose  kinetic  energy  is  p2/2/),  if  the  rotational  levels  are  spaced 
closely.  It  applies  as  well  to  the  coordinate,  in  the  case  of  a  linear  oscilla- 
tor, whose  restoring  force  is  —  kx,  and  potential  energy  fcx2/2;  the  mean 
potential  energy  of  such  an  oscillator  is  ffiT,  if  the  levels  are  spaced  closely 
enough,  though  in  many  physical  cases  which  wo  shall  meet  the  spacing  is 

not   close  enough  to  replace  summations  by 
integrations. 

3.  The  Equation  of  State  and  Specific  Heat 
of  Perfect  Monatomic  Gases.  —  Having  found 
the  distribution  of  molecular  velocities,  we  can 
calculate   the   equation   of   state   and   specific 
heat  of  perfect  monatomic  gases  by  elementary 
methods,  postponing  until  a  later  chapter  the 
iv-2.  -Diagram  to  direct  statistical  calculation  by  means  of  the 
illustrate  collisions  of  mole-   partition  function.     We  shall  again  limit  our- 

cules  with  1  sq.  cm.  of  surface    0_i  *       4.1  •    i  L  ,  i 

Of  waji.  selves  to  the  special  case  where  there  is  no 

potential    energy   and  where  the  distribution 

function  is  independent  of  position.  This  is  the  only  ease  where  we 
should  expect  the  pressure  to  be  constant  throughout  the  container.  We 
shall  find  the  pressure  by  calculating  the  momentum  carried  to  the  wall 
per  second  by  molecules  colliding  with  it.  The  momentum  transferred 
to  1  sq.  cm.  of  wall  per  second  is  the  force  acting  on  the  wall,  or  the 
pressure. 

For  convenience,  let  us  choose  the  x  axis  perpendicular  to  the  square 
centimeter  of  wall  considered,  as  in  Fig.  IV-2.  A  molecule  of  velocity 
v  =  p/m  close  to  the  wall  will  strike  the  wall  if  px  is  positive,  and  will 
transfer  its  momentum  to  the  wall.  When  it  is  reflected,  it  will  in  general 
have  a  different  momentum  from  what  it  6ngmally~had,  and  will  come 
away  with  momentum  p',  the  component  p'x  being  negativeT^-STEer 
collision,  in  other  words,  it  will  again  belong  to  the  group  ofmolecules 
near  the  wall,  but  now  corresponding  to  negative  px,  and  it  wilFhave 
taken  away  from  the  wall  the  momentum  p',  or  will"  have  given  the  wall 
the  negative  of  this  momentum.  We  can,  then,  get  all  the  momentum 
transferred  to  the  wall  by  considering  all  molecules,  both  with  positive 
and  negative  px's.  Consider  those  molecules  contained  in  "the  element 
dpxdpydpz  in  momentum  space,  and  lying  in  tHe  prism  drawn  in  Fig. 
IV-2.  Each  of  these  molecules,  and  no  others,  will  strike  the  square 
centimeter  of  wall  in  time  dt.  The  volume  of  the  prism  is  px/m  dt.  The 
average  number  of  molecules  per  unit  volume  in  the  momentum  element 
dpx  dpv  dpzj  averaged  through  the  assembly,  will  be  denoted  by  fm  dpxdpv 


SBC.  3]        THE  M AXW ELL-BOLT Z MANN  DISTRIBUTION  LAW  59 

dp,.     For  Maxwell's  distribution  law,  we  have 

fm  =  y(2irrofcrr  V5^  (3.1) 

using  Eq.  (2.4),  where  N/V  is  the  number  of  molecules  of  all  velocities 
per  unit  volume.  We  shall  not  explicitly  use  this  form  for  fm  at  the 
moment,  however,  for  our  derivation  is  more  general  than  Maxwell's 
distribution  law,  and  holds  as  well  for  the  Fermi-Dirac  and  Einstcin-Bose 
distributions,  which  we  shall  take  up  in  later  sections.  Using  the  func- 
tion fm,  the  average  number  of  molecules  of  the  desired  momentum,  in 
the  prism,  is  p^/m/m  dpx  dpi,  dp,.  Each  such  molecule  takes  momentum 
of  components  px,  py,  pz  to  the  surface.  Hence  the  total  momentum  given 
the  surface  in  time  dt  by  all  molecules  is  the  integral  over  momenta  of  the 
quantity  with  components  (pl/rn,  pxp1t/m,  pxpz/m)dtfmdpxdpvdp*. 
Dividing  by  dt,  we  have  the  force  exerted  on  the  square  centimeter  of 
surface,  which  has  components 

f  f  (V 

x  component  of  force  =    I    I    I  —fm  dpx  dpy  dpz, 

y  component  of  force  =    I    I    I  — - fm  dpx  dpu  dpz, 

z  component  of  force  =    I    I    I         fm  dp*  dpy  dpt.  (3.2) 

Now  we  shall  limit  our  distribution  function.  We  shall  assume  that 
a  molecule  with  a  given  value  of  px  is  equally  likely  to  have  positive  or 
negative  values  of  py  and  pz,  so  that  fm(px>  py}  =  f™(pxj  —  py),  etc. 
Plainly  the  Maxwell  law  (3.1)  satisfies  this  condition.  Then  the  second 
and  third  integrals  of  Eq.  (3.2)  will  be  zero,  since  the  integrands  with  a 
given  px,  py  will  have  opposite  sign  to  the  integrands  at  px,  —py,  and  will 
cancel  each  other.  The  force  on  the  unit  area,  in  other  words,  is  along  the 
normal,  or  corresponds  to  a  pure  pressure,  without  tangential  forces. 
Thus  we  finally  have 


///* 


»  dpx  dpv  dpz.  (3.3) 


Now  fm  dpx  dpy  dpz  is  the  number  of  molecules  per  unit  volume  in  the 
range  dpx  dpv  dpz.  Multiplying  by  pl/m  and  integrating,  we  have  simply 
the  sum  of  pl/m  for  all  molecules  in  unit  volume.  This  is  simply  the  sum 
of  pl/m  for  all  molecules  of  the  gas,  divided  by  V.  Let  us  assume  that  the 
distribution  function  is  one  for  which  all  directions  in  space  are  equivalent. 
This  is  the  case  with  the  Maxwell  distribution,  Eq.  (3.1),  for  this  depends 
only  on  the  magnitude  of  p,  not  on  its  direction.  Then  the  sum  of 


60  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IV 

equals  that  of  pj/m  or  pl/m.     We  have,  moreover, 

rf  +  Py  +  Sf.P^^X  kinetic  energy.  (3.4) 

m        m       m        m 

Hence,  the  sum  of  pl/m  is  two-thirds  the  total  kinetic  energy,  and  since 
there  is  no  potential  energy,  this  is  two-thirds  the  internal  energy. 
Finally,  then,  we  have 

p-2^, 

^  ~3F' 
or 

PV  =  f  U.  (3.5) 

Equation  (3.5)  gives  the  relation  predicted  by  kinetic  theory  for  the 
pressure  of  a  perfect  monatomic  gas,  in  terms  of  its  internal  energy  and 
volume. 

We  can  now  combine  Eq.  (3.5),  and  Eq.  (2.7),  giving  the  mean  kinetic 
energy  of  a  monatomic  gas,  to  find  the  equation  of  state.  From  Eq. 
(2.7),  we  have  at  once  for  N  molecules 

U  =  $NkT.  (3.6) 

Combined  with  Eq.  (3.5),  this  gives  at  once 

PV  =  NkT,  (3.7) 

as  the  equation  of  state  of  a  perfect  gas,  as  derived  by  elementary  meth- 
ods. We  should  compare  Eq.  (3.7)  with  the  gas  law  as  ordinarily  set  up 
from  experimental  measurements.  Let  us  suppose  that  we  have  n  moles 
of  our  gas.  That  is,  the  mass  of  gas  wo  are  dealing  with  is  nM,  where  M 
is  the  molecular  weight.  Then  the  usual  law  is 

PV  =  nRT,  (3.8) 

where  R,  the  gas  constant  per  mole,  is  given  alternatively  by  the  numerical 
values 


R  =  8.314  X  107  ergs  per  degree 
=  0.08205  l.-atm.  por  degree 


(3-9) 


The  law  (3.8)  expresses  not  only  Boyle's  and  Charles's  laws,  but  also 
Avogadro's  law,  stating  that  equal  numbers  of  moles  of  any  two  perfect 
gases  at  the  same  pressure  and  temperature  occupy  the  same  volume. 
Now  let  JV0  be  the  number  of  molecules  in  a  gram  molecular  weight,  a 
universal  constant.  This  is  ordinarily  called  Avogadro's  number  and  is 


SBC.  3]         THE  M  AXW  ELL-BOLT  Z  MANN  DISTRIBUTION  LAW  61 

given  approximately  by 

#o  -  6.03  X  1028,  (3.10) 

by  methods  based  on  measurement  of  the  charge  on  the  electron.     Then 
we  have 

N  =  rcAT0,  (3.11) 

MO  that  Eq.  (3.8)  is  replaced  by 

PV  =  N-^-T.  (3.12) 

-tVo 

Equation  (3.12)  agrees  with  Eq.  (3.7)  if 

n 

k  =  -^-i         or        R  =  Nok, 
NO 

so  that 

8.314  X  107 
6.03  X  102* 
=  1.379  X  10-lfi  erg  per  degree,  (3.13) 

as  was  stated  in  Chap.  Ill,  Sec.  1. 

From  the  internal  energy  (3.6)  we  can  also  calculate  the  specific  heat. 
We  have 


The  expression  (3.14),  as  we  have  mentioned  before,  gives  the  heat  capac- 
ity of  n  moles  of  gas.  The  specific  heat  is  the  heat  capacity  of  1  gin.,  or 
1/M  moles,  if  M  is  the  molecular  weight.  Thus  it  is 

3  R 

Specific  heat  per  gram  =  HT>'  (3.15) 

Very  often  one  also  considers  the  molecular  heat,  the  heat  capacity  per 
mole.  This  is 

Molecular  heat,  per  mole  =  f  72  =  2.987  cal.  per  mole.       (3.16) 

To  find  the  specific  heat  at  constant  pressure,  we  may  use  Eq.  (5.2), 
Chap.  II.  This  is 


(-Y 

? 


(3.17) 

^P/r 

which  holds  for  any  amount  of  material.     Substituting  V  =  nRT/P,  we 
have  (dV/dT)f  =  nR/P,  (dV/dP)T  -  -nRT/P\  so  that 

CP  =  CV  +  nR  =  -Jnff  =  4.968  cal.  per  mole,  (3.18) 


62  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IV 

and 

^  =  7=f  =  1.667.  (3.19) 

v/  V  O 

The  results  (3.14),  (3.18),  and  (3.19)  hold  theoretically  for  monatomic 
perfect  gases,  and  actually  they  are  approximately  true  for  real  monatomic 
gases. 

4.  The  Perfect  Gas  in  a  Force  Field. — For  two  sections  we  have  been 
considering  the  distribution  of  velocities  included  in  the  Maxwell-Boltz- 
mann  distribution  law.  Next,  we  shall  take  up  the  distribution  of  coordi- 
nates in  cases  where  there  is  an  external  field  of  force.  First,  we  should 
observe  that  on  account  of  the  form  of  Eqs.  (1.4)  and  (1.5),  the  distribu- 
tions of  coordinates  and  velocities  are  independent  of  each  other.  These 

equations  show  that  the  distribution  function  contains  one  factor  e   2mkT 

depending  on  velocities,  another  factor  e  kT  depending  on  coordinates. 
This  has  important  implications.  The  Maxwell  distribution  of  velocities. 
wfafoh  we  l^ave  discu^snd.  is  the  *ffiflic  at  any  point  of  a  gas,  even  in  an 
external  field;  and  the  variation  of  density  with  position  is  the  same  for 
the  whole  density,  or  for  the  particular  class  of  molecules  having  an v 
chosen  velocity.  We  wish,  then,  to  discuss  the  variation  of  density  com- 
ing from  the  factor  e  k^.  The  most  familiar  example  of  this  formula  is 
found  in  the  decrease  of  density  of  the  atmosphere  as  we  go  to  higher 
altitudes.  The  potential  energy  of  a  molecule  of  mass  m  at  height  ft 
above  the  earth  is  mgft,  where  g  is  the  acceleration  of  gravity.  Then  the 
density  of  gas  at  height  ft,  assuming  constant  temperature  throughout 
(which  is  not  a  good  assumption  for  the  actual  atmosphere),  is  given  by 

mgh 

Density  proportional  to  e    kr .  (4.1) 

Formula  (4.1)  is  often  called  the  barometer  formula,  since  it  gives  the 
variation  of  barometric  pressure  with  altitude.  It  indicates  a  gradual 
decrease  of  pressure  with  altitude,  going  exponentially  to  zero  at  infinite 
height. 

The  barometer  formula  can  be  derived  by  elementary  methods,  thus 
checking  this  part  of  the  Maxwell-Boltzmann  distribution  law.  Con- 
sider a  column  of  atmosphere  1  sq.  cm.  in  cross  section,  and  take  a  section 
of  this  column  bounded  by  horizontal  planes  at  heights  ft  and  ft  +  dh. 
Let  the  pressure  in  this  section  be  P;  we  are  interested  in  the  variation  of 
P  with  ft.  Now  it  is  just  the  fact  that  the  pressure  is  greater  on  the  lower 
face  of  the  section  than  on  the  upper  one  which  holds  the  gas  up  against 
gravity.  That  is,  if  P  is  the  upward  pressure  on  the  lower  face,  P  +  dP 
the  downward  pressure  on  the  upper  face,  the  net  downward  force  is  dPy 


SBC.  4]         THE  MAXWELL-BOLTZMANN  DISTRIBUTION  LAW  6 

the  net  upward  force  —  dP,  and  this  must  equal  the  force  of  gravity  on  th 
material  in  the  section.  The  latter  is  the  mass  of  the  gas,  times  g 
The  mass  of  the  gas  in  the  section  is  the  number  of  molecules  per  uni 
volume,  times  the  volume  dh,  times  the  mass  m  of  a  molecule.  The  num 
ber  of  molecules  per  unit  volume  can  be  found  from  the  gas  law,  whicl 
can  be  written  in  the  form 

P  =  ^*r,  (4.2 

where  (N/V)  is  the  number  of  molecules  per  unit  volume.  Then  we  fine 
that  the  mass  of  gas  in  the  volume  dh  is  (P/kT)m  dh.  The  differentia 
equation  for  pressure  is  then 


In  P-  const.  -,  (4.3; 

from  which,  remembering  that  at  constant  temperature  the  pressure  if- 
proportional  to  the  density,  we  have  the  barometer  formula  (4.1). 

It  is  possible  not  only  to  derive  the  barometer  formula,  but  the  whole 
Maxwell-Boltznuum  distribution  law,  by  an  extension  of  this  method, 
though  we  shall  not  do  it.1  One  additional  assumption  must  be  made, 
which  we  have  treated  as  a  consequence  ot  the  distribution  law  rather  thin 
as  an  independent  hypothesis:  that  the  mean  kinetic  o^gy  flf  ft  ™nW"]' 
is  4ferT  indepn^Hnnt  of  whore  it  may  hn  found.  Assuming  it  and  con- 
sidering the  distribution  of  velocities  in  a  gravitational  field,  we  seem  at 
first  to  meet  a  paradox.  Consider  the  molecules  that  are  found  low  in 
the  atmosphere,  with  a  certain  velocity  distribution.  As  any  one  oi 
these  molecules  rises,  it  is  slowed  down  by  the  earth's  gravitational  field, 
the  increase  in  its  potential  energy  just  equaling  the  decrease  in  kinetic- 
energy.  Why,  then,  do  not  the  molecules  at  a  great  altitude  have  lowci 
average  kinetic  energy  than  those  at  low  altitude?  The  reason  is  not 
difficult  to  find.  The  slower  molecules  at  low  altitude  never  reach  the 
high  altitude  at  all.  They  follow  parabolic  paths,  whose  turning  points 
come  at  fairly  low  heights.  Thus  only  the  fast  ones  of  the  low  molecules 
penetrate  to  the  high  regions;  and  while  they  are  slowed  down,  they  slow 
down  just  enough  so  that  their  original  excessive  velocities  are  reduced 
to  the  proper  average  value,  so  that  the  average  velocity  at  high  altitudes 
equals  that  at  low,  but  the  density  is  much  lower.  Now  this  explanation 

^cc  for  instance,  K.  F.  Herzfeid,  "Kinetische  Theorie  der  Warme,"  p.  20, 
Vieweg,  1929. 


64  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IV 

puts  a  very  definite  restriction  on  the  distribution  of  velocities.  By  the 
barometer  formula,  we  know  the  way  in  which  the  density  decreases  from 
h  to  h  +  dh.  But  the  molecules  found  at  height  A,  and  not  at  height 
h  +  dhj  are  just  those  which  come  to  the  turning  point  of  their  paths 
between  these  two  heights.  This  in  turn  tells  us  the  vertical  components 
of  their  velocities  at  any  height.  The  gravitational  field,  in  other  words, 
acts  to  spread  out  molecules  according  to  the  vertical  component  of  their 
velocities.  And  a  simple  calculation  based  on  this  idea  proves  to  lead 
to  the  Maxwell  distribution,  as  far  as  the  vertical  component  of  velocities 
is  concerned.  No  other  distribution,  in  othnr  wnrdsr  would  have  this 
special  property  of  giving  the  same  distribution  of  velocities  at  anv 
height. 

The  derivation  which  we  have  given  for  the  barometer  formula  in 
Eq.  (4.3)  can  be  easily  extended  to  a  general  potential  energy.  Let  the 
potential  energy  of  a  molecule  be  <t>.  Then  the  force  acting  on  it  is  d<t>/ds, 
where  ds  is  a  displacement  opposite  to  the  direction  in  which  the  force 
acts.  Take  a  unit  cross  section  of  height  ds  in  this  direction.  Then,  as 
before,  we  have 


In  P  =  const.  -  pp,  (4.4) 

general  formula  for  pressure,  or  density^  as  it  depends  on  potential 
energy. 


CHAPTKR  V 
THE  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS 

The  Maxwell-Boltzmann  distribution  law,  which  we  have*  derived 
and  discussed  in  the  last  chapter,  seems  like  a  perfectly  straightforward 
application  of  our  statistical  methods.  Nevertheless,  when  we  come  to 
examine  it  a  little  more  closely,  we  find  unexpected  complications,  arising 
from  the  question  of  whether  there  really  is  any  way  of  telling  the  mole- 
cules of  the  gas  apart  or  not.  We  shall  analyze  these  questions  in  the 
present  chapter,  and  shall  find  that  on  account  of  the  quantum  theory  r  t.lm 
Maxwell-Boltzinann  distribution  law  is  really  only  an  approximation 
valid  for  gases  at  comparatively  low  density,  a  limiting  case?  of  two  other 
distributions,  known  by  the  names  of  the  Fcrmi-Dirac  and  the  KlinsteiiY 
Bosc  statistics.  _  Real  gases  obey  one  or  the  other  of  these  latter  forms  of 
statistics,  some  being  governed  by  one,  some  by  the  other.  As  a  matter 
of  fact,  for  all  real  fiases  the  nnrrnp.t.innH  t.n  t.ho  MnYwp1]-fin]tzm.inn  dis- 
tribution law  which  result  from  the  quantum  statistics  an*  negligibly 
small  except  at  the  very  lowest  temperatures,  and  helium  is  the  only  fi^s 
remaining  in  the  vauor  state  at  low  enough  temperature  for  the  correc- 
tions to  be  important.  Thus  the  reader  who  is  interested  only  in  molecu- 
lar gases  may  well  feel  that  these  forms  of  quantum  statistics  are 
unnecessary.  There  is  one  respect  in  which  this  feeling  is  not  justified: 
we  shall  find  that  in  the  calculation  of  the  entropy,  it  is  definitely  wrong 
not  to  take  account  of  the  identity  of  molecules.  But  the  real  iiimortanr»n 

of  the  nnfUltllTn  fo^T^fi  nf  Jttfq.f.l*jt.lf».H  Pompfi  from  f.hfl  ffU'.f  i.ltn.f,  f.lw*  nl^nfrnnw 

in  solids  satisfy  the  Fcrmi-Uirac  statistics,  and  for  them  thf*  nn  moving] 
quantities  are  such  that  the  behavior  is  completely  different  frorp  what 
would  be  predicted  bv  the  M^wn11-**"1*"™"""  '1^+™^"+:™  law.  The 
^""tHn-RrtfiP  ^-p*™*™^  tho11^  *f  MS  applications  to  black-body  radia- 
tion,  does  not  have  the  general  importance  of  the  Fermi-Dirac  statistics. 
1.  The  Molecular  Phase  Space. — In  the  last  chapter,  we  pointed  out 
that  for  a  gas  of  N  identical  molecules,  each  of  n  degrees  of  freedom,  the 
phase  space  of  2Nn  dimensions  could  be  subdivided  into  N  subspaces, 
each  of  2n  dimensions.  We  shall  now  consider  a  different  way  of  describ- 
ing our  assembly.  We  take  simply  a  2/i-dimensional  space,  like  one  of  our 
previous  subspaces,  and  call  it  a  molecular  phase  space,  since  a  point 
in  it  gives  information  about  a  single  molecule.  This  molecular  phase 
space  will  be  divided,  according  to  the  quantum  theory,  into  cells  of 
volume  hn.  A  given  quantum  state,  or  complexion,  of  the  whole  gas  of  N 

65 


66 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  V 


molecules,  then  corresponds  to  an  arrangement  of  N  points  representing 
the  molecules,  at  definite  positions,  or  in  definite  cells,  of  the  molecular 
phase  space.  But  now  we  meet  immediately  the  question  of  the  identity 
of  the  molecules.  Are  two  complexions  to  be  counted  as  the  same,  or  as 
different,  if  they  differ  only  in  the  interchange  of  two  identical  molecules 
between  cells  of  the  molecular  phase  space?  Surely,  since  two  such 
complexions  cannot  be  told  apart  by  any  physical  means,  they  should 
not  be  counted  as  different  in  our  enumeration  of  complexions.  Yet  they 
correspond  to  different  cells  of  our  general  phase  space  of  2Nn  dimen- 
sions. In  one,  for  example,  molecule  1  may  be  in  cell  a  of  the  molecular 
phase  space,  molecule  2  in  cell  6,  while  in  the  other  molecule  1  is  in  cell 
bj  molecule  2  in  cell  a.  Tlm&JiLJj_S¥!stem.  of  identical  molecules,  it  is 
incorrect  to  assume  that  every  complexion  that  we  can  set  up  in  the 
general  phase  space,  by  assigning  each  molecule  to  a  particular  cell  of 
its  ffiKgpapp  yq  Hi.gf.iti/»+,  1™™  /vir^y  ^hor  Qfi  ^yp  tanitly  assumed  in  Chap. 


By  considering  the  molecular  phase  space,  we  can  see  how  many 
apparently  different  complexions  really  are  to  be  grouped  together  as 
one.  Let  us  describe  a  complexion  by  numbers  Nv,  representing  the 
number  of  molecules  in  the  Mi  cell  of  the  molecular  phase  space.  This 
is  a  really  valid  way  of  describing  the  complexion;  interchange  of  identical 
molecules  will  not  change  the  Nfs.  How  many  ways,  we  ask,  are  there 
of  setting  up  complexions  in  the  general  phase  space  which  lead  to  a  given 
set  of  JWs?  We  can  understand  the  question  better  by  taking  a  simple 
example.  Let  us  suppose  that  there  are  three  cells  in  the  molecular  phase 
space  and  three  molecules,  and  that  we  arc  assuming  Ni  —  1,  N%  =  2, 

TABLE  V-l 


Cell 

1 

2 

3 

,.. 

1 

2 

0 

Complexion 

a 

i! 

2  3 
3  2 

!2 

1  3 

2 

3  1 

J3 

1  2 

c 

is 

2  1 

NB  =  0,  meaning  that  one  molecule  is  in  the  first  cell,  two  in  the  second, 
and  none  in  the  third.  Then,  as  we  see  in  Table  V-l,  there  are  three 
apparently  different  complexions  leading  to  this  same  set  of  Ni's.  In 
complexion  a,  molecule  1  is  in  cell  1,  and  2  and  3  are  in  cell  2;  etc.  We 


SBC.  1]      THE  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS          67 

see  from  this  example  how  to  find  the  number  of  complexions.  First,  we 
find  the  total  number  of  permutations  of  N  objects  (the  N  molecules). 
This,  as  is  well  known,  is  TV!;  for  any  one  of  the  N  objects  can  come  first, 
any  one  of  the  remaining  (N  —  1  )  second,  and  so  on,  so  that  the  number 
of  permutations  is  N(N  —  i)(N  —  2)  .  .  .  2.1  =  Nl  In  the  case  of 
Table  V-l,  there  are  3!  =  6  permutations.  But  some  of  these  do  not 
represent  different  complexions,  even  in  the  general  phase  space,  as 
we  show  by  our  brackets;  as  for  example  the  arrangements  1,23  and  1,32 
grouped  under  the  complexion  (a).  For  they  both  lead  to  exactly  the 
same  assignment  of  molecules  to  cells.  In  fact,  if  any  N%  is  greater  than 
unity,  Nil  (in  our  case  2!  =  2)  different  permutations  of  the  N  objects 
will  correspond  to  the  same  complexion.  And  in  general,  the  number  of 
complexions  in  the  general  phase  space  which  lead  to  the  same  TVYs,  and 
hence  are  really  identical,  will  be 


Remembering  that  0!  =  1!  =  1,  we  see  that  in  our  example  we  have 
3  1/1  12  10!  =|=  3. 

If  then  we  wished  to  find  the  partition  function  for  a  perfect  gas,  using 
the  general  phase  space,  we  should  have  to  proceed  as  follows.  We 
couldset  uu  cells  in  phase  space,  each  of  volume  hNn.  but  we  could  not 
assume  that  each  of  these  represented  a  different  complexion,  or  thftt  we 
were  to  sum  over  all  these  flnlla  in  nnrnpiifing  +^  poi.*^™*  1,,^;,^ 
Rather,  each  cell  would  be  one  of  Nl/N-dNf]  -  .  .  similar  cells,  all  taken 
together  fa  ygpreseiyfr  one  single  complexion.  We  could  handle  this  .11 
we  chose^  bv  summing,  or  in  the  case  of  continuous  energy  levels  fry 
i  ntegrathifi.  over  all  cells,  but  dividinp:  the  contribution  of  each  cell  by 
the  number  (1.1).  computed  for  that  cell.  Since  this  number  can  change 
from  cell  to  cell,  this  is  a  very  inconvenient  procedure  and  cannot  be 
carried  out  without  rather  complicated  mathematical  methods.  There 
is  a  special  case,  however,  in  which  it  is  very  simple.  This  is  the  case 
where  the  gas  is  so  rare  jfchat  we  are  very  unlikely!  in  our  assembly,  to  find 
any  appreciablo^iiiinib£r  of  systems  with  more  than  a  single  molecule  in 
any  cell.  In  this  case,  each  of  the  JV^s^iiLJJifi_denQminatQr  of  formula 
(1.1)  will  be  0  or  1,  each  of  the  JVJ's  will  be  1.  and  the  number  (1.1) 
becomes  simply  Nl.  Thus,  in  this  case,  we  can  find  the  partition  function 
by  carrying  out  the  summation  or  integration  in  the  general  phase  space 
in  the  usual  way,  but  dividing  the  result  by  JVI,  and  using  the  final  value 
to  find  the  Helmholtz  free  energy  and  other  thermodynamic  quantities. 
This  method  leads  to  the  Mn.Y^p|1-Rn1t.7.mnnn  HiHf.Hhiitjpn  IRW.  and  it  is 
the  method  which  we  shall  use  later  in  Chap.  VIII,  dealing  with  thermo- 
dynamic and  statistical  properties  of  ordinary  perfect  gases.  When  we 


68  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

are  likely  to  find  JVYs  greater  than  unity,  however,  this  method  is  imprac- 
ticable, and  we  must  adopt  an  alternative  method  based  directly  on  the 
use  of  the  molecular  phase  space. 

2.  Assemblies  in  the  Molecular  Phase  Space.  —  When  we  describe  a 
system  by  giving  the  JWs,  tho  numbers  of  molecules  in  each  coll  of  the 
molecular  phase  space,  we  automatically  avoid  the  difficulties  described 
in  the  last  section  relating  to  the  identity  of  molecules.  We  now  meet 
immediately  the  distinction  between  the  Fcrmi-Dirac,  the  Einstein-Bose, 
and  the  classical  or  Boltzmann  statistics.  In  the  Einstcin-Bose  statistics. 
the  simplest  form  in  thenrv  WP  «^f  up  a  comnlexion  hv  giving  a.  set  Of 
Ni's,  and  we  say  that  any  possible  set  of  JWs,  subject  only  to  the  obvious 
restriction 


represents  a.  possible  e.orflplexion.  all  florflplexions  having  equal  A  priori 
probability.  The  Fermi-Dirac  statistics  differs  from  the  Einstein-Bose 
in  that  there  is  ayi  n^ti™™!  p^^pl^,  nnllpH  fh»  pv»1i"=jjnn  prj^piQ 
superposed  on  the  principle  of  identity  of  molennles..  The  exclusion 
nrTneiple  states  that  no  two  molecules  mav  he  in  the  same  nail  nf  the 
molecular  phase  space  at  the  same  time:  that  is.  no  one  of  t.hp.  /y.'«  nrm.v 
be  greater  than  unity.  This  principle  gains  its  importance  from  the  fact 
that  electrons  are  found  experimentally  to  obey  it.  It  is  a  principle 
which,  as  we  shall  see  later,  is  at  the  foundation  of  the  structure  of  the 
atoms  and  the  periodic  table  of  the  elements,  as  well  as  having  the  greatest 
importance  in  all  problems  involving  electrons.  In  the  Fermi-Dirac 
statistics,  then,  any  possible  set  of  JWs.  subject  to  ECL  (2.1)  and  to  the 

nrn  OT  unitvr  forms  a 


possible  complexion  nf  the  ^ysteni.  Finally  the  Boltzmann  statistics  is 
the  limiting  case  of  cither  of  the  other  types,  in  the  limit  of  low  density, 
where  so  few  molecules  ^ajiL-disijibutod  amiing—  ^o-uiisiiiszL-cells  thfl/h  the 
chance  of  finding  two  in  the  saino_c^ll  js  negligible  anvwftyl  ^i\c\  thn  difYnr- 
ence  between  the  Fcrmi-Dirac  and  the  Einstein-Bo*^  fftitjtfif  ^  ^«*»pp»«^ 
Let  us  consider  a  single  complexion,  represented  by  a  set  of  A/Ys,  in 
the  molecular  phase  space.  We  see  that  the  JVYs  are  likely  to  fluctuate 
greatly  from  cell  to  cell.  For  instance,  in  the  limiting  case  of  the  Boltz- 
mann statistics,  where  there  are  many  fewer  molecules  than  cells,  we  shall 
find  most  of  the  JVYs  equal  to  zero,  a  few  equal  to  unity,  and  almost  none 
greater  than  unity.  It  is  possible  in  principle,  according  to  the  principle 
of  uncertainty,  to  know  all  the  A/Ys  definitely,  or  to  prepare  an  assembly 
of  systems  all  having  the  same  JVYs.  But  for  most  practical  purposes 
this  is  far  more  detailed  information  than  we  require,  or  can  ordinarily 
give.  We  have  found,  for  instance,  that  the  translational  energy  levels 


SEC.  2)       THE  FERMI-DIRAC  AND  El N STEIN-BOISE  STATISTICS          69 

of  an  ordinary  gas  are  spaced  extremely  close  together,  and  while  there  is 
nothing  impossible  in  principle  about  knowing  which  levels  contain 
molecules  and  which  do  not,  still  practically  we  cannot  tell  whether 
a  molecule  is  in  one  level  or  a  neighboring  one.  In  other  words,  for 
this  case,  for  all  practical  purposes  the  scale  of  our  observation  is  much 
coarser  than  the  limit  set  by  the  principle  of  uncertainty.  Let  us,  then, 
try  to  set  up  an  assembly  of  systems  reflecting  in  some  way  the  actual 
errors  that  we  are  likely  to  make  in  observing  molecular  distributions. 
Let  us  suppose  that  really  we  cannot  detect  anything  smaller  than  a 
group  of  G  cells,  where  G  is  a  rather  large  number,  containing  a  rather 
large  number  of  molecules  in  all  the  systems  of  our  assembly.  And  let  us 
assume  that  in  our  assembly  the  average  number  of  molecules  in  the  ith 
cell,  one  of  our  group  of  G,  isffit,  a  quantity  that  ordinarily  will  b<»  a  frac- 
tion rather  than  an  integer  In  the  particular  case  01  ttie  uoit/mami 
statistics.  N .  ^jll  b^  n.  f motion  much  less  than  unity  \  in  the  Feriui-Dirac . 
statistics  it  will  be  les^  than  unity,  but  not  necessarily  much  less;  while  in 
the  Einstein-Bose  statistics  it  can  have  any  value.  We  shall  now  try  to 
set  up  an  assembly  leading  to  these  postulated  values  of  the  N^.  To  do 
this,  we  shall  find  all  the  complexions  that  lead  to  the  postulated  $Ys,  in 
the  sense  of  having  NtG  molecules  in  the  group  of  G  cells,  and  we  shall 
assume  that  these  complexions,  and  these  only,  are  represented  in  the 
assembly  and  with  equal  weights.  Our  problem,  then,  is  to  calculate  the 
number  of  complexions  consistent  with  a  given  set  of  JVYs,  or  the  thermo- 
dynamic  probability  W  of  the  distribution,  in  Boltzinann's  sense,  as 
described  in  Chap.  Ill,  Sec.  1.  Having  found  the  thermodynamic  prob- 
ability, we  can  compute  the  entropy  of  the  assembly  by  the  fundamental 
relation  (1.3)  of  Chap.  Ill,  or 

8  =  *  In  W.  (2.2) 

For  actually  calculating  the  thermodynamic  probability,  we  must 
distinguish  between  the  Fermi-Dirac  and  the  Einstein-Bose  statistics. 
First  we  consider  the  Fermi-Dirac  case.  We  wish  the  number  of  ways  of 
arranging  N£f  molecules  in  G  cells,  in  such  a  way  that  we  never  have 
more  than  one  molecule  to  a  cell.  To  find  this  number,  imagine  G  coun- 
ters, of  which  N£f  are  labeled  1  (standing  for  1  molecule),  and  the  remain- 
ing (1  —  ftt)G  are  labeled  0  (standing  for  no  molecules).  If  we  put  one 
counter  in  each  of  the  G  cells,  we  can  say  that  the  cells  which  have  a 
counter  labeled  1  in  them  contain  a  molecule,  the  others  do  not.  Now 
there  are  G\  ways  of  arranging  G  counters  in  G  cells,  one  to  a  cell,  as  we 
have  seen  in  the  last  section.  Not  all  of  these  G\  ways  of  arranging  the 
counters  lead  to  different  arrangements  of  the  molecules  in  the  cells, 
however,  for  the  N$  counters  labeled  1  are  identical  with  each  other,  and 


70  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

those  labeled  0  are  identical  with  each  other.  For  a  given  assignment  of 
molecules  to  cells,  there  will  be  (N£f)  1  ways  of  rearranging  the  counters 
labeled  1,  and  [(1  —  Ni)G]\  ways  of  rearranging  those  labeled  zero,  or 
(NlG)l[(l  —  Nl)G]\  arrangements  in  all,  all  of  which  lead  to  only  one 
complexion.  Thus  to  get  the  total  number  of  complexions,  we  must 
divide  G\  by  this  quantity,  finding 

Number  of  complexions  of  NjB  atoms  in  G  cells  =  —  _ 


-•  --        (2.3) 

We  shall  now  rewrite  Eq.  (2.3),  using  Stirling's  theorem.    This  states 
that,  for  a  large  value  of  N, 


Nl  =  V'lirN  ~     •  (2.4) 

Stirling's  formula  is  fairly  accurate  for  values  of  N  greater  than  10;  for 
still  larger  JV's,  where  N\  and  (N/e)N  are  very  large  numbers,  the  factor 
is  so  near  unity  in  proportion  that  it  can  be  omitted  for  most 


purposes,  so  that  we  can  write  Nl  simply  as  (N/e)N.     Adopting  this 
approximation,  we  can  rewrite  Eq.  (2.3)  as 

Number  of  complexions  of  N1G  atoms  in  G  cells 


[1 
#/'(!     - 


Equation  (2.5)  is  of  an  interesting  form:  being  a  quantity  independent 
of  G,  raised  to  the  G  power,  we  may  interpret  it  as  a  product  of  terms,  one 
for  each  cell  of  the  molecular  phase  space.  Now  to  get  the  whole  number 
of  complexions  for  the  system,  we  should  multiply  quantities  like  (2.5),  for 
each  group  of  G  cells  in  the  whole  molecular  phase  space.  Plainly  this 
will  give  us  something  independent  of  the  exact  way  we  divide  up  the 
cells  into  groups,  or  independent  of  (?,  and  we  find 


where  JJ  indicates  a  product  over  all  cells  of  the  molecular  phase  space. 


SBC.  2]      THE  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS         71 


Using  Eq.  (2.2),  we  then  have 
S  =  -*5[#<  In 


-  #    In   1  - 


(2.7) 


as  the  expression  for  entropy  in  the  Formi-Dirac  statistics  in  terms  of  the 
average  number  Ni  of  molecules  in  each  cell. 

For  the  Einstein-Bose  statistics,  we  wish  the  number  of  ways  of 
arranging  Nfr  molecules  in  G  cells,  allowing  as  many  molecules  as  we 
please  in  a  cell.  This  number  ran  be  shown  to  be 

Number  of  complexions  of  N1G  atoms  in  G  cells 

_  (#/?+(?  -1)1 


-  1)! 


' 


We  can  easily  make  Eq.  (2.8)  plausible,1  though  without  really  proving 
it,  by  an  example.  Let  us  take  N1G  =  2,  G  =  3,  and  make  a  table,  as 
Table  V-2,  showing  the  possible  arrangements: 

TABLE  V-2 


Cell  1 

2 

'      3 

11 

0 

0 

1 

1 

0 

0 

1 

I 

0 

I 

0 

1 

0 

1 

0 

1 

1 

0 

0 

1 

0 

11 

0 

0 

1 

1 

0 

0 

1 

1 

0 

1 

0 

1 

0 

0 

11 

0 

0 

1 

1 

In  the  first  three  columns  of  Table  V-2,  we  indicate  the  three  colls,  and 
indicate  each  of  the  two  molecules  by  a  figure  1,  showing  the  six  possible 

arrangements    6  =  -  ^  ~  —    -v  .-"   ;  following  that,  we  give  a  scheme  with 
L  *H«J  —  l;i   J 


_ 

four  columns  (4  =  NtG  +  G  —  1  =2  +  3  —  1)  in  which  we  give  all  the 
possible  arrangements  of  the  two  1's,  two  O's,  with  one  in  each  column 
(2  =  NtG,  2  =  G  —  1).  In  the  general  case,  the  number  of  such  arrange- 
ments is  given  just  by  Eq.  (2.8),  as  wo  can  see  by  arguments  similar  to 
those  used  in  deriving  formula  (1.1).  But  the  four-column  arrangement 
of  Table  V-2  corresponds  exactly  witli  the  three-column  one,  if  we  adopt 
the  convention  that  two  successive  IV  in  the  four-column  scheme  belong 
in  the  same  cell.  It  is  not  hard  to  show  that  the  same  sort  of  correspond- 
ence holds  in  the  general  case  and  thus  to  justify  Eq.  (2.8). 

Applying  Stirling's  theorem  to  Eq.  (2.8)  and  neglecting  unity  com- 
pared to  G,  we  now  have 

1  For  proof,  as  well  as  other  points  connected  with  quantum  statistics,  see  L. 
Brillouin,  "Die  Quantenstatistik,"  pp.  129  iT  ,  Julius  Springer,  1931. 


72  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

Number  of  complexions  of  N&f  atoms  in  G  cells 


[ 


As  with  the  Fermi-Dirac  statistics,  this  is  a  product  of  terms,  one  for 
each  of  the  G  cells;  to  find  the  whole  number  of  complexions,  or 
the  thermodynamic  probability,  we  have 

w  = 

N,»- 
and 

S  =  -fc[#,  In  ft,  -  (1  +  #,)  In  (1  +  #,)]•  (2.11) 


Iii  Eqs.  (2.7)  and  (2.11),  we  have  found  the  general  expressions  for 
the  entropy  in  the  Fermi-Dirac  and  Ernst  ein-Bose  statistics.  From  either 
one,  we  can  find  the  entropy  in  the  Boltzmann  statistics  by  passing  to  the 
limit  in  which  all  JVVs  are  very  small  compared  to  unity.  For  small 
Nt,  In  (1  ±  Nt)  approaches  ±Ni,  and  (1  ±  JV»)  can  be  replaced  by  unity. 
Thus  either  Eq.  (2.7)  or  (2.11)  approaches 

,  ln  ^*  ~  ^*).  (2.12) 

Equation  (2.12)  expresses  the  form  of  the  ontropy  for  the  Boltzmann 
statistics. 

3.  The  Fermi-Dirac  Distribution  Function.  —  In  the  preceding  section, 
we  have  set  up  assemblies  of  systems  satisfying  either  the  Formi-Dirac 
or  the  Einstein-Bose  statistics  and  having  an  arbitrary  average  number  of 
molecules  Nl  in  the  ith  coll  of  molecular  phase  space.  We  have  found  the 
thermodynamic  probability  and  the  entropy  of  such  an  assembly.  These 
assemblies,  of  course,  do  not  correspond  to  thermal  equilibrium  and,  as 
time  goes  on,  the  effect  of  collisions  of  molecules  will  be  to  change  the 
numbers  fft  gradually,  with  an  approach  to  a  steady  state.  In  the  next 
chapter  we  shall  consider  this  process  specifically,  really  following  in 
detail  the  irreversible  approach  to  a  steady  state.  We  shall  verify  then, 
as  we  could  assume  from  our  general  knowledge,  that  during  the  irreversi- 
ble process  the  entropy  will  gradually  increase,  until  finally  in  equilibrium 
it  reaches  the  maximum  value  consistent  with  a  constant  value  of  the 


SBC.  3]       THE  FERMI-D1RAC  AND  EIN8TE1N-BO8E  STATISTICS          73 

total  energy.  But,  for  the  moment,  we  can  assume  this  final  condition 
and  use  it  to  find  the  equilibrium  distribution.  We  ask,  in  other  words, 
what  set  of  Ni's  will  give  the  maximum  entropy,  subject  to  a  constant 
internal  energy?  We  can  solve  this  problem,  as  we  solved  a  similar  one  in 
Chap.  Ill,  Sec.  5,  by  the  method  of  undetermined  multipliers. 

For  the  Fermi-Dirac  statistics,  we  wish  to  make  the  entropy  (2.7)  a 
maximum,  subject  to  a  constant  value  of  the  energy.  Rather  than 
impose  just  this  condition,  we  employ  the  thermodynamically  equivalent 
one  of  making  the  function  A  =  U  —  TS  a  minimum  for  changes  at 
constant  temperature,  as  discussed  in  Chap.  II,  Sec.  3.  This  is  essentially 
a  form  of  the  method  of  undetermined  multipliers,  the  constants  multiply- 
ing U  and  S  being  respectively  unity  and  —  T.  As  in  Chap.  IV,  Sec.  1,  we 
let  the  energy  of  a  molecule  in  the  ith  state  be  «».  Then  the  average 
energy  over  the  assembly  is  clearly 

U  = 

the  summation  being  over  the  cells  of  the  molecular  phase  space.  Using 
Eq.  (2.7)  for  the  entropy,  we  then  have 

A  =  ][#tct  +  kTNi  In  Nl  +  kT(\  -  #,)  In  (1  -  Nt)].       (3.1) 


Now  we  find  the  change  of  A,  when  the  $Vs  are  varied,  keeping  tem- 
perature and  the  e»'s  fixed.     We  find  at  once 


dA  =  0  =  +  kT  ln  —  T^  (3-2) 

t 

as  the  condition  for  equilibrium.     This  must  be  satisfied,  subject  only 
to  the  condition 

i  =  0,  (3.3) 


i 

expressing  the  fact  that  the  changes  of  the  Ni's  are  such  that  the  total 
number  of  molecules  remains  fixed.  The  only  way  to  satisfy  Eq.  (3.2), 
subject  to  Eq.  (3.3),  is  to  have 

e<  +  kT  In  -  *—  =  €<>  =  const.,  (3.4) 

1  -  Ni 

independent  of  i.  For  then  the  bracket  in  Eq.  (3.2)  can  be  taken  outside 
the  summation  sign,  and  Eq.  (3.3)  immediately  makes  the  whole  expres- 
sion vanish. 


74  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

Using  Eq.  (3.4),  we  can  immodiatcly  solve  for  N*.     We  have 

N< 


/ 
N\ 


l  +  e   kT       =  e  ~kT 
<?-*> 

o'kT 


(3'5) 


Equation  (3.5)  expresses  the  Fermi  distribution  law,  which  we  shall  now 
proceed  to  discus:,. 

First,  let  us  see  that  the  Fermi-Dirac  distribution  law  reduces  to  the 
Maxwell-Boltzmann  law  in  the  limit  when  the  NSs  are  small.  In  that 
case,  it  must  be  that  the  denominator  in  Eq.  (3.5)  is  large  compared  to 

the  numerator.  But  if  e  k?  +1  is  large  compared  to  1,  the  numera- 
tor, it  must  be  that  1  can  be  neglected  compared  to  the  exponential. 
Thus,  in  this  limit,  we  can  write 

Ni  =  e^,  (3.6) 

which  is  the  Maxwell-Boltzmann  law,  Eq.  (1.4)  of  Chap.  IV,  if  €0  is  prop- 
erly chosen.  We  notice  that  even  if  the  temperature  is  low,  so  that  some 
JVVs  are  not  small,  still  the  states  of  high  energy  will  have  large  values  of 

e» 

ckT.  Thus  the  argument  we  have  just  used  will  apply  to  these  states, 
and  for  them  the  Maxwell-Boltzmann  distribution  will  be  correct,  even 
though  it  is  not  for  states  of  low  energy. 

The  quantity  «o  is  to  be  determined  by  the  condition  that  the  total 
number  of  particles  is  N.  Thus  we  have 

N-^?,Ni  =  ^-Trrl (3-7) 


Since  the  e»'s  are  determined,  Eq.  (3.7)  can  be  satisfied  by  a  proper  choice 
of  €0.  Unfortunately,  Eq.  (3.7)  cannot  be  solved  directly  for  c0,  and  it  is  a 
matter  of  considerable  difficulty  to  evaluate  this  important  quantity.  It 
is  not  hard,  however,  to  see  how  N*  behaves  as  a  function  of  €t,  particu- 
larly for  low  temperatures.  When  ct  —  eo  is  negative,  or  for  energies 
below  €o,  the  exponential  in  Eq.  (3.5)  is  less  than  unity,  becoming  rapidly 
very  small  as  the  temperature  decreases.  Thus  for  these  energies  the 
denominator  is  only  slightly  greater  than  unity,  and  ft*  only  slightly  less 
than  unity.  On  the  other  hand,  when  a  —  eo  is  positive,  for  energies 
above  €0,  the  exponential  is  greater  than  unity,  becoming  rapidly  large  as 


SEC.  3]       THE  FERMI-DIRAC  AND  EINSTEIN-ROSE  STATISTICS 


75 


the  temperature  decreases.  In  this  case  we  can  almost  neglect  unity 
compared  to  the  exponential,  and  we  have  the  case  of  the  last  paragraph, 
where  the  Boltzmann  distribution  is  approximately  correct,  in  the  form 

(3.6).     In  Fig.  V-l  we  show  the  function  -j—( oy as  a  function  of  e, 

e  w'  +  1 

for  several  temperatures.  At  T  =  0,  the  function  drops  from  unity  to 
zero  sharply  when  e  =  e0,  while  at  higher  temperatures  it  falls  off 
smoothly.  For  large  values  of  €,  it  approximates  the  exponential  falling 
off  of  the  Boltzmann  distribution. 


-7    -6   -5  -4   -3   -2 


67 


FIG.  V-l. — Fermi  distribution  function,  as  function  of  energy,  for  several  temperatures. 
Curve  a,  kT  =  0;  6,  kT  =  1;  c,  kT  =  2.5 

One  important  feature  of  the  function  (3.5)  is  the  following: 


1  - 


kT 


kT 


1-1 


+ 


(3.8) 


kT 


+  1 


That  is,  in  Fig.  V-l,  the  distribution  function  at  any  point  to  the  right 
of  €0  is  equal  to  the  difference  between  the  function  and  unity,  the  same 
distance  to  the  left  of  e0,  and  vice  versa.  The  curve,  in  other  words,  is 
symmetrical  with  change  of  sign  about  the  point  €  =  €o  and  ordinate  £. 
From  this  it  follows  that  €0  is  approximately  constant,  for  small  tempera- 
tures at  least.  For  the  summation  in  Eq.  (3.7),  which  must  give  N 
independent  of  temperature,  is  found  as  follows.  Along  the  axis  of 
abscissae  in  Fig.  V-l,  we  mark  the  various  energy  levels  of  the  problem. 
At  each  energy  level  we  erect  a  line,  extending  up  to  the  distribution 
curve.  The  sum  of  the  lengths  of  all  these  lines  is  the  summation  desired. 
We  must  now  adjust  €0,  moving  the  curve  to  the  left  or  right,  so  that  the 
sum  equals  N.  At  the  absolute  zero  this  is  perfectly  simple:  we  simply 
count  up  to  the  Nth  energy  level  from  the  bottom,  and  put  €0  somewhere 
between  the  Nth  and  the  (N  -f  l)st  levels.  At  a  higher  temperature, 


76  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

suppose  we  try  the  same  value  of  «0  and  see  if  it  is  correct.  Then  the  sum- 
mation will  change  by  subtraction  of  the  part  of  the  lines  above  the  distri- 
bution curve  to  the  left  of  €o  and  by  addition  of  those  below  the  distribution 
curve  to  the  right.  These  areas  are  equal  by  the  result  we  have  just  found. 
Thus,  if  levels  come  with  the  same  density  to  the  left  and  to  the  right  of 
co,  the  summation  will  again  be  N,  and  the  same  value  of  €o  will  be  correct. 
In  the  next  section  we  find  how  much  co  will  change  with  temperature  if 
the  density  of  levels  changes  with  the  energy,  as  of  course  it  will  to  some 
extent. 

4.  Thermodynamic  Functions  in  the  Fermi  Statistics. — Having 
derived  the  Fermi-Dirac  distribution  law,  we  shall  go  on  to  find  some  of 
its  properties  in  the  case  of  low  temperatures,  the  important  case  in  the 
practical  applications  to  electrons  in  metals,  where  for  the  present  pur- 
poses temperatures  even  of  several  thousand  degrees  can  be  regarded  as 
low.  The  distribution  function,  as  we  see  in  Fig.  V-l,  changes  with 
temperature  mostly  in  the  immediate  neighborhood  of  the  energy  c0.  If, 
then,  we  know  the  distribution  of  energy  levels  in  the  neighborhood  of 
c0,  we  can  find  the  variation  of  the  thermodynamic  functions  with  tem- 
perature. We  carry  out  that  analysis  in  the  present  section,  assuming 
that  the  energy  levels  are  distributed  continuously  in  energy,  an  approxi- 
mately correct  assumption  in  the  cases  with  which  we  shall  deal. 

Let  the  value  of  c0  at  the  absolute  zero  of  temperature  be  c0o.  We 
know  how  to  find  it  from  Sec.  3,  simply  by  counting  up  N  levels  from  the 
lowest  level.  First  we  shall  try  to  find  how  eo  depends  on  temperature. 
We  shall  assume  that  the  number  of  energy  levels  between  c  and  e  +  d*  is 

« -  [(£). + $),<•  -  •••>  +  •  •  •  >•      <"> 

a  Taylor's  expansion  of  the  function  dN/de  about  the  point  e  =  e0o,  at 
which  the  derivatives  (dN/dt)^  and  (dW/dc2)0  are  evaluated.  With  this 
assumption,  all  our  summations  can  be  converted  into  integrations.  We 
write  the  summation  of  Eq.  (3.7)  in  the  form  of  an  integration;  instead  of 
using  just  this  form,  we  find  the  difference  between  the  summation  and 
that  at  the  absolute  zero,  which  should  give  a  difference  of  zero.  Thus 
we  have 


0  = 


> 


The  term  —1  in  the  first  integral  takes  care  of  the  summation  at  the 
absolute  zero,  where  the  Fermi  function  is  unity  for  energies  less  than  «oo, 


SEC.  4]       THE  FUHM1-D1HAV  AND  E1NSTJK1N-BOSE  STATISTICS          77 

zero  for  higher  energies.  In  the  first  integral,  we  use  Eq.  (3.8).  Then  in 
the  first  we  make  tho  change  of  variables  u  =  —  (e  —  €0),  and  in  the 
second  u  =  (e  —  €0).  The  two  integrals  then  combine.  We  retain  only 
the  terms  necessary  for  a  first  approximation;  this  means,  it  is  found, 
that  in  the  term  (dW/d€2)0  we  can  neglect  the  distinction  between  c0  and 
€0o,  though  this  distinction  is  essential  in  tho  term  in  (dN/di)^  In  this 
way  we  find 

~<  u       7  t4»\ 

-u  ---  du.        (4.3) 


In  the  first  integral,  through  the  very  small  range  from  c0o  —  to  to  c0  —  e0o, 
we  can  replace  the  integrand  by  its  value  when  u  =  0,  or  £.  Thus  the 
first  term  becomes  (rfJV/de)o(e0  —  e0o).  To  reduce  the  second  integral 

__  u 

to  a  familiar  form,  we  let  x  =  e  *r,  u  =  —  kT  In  x,  du  =  —kTdx/x. 
The  integral  then  becomes 

(4.4) 


f  *-^—du  =  -(*T)>  f  'iqr^r  -  g 

*/o     *r  */        ^ 


the  integral  in  Eq.  (4.4)  being  tabulated  for  instance  in  B.  O.  Peirce's 
"Short  Table  of  Integrals."     Then  Eq.  (4.3)  becomes 


or 


(4.5) 


Equation  (4.5)  represents  €0  by  the  first  two  terms  of  a  power  series  in  the 
temperature,  and  the  approximations  we  have  made  give  the  term  in 
T2  correctly,  though  we  should  have  to  be  more  careful  to  get  higher 
terms.  We  see,  as  we  should  expect  from  the  last  section,  that  if 
(dW/de2)o  =  0,  so  that  the  distribution  of  energy  levels  is  uniform  at 
€0,  c0  will  be  independent  of  temperature  to  the  approximation  we  are 
using. 

Next,  let  us  find  the  internal  energy  in  the  same  sort  of  way.     Written 
as  a  summation,  it  is 

u  =        —  (4-6) 

+  1 


78  INTRODUCTION*™  CHEMICAL  PHYSICS  [CHAP.  V 

Here  again,  in  converting  to  an  integration,  we  shall  find,  not  [/,  but 
U  —  f/o,  where  £70  is  the  value  at  the  absolute  zero.  Then  we  find  at 
once  that  the  integral  expression  for  it  is  exactly  like  the  integral  in  Eq. 
(4.2),  only  with  an  additional  factor  e  in  the  integrand  of  each  integral. 
The  leading  term  hero,  however,  comes  from  the  term  (dAT/de)o,  and  we 
can  neglect  the  terms  in  (rfW/ck'2)0.  Furthermore,  in  the  integrals  we 
retain,  we  can  neglect  the  difference  between  €o  and  eoo.  Then  we  have 


and 

u-u'  +  (f  )l(fcr)2'  (4-7) 

again  correct  to  terms  in  T2.     From  the  internal  energy  we  can  find  the 
heat  capacity  CV,  by  the  equation 


c"  -  (wl 


-tk'T.  (4.8) 

We  notice  that  at  low  temperatures  the  specific  heat  of  a  system  with 
continuous  energy  levels,  obeying  the  Fermi  statistics,  is  proportional 
to  the  temperature.  We  shall  later  see  that  this  formula  has  applications 
in  the  theory  of  metals. 

Let  us  next  find  the  entropy.     We  can  get  a  general  formula  from 
Eq.  (2.7).     This  can  be  rewritten 


5  -  -kN>  In     -r  -  k         In  (1  -  ft.) 


(4'9) 


where  we  have  used  the  Fermi  distribution  law.  Replacing  the  summa- 
tion in  Eq.  (4.9)  by  an  integration  and  using  Eqs.  (4.5)  and  (4.7)  for  c0 
and  £7,  we  can  compute  S.  The  calculation  is  a  little  involved,  however, 
and  it  is  easier  to  use  the  relation 


SEC.  4]  THE  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS  79 
from  which 

8  -  (f  XT**  «•"» 

In  the  integration  leading  to  Eq.  (4.10),  we  have  used  the  fact  that  S  -  0 
at  the  absolute  zero.  This  follows  directly  from  Eq.  (2.7)  for  the  entropy. 
For  each  term  in  the  entropy  is  of  the  form  N  In  N  or  (1  —  N)  In  (1  —  N), 
and  at  the  absolute  zero  each  value  of  N  is  cither  1  or  0,  so  that  each  of 
these  terms  is  1  In  1  or  0  In  0,  either  of  which  is  zero. 

From  the  internal  energy  and  the  entropy  we  can  find  the  Helmholtz 
free  energy 

A  =  U  -  T8 

_-A  _>-«) 

In  (e.      *T~    +  1)  (4.11) 


-  (f 


where  Eq.  (4.11)  is  derived  from  Eq.  (4.9),  and  Eq.  (4.12)  from  Eqs.  (4.7) 
and  (4.10).  By  differentiating  the  function  A  with  respect  to  tempera- 
ture at  constant  volume,  we  get  the  negative  of  the  entropy,  as  we  should. 
By  differentiating  with  respect  to  the  volume  at  constant  temperature,  we 
get  the  negative  of  the  pressure,  and  hence  can  find  the  equation  of  state. 
So  far,  we  have  not  mentioned  the  dependence  of  any  of  our  functions  on 
the  volume.  Surely,  however,  the  stationary  states  and  energy  levels  of 
the  particles  will  depend  on  the  volume,  though  not  explicitly  on  the 
temperature.  Hence  Uo  and  (rf#/rfe)0  are  to  be  regarded  as  functions  of 
the  volume.  The  functional  dependence,  of  course,  cannot  be  given  in  a 
general  discussion,  applicable  to  all  systems,  such  as  the  present  one. 
Using  this  fact,  then,  we  have 


The  first  term,  the  leading  one  ordinarily,  is  independent  of  temperature, 
and  the  second,  a  small  additional  one  which  can  be  of  either  sign,  is 
proportional  to  the  square  of  the  temperature.  Thus,  the  equation  of 
state  is  very  different  from  that  of  a  perfect  gas  on  Boltzmann  statistics. 
It  is  to  be  borne  in  mind,  however,  that  this  formula,  like  all  those  of  the 
present  section,  applies  only  at  low  temperatures  and  is  only  the  beginning 
of  a  power  series.  At  high  temperatures  the  statistics  reduce  to  the 
Boltzmann  statistics,  as  we  have  seen,  and  the  equation  of  state  of  a  Fermi 


80  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

system  and  the  corresponding  system  obeying  Boltzmann  statistics  must 
approach  each  other  at  high  temperature. 

5.  The  Perfect  Gas  in  the  Fermi  Statistics. — As  an  example  of  the 
application  of  the  Fermi  statistics,  we  can  consider  the  perfect  gas.  In 
Chap.  IV,  Sec.  1,  we  have  found  the  number  of  energy  levels  for  a  mole- 
cule of  a  perfect  gas,  in  the  energy  range  de.  Rewriting  Eq.  (1.10)  of  that 
chapter,  we  have  at  once 


**L  =  ZLL(%m\*&  (51) 

ae          /r 

and 

If  we  substitute  €0o  for  e  in  Eqs.  (5.1)  and  (5.2),  we  have  the  quantities 
(dN/de)o  and  (rfW/dc2)0  of  the  previous  section.  To  find  eoo,  we  note 
that  from  Eq.  (1.9),  Chap.  IV,  the  number  of  states  with  energy  less  than 
€  is  (4irV/3Ji8)(2w€)w.  Remembering  that  there  are  just  N  states  with 
energy  less  than  eoo,  this  gives 


We  notice,  as  is  natural,  that  eoo,  the  highest  occupied  energy  level  at  the 
absolute  zero,  increases  as  the  number  of  particles  N  increases.  It  is 
important  to  notice,  however,  that  it  is  the  density  of  particles,  N/V,  that 
is  significant,  not  the  absolute  number  of  particles  in  the  system.  In  a 
gas  obeying  the  Fermi  statistics,  the  particles  cannot  all  have  zero  energy 
at  the  absolute  zero,  as  they  would  in  the  Boltzmann  statistics;  but  since 
there  can  be  only  one  particle  in  each  stationary  state,  there  is  an  energy 
distribution  up  to  the  maximum  energy  eoo.  Let  us  see  how  large  this 
is,  in  actual  magnitude.  We  can  hardly  be  interested  in  cases  where 
N/V  represents  a  density  much  greater  than  the  number  of  atoms  of  a 
solid  per  unit  volume.  Thus  for  example  let  N/V  be  one  in  a  cube  of 
3  X  lO"8  cm.  on  a  side,  or  let  it  equal  &  X  1024.  Let  us  make  the 
calculation  in  kilogram  calories  per  gram  mole  (1  kg.-cal.  equals  1000  cal. 
or  4.185  X  1010  ergs,  one  mole  contains  6.03  X  1023  molecules),  and  let 
us  do  it  first  for  an  atom  of  unit  molecular  weight,  for  which  one  molecule 
weighs  1.66  X  10~~24  gm.  Then  we  have 


6.03  X  1023 

coo  — 


[3  X  (6.61  X  1Q-27)3T 
10-"L    47r  X  27  X  10-24    J 


4.185  X  10102  X  1.66  X  10- 
=  0.081  kg.-cal.  per  gram  mole.  (5.4) 

There  do  not  seem  to  be  any  ordinary  gases  in  which  the  energy  calculated 
from  Eq.  (5.4)  is  appreciable.     Hydrogen  H2  and  helium  He  both  satisfy 


SBC.  5J      THE  FERMI-DIRAC  AND  E1NSTEIN-BOSE  STATISTICS         81 

the  Einstein-Bose  statistics  instead  of  the  Fermi,  and  so  are  not  suitable 
examples.  With  any  heavier  gas,  the  mass  that  comes  in  the  denominator 
of  Eq.  (5.3)  would  reduce  the  value  to  a  few  gram  calories  per  mole,  a 
value  small  compared  to  the  internal  energy  which  the  gas  would  acquire 
in  even  a  few  degrees  with  normal  specific  heat  as  given  by  the  Boltzmann 
statistics,  in  Chap.  IV,  Sec.  3.  The  one  case  where  the  Fermi  statistics 
is  of  great  importance  is  with  the  electron  gas,  on  account  of  the  very 
small  mass  of  the  electron.  The  atomic  weight  of  the  electron  can  be 
taken  to  be  TSTV-  Then  to  get  coo  for  an  electron  gas  of  the  density 
mentioned  above,  we  multiply  the  figure  of  Eq.  (5.4)  by  1813,  obtaining 

€0o  =  148  kg.-cal.  per  gram  mole  =  6.4  electron  volts.         (5.5) 

The  value  (5.5),  instead  of  being  small  in  comparison  with  thermal  magni- 
tudes like  (5.4),  is  of  the  order  of  magnitude  of  large  heats  of  dissociation 
or  ionization  potentials,  and  enormously  large  compared  with  thermal 
energies  at  ordinary  temperatures.  Thus  in  an  electron  gas  at  high  den- 
sity, the  fastest  electrons,  even  at  the  absolute  zero,  will  be  moving  with 
very  high  velocities  and  very  large  energies. 

We  can  next  use  Eq.  (4.5)  to  find  c0  at  other  temperatures.     We  have 
at  once 


From  Eq.  (5.6)  we  note  that  in  ordinary  gases,  where  too  is  of  the  order 
of  magnitude  of  kT  for  a  low  temperature,  the  term  in  712  will  be  large, 
showing  that  the  series  converges  slowly.  On  the  other  hand,  in  an  elec- 
tron gas,  where  eoo  is  very  large  compared  to  fcT,  the  series  converges 
rapidly  at  ordinary  temperatures,  and  €0  is  approximately  independent  of 
temperature,  decreasing  slightly  with  increasing  temperature. 

The  quantity  C/o,  the  internal  energy  at  the  absolute  zero,  is  easily 
found,  from  the  equation 


(5.7) 

using  Eq.  (5.1).  Thus  the  mean  energy  of  a  particle  at  the  absolute  zero 
is  three-fifths  of  the  maximum  energy.  From  Eq.  (4.7)  we  can  then  find 
the  internal  energy  at  any  temperature,  finding 


82  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  V 

The  heat  capacity  is  given  by 

Cv  «  M£—  •  (5.9) 

*  coo 

For  an  electron  gas  at  ordinary  temperature,  this  is  a  small  quantity.  An 
ordinary  gas  would  have  a  heat  capacity  fJVfc.  This  value  is  the  ordinary 
value,  multiplied  by  (7r2/3)(fc!T/€oo).  Now  at  1000°  abs.,  for  instance,  a 
fairly  high  temperature,  kT  is  about  2  kg.-cal.  per  gram  mole,  whereas  we 
have  seen  in  Eq.  (5.5)  that  e0o  is  of  the  order  of  148  kg.-cal.  per  gram  mole. 
Since  ir2/3  is  about  3,  this  means  that  the  electronic  specific  heat  at  1000° 
abs.  is  about  four  per  cent  of  that  of  a  perfect  gas  on  the  Boltzmann 
statistics,  while  at  room  temperature  it  would  be  only  a  little  over  1  per 
cent. 

The  other  thermal  quantities  arc  easily  found.  As  we  see  from 
Eqs.  (4.8)  and  (4.10),  the  entropy  equals  the  specific  heat,  up  to  terms 
linear  in  the  temperature,  so  that  the  entropy  of  the  perfect  gas  is  given 
by  Eq.  (5.9)  at  low  temperatures.  And  the  function  A,  using  Eq.  (4.12), 
is 

Q  ir2  IcT 

A    =   ^€00   -   NkT^-r—'  (5.10) 

O  4  €00 

To  find  the  pressure,  we  wish  to  know  A  explicitly  as  a  function  of  volume. 
Substituting  for  €0o  from  Eq.  (5.3),  we  have 


A  - 

We  can  now  differentiate  to  get  the  pressure: 


-GO 


-  (5-12) 

as  we  can  see  by  substituting  for  €00  in  Eq.  (5.8).  Equation  (5.12),  stating 
that  PV  =  ft/,  is  the  equation  found  in  Eq.  (3.5),  Chap.  IV,  by  a  kinetic 
method,  without  making  any  assumptions  about  the  distribution  of 
velocities.  It  must  therefore  hold  for  the  Fermi  distribution  as  well  as 
for  the  Boltzmann  distribution,  and  it  was  really  not  necessary  to  make  a 
special  calculation  of  the  equation  of  state  at  all.  It  is  obvious,  however, 
that  the  final  equation  of  state  is  very  different  from  that  of  the  perfect 
gas  on  the  Boltzmann  statistics,  on  account  of  the  very  different  relation 
which  we  have  here  between  internal  energy  and  temperature.  Since 


SEC.  6]      THE  FERMI-DIRAC  AND  EINSTEIN-BOSE  STATISTICS         83 

the  internal  energy  is  very  large  at  the  absolute  zero,  but  increases  only 
slowly  with  rising  temperature,  with  a  term  proportional  to  T72,  the  same 
is  true  here  of  PV.  Thus,  in  the  example  used  above,  the  pressure  is 
149,000  atm.  at  the  absolute  zero.  We  note  that,  in  contrast  to  the 
Boltzmann  statistics,  the  internal  energy  here  depends  strongly  on  the 
volume,  as  Eqs.  (5.8)  and  (5.3)  show;  thus  the  gas  does  not  obey  Boyle's 
law.  This  dependence  of  internal  energy  on  volume  is  an  interesting 
thing,  for  it  does  not  indicate  in  any  way  the  existence  of  forces  between 
the  particles,  which  we  are  neglecting  here  just  as  in  the  Boltzmann  theory 
of  a  perfect  gas.  The  kinetic  energy  is  what  depends  on  the  volume,  on 
account  of  the  dependence  of  eoo  on  volume. 

6.  The  Einstein-Bose  Distribution  Law.  —  We  can  find  the  Einstein- 
Bose  distribution  law,  proceeding  by  exact  analogy  with  the  methods  of 
Sec.  3  but  using  the  expression  (2.11)  for  the  entropy.  Thus  for  the 
function  A  we  have 


A  =  2)L/V\«<  +  kTNv  In  Ni  -  kT(\  +  NJ  In  (1  +  ft.)].       ,g  ^ 
i 

Varying  the  NS*  and  requiring  that  ^4  be  a  minimum  for  equilibrium,  we 
have 

dA  =  0  =  ^5(  €.  +  kT  In  TTVV^»-  (6-2) 

^MM    V  1       — T-     TV      / 

As  in  Sec.  3,  Eq.  (6.2)  must  be  satisfied,  subject  to  the  condition 

_    V,  =  0, 

leading  to  the  relation 

et  +  kT  In  -  -j-^r  =  €o  =  const,  (6.3) 

1  ~r  ^Vi 

Solving  for  Ni,  as  in  the  derivation  of  Eq.  (3.5),  we  have 

N*  =  ~K-3 (6.4) 

e  kT    -  1 

Equation  (6.4)  expresses  the  Einstein-Bose  distribution  law.    As  with 
the  Fermi-Dirac  law,  the  constant  co  is  to  be  determined  by  the  condition 


^'  =  2  "^^ 

'  ' 


kT  '   _ 


84 


INTRODUCTION  TO  CHEMICAL  I'/IYtiWS 


[CHAP.  V 


We  can  show,  as  we  did  with  the  Fermi-Dirac  statistics,  that  the 
distribution  (6.5)  approaches  the  Maxwell-Boltzmann  distribution  law  at. 
high  temperatures.  It  is  no  easier  to  make  detailed  calculations  with  the 
Einstein-Bose  law  than  with  tho  Fermi-Dirac  distribution,  and  on  account 
of  its  smaller  practical  importance  we  shall  not  carry  through  a  detailed 


-2 


kT 


FIG.    V-2.  —  Distribution    functions   for    Fcrmi-Dirac   statistics    (a)  ;    Maxwell-Boltzmann 
statistics  (&);  and  lOinstom-Bose  statistics  (f). 

discussion.     It  is  interesting,  however,  to  compare  the  three  distribution 
laws.     This  is  done  in  Fig.  V-2,  where  we  plot  the  function    /^x—  -— 


representing  the  Einstcin-Bose  law,  l/e  kT  representing  the  Maxwell- 
Boltzmann,  and  -(,_«,) representing  the  Fermi-Dirac,  all  as  functions 

•  z.y         i      -t 

e  kl     +1 

of  €.  We  observe  that  the  curve  for  the  Einstein-Boso  distribution 
becomes  asymptotically  infinite  as  c  approaches  eo.  From  this  and  from 
Eq.  (6.5),  it  follows  that  e0  must  lie  lower  than  any  of  the  energy  levels  of  a 
system,  in  contrast  to  the  case  of  the  Fermi-Dirac  distribution.  We  see 
that  the  Maxwell-Boltzmann  distribution  forms  in  a  certain  sense  an 
intermediate  case  between  the  two  other  distributions.  The  Fermi-Dirac 
statistics  tends  to  concentrate  the  molecules  more  in  the  higher  energies, 
having  fewer  molecules  in  proportion  in  the  lower  energies  than  in  the 
Maxwell-Boltzmann  statistics.  On  the  contrary,  the  Einstein-Boso 
statistics  tends  to  have  more  molecules  in  the  lower  energies.  As  a  mat- 
ter of  fact,  more  elaborate  study  of  the  Einstein-Bose  distribution  law 
shows  that  the  concentration  of  molecules  in  the  low  states  is  so  extreme 
that  at  low  enough  temperatures  a  phenomenon  of  condensation  sets 
in,  somewhat  analogous  to  ordinary  changes  of  phase  of  a  real  gas.  From 


SBC.  6J      THE  FERMI-DIRAC  AND  E1NSTEIN-BOSE  STATISTICS          85 

these  properties  of  the  distribution  laws,  we  can  see  that  in  some  super- 
ficial ways  the  effect  of  the  Fermi-Dirac  statistics  is  similar  to  that  of 
repulsive  forces  between  the  molecules,  leading  to  a  large  pressure  even 
at  the  absolute  zero,  while  the  effect  of  the  Einstein-Bose  statistics  is 
similar  to  that  of  attractive  forces,  leading  to  condensation  into  a  phase 
resembling  a  liquid. 

The  real  gases  hydrogen  and  helium  obey  the  Einstein-Bose  statistics, 
and  there  are  indications  that  at  temperatures  of  a  few  degrees  absolute 
the  departures  from  the  Maxwell-Boltzmann  statistics  are  appreciable. 
Of  course,  the  molecules  have  real  attractive  forces,  but  the  effect  of  the 
statistics  is  to  help  those  forces  along,  producing  condensation  at  some- 
what higher  temperature  than  would  otherwise  be  expected.  The  sugges- 
tion has  even  been  made  that  the  anomalous  condensed  phase  He  II,  a 
liquid  persisting  to  the  absolute  zero  at  ordinary  pressures  and  showing 
extraordinarily  low  viscosity,  may  bo  tho  condensed  phase  of  tho  Einstoin- 
Bose  statistics.  For  other  gases  than  hydrogen  and  helium,  the  inter- 
molecular  attractions  are  so  much  greater  than  the  effect  of  tho 
Einstein-Bose  statistics  that  they  liquefy  at  temperatures  too  high  to 
detect  departures  from  the  Maxwell-Boltzmann  law.  Aside  from  these 
gases,  the  only  important  application  of  the  Einstein-Bose  statistics 
eomes  in  the  theory  of  black-body  radiation,  in  which  it  is  found  that 
photons,  or  corpuscles  of  radiant  energy,  obey  the  Einstein-Bose  statistics, 
leading  to  a  simple  connection  between  the  Einstein-Bose  distribution 
law  and  the  Planck  law  of  black-body  radiation,  which  we  shall  discuss 
in  a  later  chapter. 


CHAPTER  VI 

THE  KINETIC  METHOD  AND  THE  APPROACH  TO  THERMAL 

EQUILIBRIUM 

In  the  preceding  chapters,  we  have  taken  up  in  a  very  general  v\ay 
thermodynamics  and  statistical  mechanics,  including  some  applications 
to  perfect  gases.  Both,  as  we  have  seen,  are  very  general  and  powerful 
methods,  hut  both  are  limited,  as  far  as  quantitative  predictions  are 
concerned,  to  systems  in  thermal  equilibrium.  The  kinetic  theory,  some 
of  whose  methods  we  shall  use  in  this  chapter,  is  not  so  limited.  It  can 
handle  the  rates  of  molecular  processes  and  incidentally  treats  thermal 
equilibrium  by  looking  for  a  steady  state,  in  which  the  rate  of  change  of 
any  quantity  is  zero.  But  it  has  disadvantages  compensating  this  groat 
advantage:  it  is  much  more  complicated  and  much  less  general  than 
thermodynamics  and  statistical  mechanics.  For  this  reason  we  shall  not 
pretend  to  give  any  methods  of  handling  an  arbitrary  problem  by  kinetic 
theory.  We  limit  ourselves  to  a  very  special  case,  the  perfect  monatomic 
gas,  and  wo  shall  not  oven  make  any  quantitative  calculations  for  it. 
Later  on,  in  various  parts  of  the  book,  we  shall  handle  other  special 
problems  by  the  kinetic  method.  Always,  we  shall  find  that  an  actual 
calculation  of  the  rate  of  a  process  gives  us  a  better  physical  insight  into 
what  is  going  on  than  the  more  general  methods  of  statistical  mechanics. 
But  generally  we  shall  find  that  the  kinetic  methods  do  not  go  so  far,  and 
always  they  are  more  complicated.  Our  problem  in  this  chapter  is  to 
'  investigate  thermal  equilibrium  in  a  perfect  monatomic  gas  by  the  kinetic 
method.  We  sot  up  an  arbitrary  state  of  a  gas  and  investigate  how  it 
changes  as  time  goes  on.  We  compute  its  entropy  at  each  stage  of  the 
process,  showing  that  in  fact  the  entropy  increases  in  the  irreversible 
process  by  which  the  arbitrary  distribution  changes  over  to  thermal 
equilibrium,  and  we  can  actually  find  how  fast  it  increases,  which  we 
could  not  do  by  our  previous  methods.  Finally  by  looking  for  the  final 
state,  in  which  the  entropy  can  no  longer  increase,  we  get  the  condition 
for  thermal  equilibrium  and  show  that  it  agrees  with  the  condition  derived 
from  statistical  mechanics  and  the  canonical  assembly. 

1.  The  Effect  of  Molecular  Collisions  on  the  Distribution  Function 
in  the  Boltzmann  Statistics. — Let  us  set  up  a  distribution  in  the  molecular 
phase  space,  as  described  in  Chap.  V,  Sec.  1.  We  consider,  not  a  single 
state  of  the  gas,  but  an  assembly  of  states,  as  set  up  in  Sec.  2  of  that  same 

86 


SBC.  1]  THE  KINETIC  METHOD  87 

chapter,  defined  by  the  average  number  JVt  of  molecules  in  the  ith  cell 
of  the  molecular  phase  space,  and  having  the  entropy  given  in  Eq.  (2.7), 
(2.li),  or  (2.12)  of  Chap.  V.  We  start  with  an  arbitrary  set  of  JVVs,  and 
ask  how  they  change  as  time  goes  on.  The  changes  of  $Vs  arise  in  two 
ways.  First,  there  are  thoso  changes  that  would  be  present  even  if  there 
were  no  collisions  between  molecules.  A  molecule  with  a  certain  velocity 
moves  from  point  to  point,  and  hence  from  cell  to  cell,  on  account  of 
that  velocity.  And  if  the  molecule  is  actod  on  by  an  external  force 
field,  which  changes  its  momentum,  it  goes  from  cell  to  cell  for  that 
reason  too.  These  changes  are  in  the  nature  of  streamline  flows  of  the 
representative  points  of  the  molecules  in  the  molecular  phase  space.  Wo 
shall  discuss  them  later  and  shall  show  that  they  do  not  result  in  any 
change  of  entropy.  Secondly,  there  are  changes  of  $Ys  on  account  of 
collisions  between  molecules.  These  are  the  changes  resulting  in  irrever- 
sible approach  to  a  random  distribution  and  in  an  increase  of  entropy. 
Since  they  are  for  our  present  purposes  the  most  interesting  changes,  we 
consider  them  first. 

Consider  two  molecules,  one  in  the  itli  cell,  one  in  thejth,  of  molecular 
phase  space.  If  these  cells  happen  to  correspond  to  the  same  value  of  the 
coordinates,  though  to  different  values  of  the  momenta,  there  is  a  chance 
that  the  molecules  may  collide.  In  the  process  of  collision,  the  represen- 
tative points  of  the  molecules  will  suddenly  shift  to  two  other  cells,  say 
the  kth  and  Zth,  having  practically  the  same  coordinates  but  entirely 
different  momenta.  The  momenta  will  be  related  to  the  initial  values;  for 
the  collision  will  satisfy  the  conditions  of  conservation  of  energy  and 
conservation  of  momentum.  These  relations  give  four  equations  relating 
the  final  momenta  to  the  initial  momenta,  but  since  there  are  six  com- 
ponents of  the  final  momenta  for  the  two  particles,  the  four  equations 
(conservation  of  energy  and  conservation  of  three  components  of  momen- 
tum) will  still  leave  two  quantities  undetermined.  For  instance,  we  may 
consider  that  the  direction  of  one  of  the  particles  after  collision  is  undeter- 
mined, the  other  quantities  being  fixed  by  the  conditions  of  conservation. 

We  now  ask,  how  many  collisions  per  second  are  there  in  which  mole- 
cules in  the  ith  and  jth  cells  disappear  and  reappear  in  the  kth  and  Ith 
cells?  We  can  be  sure  that  this  number  of  collisions  will  be  proportional 
both  to  the  number  of  molecules  in  the  ith  and  to  the  number  of  molecules 
in  the  jth  cell.  This  is  plain,  since  doubling  the  number  of  either  type  of 
molecule  will  give  twice  as  many  of  the  desired  sort  that  can  collide,  and 
so  will  double  the  number  of  collisions  per  unit  time.  In  the  case  of  the 
Boltzmann  statistics,  which  we  first  consider,  the  number  of  collisions  will 
be  independent  of  the  number  of  molecules  in  the  kth  and  Ith  cells,  though 
we  shall  find  later  that  this  is  not  the  case  with  the  Fermi-Dirac  and 
Einstein-Bose  statistics.  We  can  then  write  the  number  of  collisions  of 


88  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

the  desired  type  in  unit  time,  averaged  over  the  assembly,  as 

A^fWi.  (1.1) 

The  coefficient  A\{  will  of  course  depend  on  the  momenta  associated  with 
all  four  cells,  and  in  particular  will  be  zero  if  these  momenta  do  not  satisfy 
tho  conditions  of  conservation.  It  will  also  depend  on  the  properties 
of  the  atom.  For  instance,  it  is  obvious  that  the  larger  the  molecules 
are,  the  more  likely  they  are  to  collide,  and  the  larger  the  A's  will  be.  We 
do  not  have  to  go  more  into  details  of  the  A's  for  our  present  purposes, 
however. 

In  addition  to  these  collisions,  we  shall  have  to  consider  what  we  shall 
call  an  inverse  collision.  This  is  one  in  which  the  molecules  before  colli- 
sion are  in  the  cells  k  and  Z,  and  after  collision  are  in  cells  i  and  j.  The 
number  of  such  collisions  per  unit  time,  by  the  same  argument  as  before, 
will  be 

A"ftjti.  (1.2) 

Now  we  ask,  what  relation,  if  any,  is  there  between  the  two  coefficients 
Atf  and  A1**  of  the  direct  and  inverse  collisions?  The  answer  to  this 
question  is  simple  but  not  very  easy  to  justify.  It  is  this:  if  the  cells  are 
all  of  the  same  size,  as  we  are  assuming,  we  have  simply 

AH  =  A?f  (1.3) 

In  case  the  collision  takes  place  according  to  Newtonian  mechanics,  the 
relation  (1.3)  can  be  proved  by  means  of  Liouville's  theorem.  In  quan- 
tum mechanics,  Eq.  (1.3)  is  practically  one  of  the  postulates  of  the  theory, 
following  directly  from  quantum  mechanical  calculations  of  transition 
probabilities  from  one  state  to  another.  For  our  present  purpose,  con- 
sidering that  this  is  an  elementary  discussion,  we  shall  simply  assume  the 
correctness  of  relation  (1.3).  This  relation  is  sometimes  called  the  prin- 
ciple of  microscopic  reversibility. 

We  are  now  in  position  to  find  how  our  distribution  function  changes 
on  account  of  collisions.  Let  us  consider  a  certain  cell  i}  and  ask  how  the 
average  number  Ni  of  molecules  in  this  cell  changes  with  time,  on  account 
of  collisions.  In  the  first  place,  whenever  a  molecule  in  the  cell  collides 
with  another  molecule  in  any  other  cell,  the  first  molecule  will  be  removed 
from  the  ith  cell,  and  tho  number  of  molecules  in  this  cell  will  be  dimin- 
ished by  one.  But  the  whole  number  of  collisions  of  a  molecule  in  cell 
ij  with  all  other  molecules,  per  second,  is 


where  we  are  summing  over  all  other  types  of  molecule  j  with  which  the 


SBC.  2]  THE  KINETIC  METHOD  89 

original  molecule  can  collide,  and  over  all  possible  states  k  and  I  into  which 
the  molecules  can  be  sent  by  collision.  On  the  other  hand,  it  is  possible 
to  have  a  collision  of  two  molecules  having  quite  different  momenta,  such 
that  one  of  the  molecules  after  collision  would  be  in  cell  i.  This  would 
result  in  an  increase  of  unity  in  the  number  of  molecules  in  the  coll  i.  The 
number  of  such  collisions  per  second  is 


(1.5) 
3ki 

where  we  have  used  the  result  of  Eq.  (1.3).     Thus  the  total  change  in  ffl 
per  second  is  given  by 


(1.6) 


2.  The  Effect  of  Collisions  on  the  Entropy.  —  Equation  (1.6)  represents 
the  first  part  of  our  derivation  of  the  effect  of  collisions  in  producing  an 
irreversible  change  in  the  distribution  and  hence  in  increasing  the  entropy. 
Now  we  must  go  back  to  the  definition  of  the  entropy  in  Eq.  (2.12)  in 
Chap.  V,  and  find  how  much  S  changes  per  unit  time  on  account  of  the 
collisions.  Differentiating  that  equation  with  respect  to  the  time,  we 
have  at  once 

dS 


In  AT.f  '.  (2.1) 


t 
Substituting  from  Eq.  (1.6),  this  becomes 


I*'/  In    AT  (W  N  .   A/.  3v.^  fO  9^ 

,.     —  /v     s-  ,  tljci  ill  IV  i\l\  ilV  j   —  IMjelMi).  \£i.L) 

at          <*** 

We  notice  that  the  fourfold  summation  over  i,  j,  ky  I  is  perfectly  sym- 
metrical in  i  and  j;  they  are  simply  the  indices  of  the  two  colliding  particles 
before  collision.  We  could  interchange  their  names,  and  could  equally 
well  write 


ln  R>(8&  -  #*#')•  (2-3) 


90  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

By  Eq.  (1.3),  this  could  also  be  written 

But  in  Eq.  (2.4),  we  can  interchange  the  names  of  i,  j  with  k,  Z,  obtaining 

dS         ^^ 

dt         <-*— J    kl  l    3 

lyi  In  Nk(fflNj  -  NkNi).  (2.5) 

Finally,  interchanging  the  role  of  the  fcth  and  Ith  atoms,  we  have 
dS  "SJ  , . . ,     i 

Tt  =  -*.^*«ln' 

We  now  have,  in  Eqs.  (2.2),  (2.3),  (2.5),  and  (2.6),  four  equivalent 
ways  of  writing  dS/dt.  We  add  these  equations  and  divide  by  4,  obtain- 
ing the  final  form 

J.Q       ^'%n 

\JUkJ  Iv    ^W  '     A  •  •  /»         -jTV        ,      i         -Tr  i         iTV  i 

__^_   — —  ^^  v4  '  (  ITI  W  •    I     In  7V  •      -   \\\  /Vt      -  ifi 

dt       4-*— J    kl          *  7 


-  ln  N*Ni)(ff<fli  -  ^*^»)-  (2-7) 


The  result  of  Eq.  (2.7)  is  a  very  remarkable  one.  Each  term  of  the 
summation  is  a  product  of  a  coefficient  A  (which  is  necessarily  positive), 
and  a  factor  of  the  form 

(In  z  -  In  T/)  (*-</),  (2.8) 


where  x  =  NiNj,  y  =  NkNi.  But  the  factor  (2.8)  is  necessarily  positive. 
If  x  >  y,  so  that  x  —  y  is  positive,  then  In  x  >  In  y,  so  that  the  other 
factor  In  x  —  In  y  is  positive  as  well.  On  the  other  hand,  if  x  <  y,  so 
that  x  —  y  is  negative,  In  x  —  In  y  is  also  negative,  so  that  the  product 
of  the  two  factors  is  again  positive.  Thus,  every  term  of  the  summation 
(2.7)  is  positive,  and  as  a  result  the  summation  is  positive.  The  only 
way  to  avoid  this  is  to  have  each  separate  term  equal  to  zero;  then  the 
whole  summation  is  zero.  But  if  the  summation  is  positive,  this  means 
that  the  entropy  S  is  increasing  with  time.  Thus  we  have  proved  Boltz- 
mann's  famous  theorem  (often  called  the  H  theorem,  because  he  called 
the  summation  of  Eq.  (2.12),  Chap.  V,  by  the  symbol  H,  setting 
S  =  —kH)  :  the  entropy  S  continually  increases,  on  account  of  collisions, 


SBC.  2]  THE  KINETIC  METHOD  91 

unless  it  has  already  reached  a  steady  state,  for  which  the  condition  is 

N.Nj  -  RkNt  -  0  (2.9) 

for  every  set  of  cells  i, ,7,  &,  I  between  which  a  collision  is  possible  (that 
is,  for  which  Aft  ^  0). 

By  comparison  with  Eq.  (1.6),  we  see  that  Eq.  (2.9)  leads  to  the 
condition  that  dN%/dt  should  be  zero,  by  demanding  that  each  separate 
term  of  Eq.  (1.6)  should  be  zero.  That  is,  in  equilibrium,  the  collision 
in  which  atoms  i  and  j  collide  to  give  atoms  k  and  Z,  together  with  the 
inverse  to  this  type  of  collision,  by  themselves  give  no  net  change  in  the 
numbers  of  atoms  in  the  various  states,  the  number  of  direct  collisions 
just  balancing  the  number  of  inverse  collisions.  This  condition  is  called 
the  condition  of  detailed  balancing.  It  is  a  general  characteristic  of 
thermal  equilibrium  that  this  detailed  balancing  should  hold  and,  as  we 
have  seen,  it  follows  directly  from  the  second  law  in  its  statistical  form. 

We  may  now  rewrite  Eq.  (2.9)  in  the  form 

fftffi  =  NkNi, 
or 

In  #t  +  In  Nj  =  In  Nk  +  In  NI.  (2.10) 

This  holds  for  every  transition  for  which  A&  7*  0;  that  is,  for  every  colli- 
sion satisfying  the  laws  of  conservation  of  energy  and  momentum.  Using 
the  notation  of  Sec.  3  in  Chap.  IV,  we  can  let  the  average  number  of 
molecules  in  an  element  dx  dy  dz  dpx  dpy  dpz  of  the  molecular  phase  space 
be/m  dx  dy  dz  dpx  dpy  dpz.  According  to  Chap.  Ill,  Sec.  3,  the  volume  of 
molecular  phase  space  associated  with  one  cell  is  A8.  Then  we  have  the 
relation 

ffi  =  h*fm,  (2.11) 

where /m  is  to  be  computed  in  the  ith  cell.  We  now  substitute  Eq.  (2.11) 
in  Eq.  (2.10),  writing  that  equation  in  terms  of  the/m's.  Since  all  four  of 
our  cells  must  refer  to  the  same  point  of  coordinate  space,  since  molecules 
cannot  collide  unless  they  are  in  contact,  we  write  fm  merely  as 
fm(pxpypz),  or/(p)  for  short.  Then  we  have 

ln/(p,)  +  ln/(p/)  =  ln/(p*)  +  ln/(p«).  (2.12) 

Equation  (2.12)  states  that  there  is  a  certain  function  In /of  the  momen- 
tum of  a  molecule,  such  that  the  sum  of  the  functions  of  the  two  molecules 
before  collision  equals  the  sum  of  the  functions  of  the  two  after  collision. 
That  is,  the  total  amount  of  this  function  is  conserved  on  collision.  But 
there  are  just  four  quantities  that  have  this  property:  the  energy  and  the 
three  components  of  momentum.  Any  linear  function  of  these  four 
quantities  will  also  be  conserved,  and  it  can  be  proved  that  this  is  the  most 


92  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

general  function  which  has  this  property.     Thus  we  may  conclude  that 

In  /(p)  »  A^  +  Bpx  +  CPy  +  Dpz  +  E,  (2.13) 

where  A,  B,  C,  Z),  E  are  arbitrary  constants.  Substitution  of  Eq.  (2.13) 
in  Eq.  (2.12)  shows  at  once  that  Eq.  (2.12)  is  satisfied.  We  have  been 
able,  in  other  words,  to  get  a  general  solution  of  the  problem  of  the  dis- 
tribution function  in  equilibrium. 

The  linear  combination  of  Eq.  (2.13)  can  be  rewritten  in  the  form 

[(PX  ~~  Pxo)*  +  (PV  ""  Pvo)2  +  (PZ  "  p§o)f] 


.  _ 
2m  /       "  w*y          roM'    "      mkT 


where  T  (later  to  be  identified  with  the  temperature),  px0,  py0,  p«o,  /o  are 
arbitrary  constants,  whose  relation  to  the  constants  of  Eq.  (2.13)  is  evi- 
dent from  Eq.  (2.14).  Thus  we  have 

-((p»-pxo)2  +  (7>v-pyo)2  +  (pe-p.o)2)  ,  ex 

/(p)  =/oe  ~~2^r~        "  (2.15) 

3.  The  Constants  in  the  Distribution  Function.  —  In  Eq.  (2.15),  we 
have  found  a  distribution  function  satisfying  the  condition  of  thermal 
equilibrium  and  containing  five  arbitrary  constants,  T,  px0,  pyQj  pzQ,  /0. 
Since  our  calculation  has  been  entirely  for  a  single  point  of  ordinary  space, 
these  five  quantities,  for  all  we  know,  may  vary  from  point  to  point  or  be 
functions  of  position.  Shortly  we  shall  find  how  the  quantities  must  vary 
with  position  in  order  to  have  thermal  equilibrium.  We  may  anticipate 
by  stating  the  results  which  we  shall  find  and  giving  their  physical 
interpretation. 

In  the  first  place,  we  shall  find  that  the  four  quantities  T,  px^  pv0,  pzQ 
must  be  constant  at  all  points  of  space,  for  equilibrium.  By  comparison 
with  Eq.  (2.4)  of  Chap.  IV,  the  formula  for  the  Maxwell  distribution  of 
velocities,  we  see  that  T  must  be  identified  with  the  temperature,  which 
must  not  vary  from  point  to  point  in  thermal  equilibrium.  The  quan- 
tities PXO,  pVQ,  pZQ  are  the  components  of  a  vector  representing  the  mean 
momentum  of  all  the  molecules.  If  they  are  zero,  the  distribution  (2.15) 
agrees  exactly  with  Eq.  (2.4)  of  Chap.  IV.  If  they  are  not  zero,  however, 
Eq.  (2.15)  represents  the  distribution  of  velocities  in  a  gas  with  a  certain 
velocity  of  mass  motion,  of  components  pXQ/m,  p^o/w,  pzo/in.  The  quan- 
tities px  —  p*o,  etc.,  represent  components  of  momentum  relative  to  this 
momentum  of  mass  motion,  and  the  relative  distribution  of  velocities  is  as 


SBC.  3]  THE  KINETIC  METHOD  93 

in  Maxwell's  distribution.  Since  ordinarily  we  deal  with  gas  without 
mass  motion,  we  ordinarily  set  pxQ,  pyQ,  and  p*o  equal  to  zero.  In  general, 
we  shall  not  have  thermal  equilibrium  unless  the  velocity  of  mass  motion 
is  independent  of  position;  otherwise,  as  we  can  seo  physically,  there 
would  be  the  possibility  of  viscous  effects  between  different  parts  of  the 
gas.  We  have  now  considered  the  variation  of  77,  pxo,  pyQ,  pz0  with  posi- 
tion and  have  shown  that  they  are  constants.  Finally  wo  consider  /0. 
Our  analysis  will  show  that  if  the  potential  energy  of  a  molecule  is  <£,  a 
function  of  position,  we  must  have 

—  tf> 
/o  =  const.  X  ek~f.  '  (3-1) 

Equation  (3.1)  shows  that  the  density  varies  with  position  just  as 
described  in  Chap.  IV,  Sec.  4.  Thus,  with  these  interpretations  of  the 
constants,  we  see  that  P]q.  (2.15),  representing  the  distribution  function 
which  we  find  by  the  kinetic  method  for  the  bteady  state  distribution,  is 
exactly  the  same  that  we  found  previously,  in  Chap.  IV,  by  statistical 
methods. 

We  shall  now  prove  the  results  that  we  have  mentioned  above,  regard- 
ing the  variation  of  T,  px0,  py^  pzQ,  /0  with  position.  We  stated  in  Sec.  1 
that  a  molecule  could  shift  from  point  to  point  in  phase  spare  not  only  on 
account  of  collisions,  which  we  have  considered,  but  also  on  account  of  the 
velocity  and  of  external  force  fields,  and  that  these  shifts  were  in  the 
nature  of  streamline  flows  in  the  phase  space  and  did  not  correspond  to 
changes  of  entropy.  We  must  now  analyze  these  motions  of  the  mole- 
cules. We  shall  assume  classical  mechanics  for  this  purpose;  the  energy 
levels  of  a  perfect  gas,  as  we  have  seen  in  Chap.  IV,  Sec.  1,  are  spaced  so 
closely  together  that  the  cells  in  phase  space  can  be  treated  as  continuous. 
Let  us,  then,  take  a  volume  element  dx  dy  dz  dpx  dpv  dpz  in  molecular 
phase  space,  and  find  the  time  rate  of  change  of  fm  dx  dy  dz  dpx  dpy  dpg, 
the  number  of  molecules  in  the  volume  clement,  for  all  reasons  except 
collisions.  First,  we  consider  the  number  of  molecules  entering  the 
element  over  the  surface  perpendicular  to  the  x  axis,  whose  (five-dimen- 
sional) area  is  dy  dz  dpx  dpy  dpz.  The  component  of  velocity  of  each 
molecule  along  the  x  axis  is  px/m.  Then,  using  an  argument  similar 
to  that  of  Chap.  IV,  Sec.  3,  the  number  of  molecules  entering  the  element 
over  this  surface  per  second  is  the  number  contained  in  a  prism  of  base 
dy  dz  dpx  dpy  dpz  and  altitude  px/m.  This  is  the  volume  of  the  prism 
[(Px/rn)dy  dz  dpx  dpy  dpz]  times  the  number  of  molecules  per  unit  volume 
in  phase  space,  or  /m.  Hence  the  required  number  is 


(P.\ 

•w 


\dy  dz  dpx  dpy  dpg. 


94  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

But  a  similar  number  of  molecules  will  leave  over  the  face  at  x  +  dx. 
The  only  difference  will  be  that  in  computing  this  number  we  must  use 
fm  computed  at  x  +  dx,  or  must  use 

fm(x  +  dx,  yzpxpypg)  =  fm(xyzpxpypz)  +  dx-J£(xyzpxpvpt)  +  -  -  •  , 

the  first  two  terms  of  a  Taylor's  expansion.  Thus  the  net  number  enter- 
ing over  the  two  parallel  faces  perpendicular  to  x  is 


y  dz  dpx  dpv  dpz, 

fit     uX 

and  for  the  three  sets  of  faces  perpendicular  to  x,  y,  z,  we  have  the  net 
number 

-fe  -/;  +  5  £  +  £  If)"*  *  *  *•  <"»  *-     <M> 

In  a  similar  way  we  can  consider  the  face  perpendicular  to  the  px  axis  in 
the  phase  space.  The  component  of  velocity  of  a  representative  point 
along  the  px  axis  in  this  space  is  by  definition  simply  the  time  rate  of 
change  of  px;  that  is,  by  Newton  's  second  law,  it  is  the  x  component  of 
force  acting  on  a  molecule.  If  the  potential  energy  of  a  molecule  is  <£, 
this  component  of  force  is  —d<t>/djc.  Thus,  the  number  of  molecules 
entering  over  the  face  perpendicular  to  the  px  axis  is 


/ml  -fa)dx  d1l  dz  dPy  dP*> 
and  the  net  number  entering  over  the  faces  at  px  and  at  px  +  dpx  is 

f^JL      3f 

~dx  dp    X  dy  <iZ  dpx  dj)y  flp" 

We  have  three  such  terms  for  the  three  components  of  momentum,  and 
combining  with  Eq.  (3.2),  we  have  for  the  total  change  of  the  number  of 
molecules  in  the  volume  element  per  second  the  relation 


*-   <3-8) 

Having  found  the  change  in  the  distribution  function,  in  Eq.  (3.3),  we 
shall  first  show  that  it  involves  no  change  of  entropy.  The  physical 
reason  is  that  it  corresponds  to  a  streamline  motion  in  phase  space,  result- 
ing in  no  increase  of  randomness.  We  use  Eq.  (2.1)  for  the  change  of 


SEC.  3] 


THE  KINETIC  METHOD 


entropy  with  time,  Eq.  (2.11)  for  the  relation  between 
Eq.  (3.3)  for  dfm/dt.    Then  we  have 


and  /„,,  and 


\dx  •  dp9.     (3.4) 


Each  of  the  integrals  over  the  coordinates  is  to  be  carried  to  a  point  out- 
side the  container  holding  the  gas,  each  integral  over  momenta  to  infinity. 
In  Eq.  (3.4),  each  term  can  be  integrated  with  respect  to  one  of  the  vari- 
ables. Thus  the  first  term,  as  far  as  the  integration  with  respect  to  x  is 
concerned,  can  be  transformed  by  the  relation 


f 

•  ' 


=       Infndfm  =  (/wln/w-, 


(3.5) 


At  both  limits  of  integration,  fm  =  0,  since  the  limits  lie  outside  the  con- 
tainer, so  that  the  integral  vanishes.  A  similar  transformation  can  be 
made  on  each  term,  leading  to  the  result  that  the  changes  of  fm  we  are  now 
considering  result  in  no  change  of  entropy.  This  justifies  our  analysis  of 
Sec.  2,  in  which  we  treated  the  change  of  entropy  as  arising  entirely  from 
collisions. 

Now  we  can  use  our  condition  (3.3)  to  find  the  variation  of  our  quanti- 
ties /o,  T,  pxo,  py0,  pz()  of  Sec.  2  with  position.  In  thermal  equilibrium,  we 
must  have  dfm/dt  =  0.  Thus  Eq.  (3.3)  gives  us  a  relation  involving  the 
various  derivatives  of  fm.  We  substitute  for/m  from  Eq.  (2.15),  treating 
the  quantities  just  mentioned  as  functions  of  position.  Then  Eq.  (3.3) 
becomes,  canceling  the  exponential, 


0  =  —  £l?^L0 
m  dx 

/o 


P*: 
m 


Pzdfo 
m  dz 

dpXQ 


-F + <"•  -  *•> «, 


da- 


P 


+  P. 


^ +IP.-  r^ 


(Px  ~ 


tg + (p,  _  ^ 


/o 


~  P«o)2  +  (py  - 


+  (Pz  - 


dT 


dz 


dT 


-  P-O    +  (PV  "  p*o)    +  (p*  "  p*o)       (3-6) 


Equation  (3.6)  must  be  satisfied  for  any  arbitrary  values  of  the  momenta. 
Since  it  is  a  polynomial  in  px,  pv)  p«,  involving  terms  of  all  degrees  up  to 


96  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

the  third,  and  since  a  polynomial  cannot  be  zero  for  all  values  of  its 
argument  unless  the  coefficient  of  each  term  is  zero,  we  can  conclude  that 
the  coefficient  of  each  power  of  each  of  the  p's  in  Eq.  (3.6)  must  be  zero. 
From  the  third  powers  we  see  at  once  that 

dT  -  i\          —  -  c\          dT  -  ft  tt  7^ 

te  ~  °'       ay  "  0|      °ei  ~  °'  (3'7) 

or  the  temperature  is  independent  of  position.  From  the  second  powers 
we  then  see  that  the  derivatives  of  the  form  dpxQ/dx  must  all  be  zero,  or 
the  average  momentum  is  independent  of  position.  We  are  left  with  only 
the  first  and  zero  powers.  From  the  zero  powers,  we  sec  that  either 


PxO    =    PV0    =   PzQ    =   0, 

or 

d<t>        d<t>        d<t>  _ 


,-  ~v 

(3-8) 


That  is,  if  there  is  an  external  force  field,  there  can  be  no  mass  motion  of 
the  gas,  for  in  this  case  the  external  field  would  do  work  on  the  gas  and  its 
energy  could  not  be  constant.  Then  we  are  left  with  the  first  power 
terms.  The  coefficient  of  px,  for  instance,  gives 

fadx   =  ~~kfdx'  '3'9' 

with  similar  relations  for  the  y  and  z  components.  Equation  (3.9)  can 
be  rewritten 

"-* 


dx  dx     ' 

In  /o  =  ^  +  const., 

-» 
/o  =  const.  ekT , 

or  Eq.  (3.1).  Thus  we  have  proved  all  the  results  regarding  our  distribu- 
tion function  that  we  have  mentioned  earlier  in  the  section,  and  have 
completed  the  proof  that  the  Maxwell-Boltzmann  distribution  law  is  the 
only  one  that  will  not  be  affected  by  collisions  or  the  natural  motions  of 
the  molecules  and,  therefore,  must  correspond  to  thermal  equilibrium. 

4.  The  Kinetic  Method  for  Fermi -Dirac  and  Einstein -Bose  Statistics. 
The  arguments  of  the  preceding  sections  must  be  modified  in  only  two 
ways  to  change  from  the  Boltzmann  statistics  to  the  Fermi-Dirac  or 
Einstein-Bose  statistics.  In  the  first  place,  the  law  giving  the  number 
of  collisions  per  unit  time,  Eq.  (1.1),  must  be  changed.  Secondly,  as 


SBC.  4]  THE  KINETIC  METHOD  97 

we  should  naturally  expect,  we  must  use  the  appropriate  formula  for 
entropy  with  each  type  of  statistics.  First,  we  consider  the  substitute 
for  the  law  of  collisions.  Clearly  the  law  (1.1),  giving  AHftJtj  collisions 
per  second  in  which  molecules  in  states  i  and  j  collide  to  give  molecules  in 
states  k  and  I  cannot  be  correct  for  Fermi-Dirac  statistics.  For  the 
fundamental  feature  of  Fermi-Dirac  statistics  is  that  if  the  fcth  or  Ith 
stationary  states  happen  to  be  already  occupied  by  a  particle,  there  is  no 
chance  of  another  particle  going  into  them.  Thus  our  probability  must 
depend  in  some  way  on  the  number  of  particles  in  the  kill  and  Ith  states, 
as  well  as  the  ith  and  jth.  Of  course,  the  kth  state  can  have  either  no 
particles  in  it,  or  one;  never  more.  Thus  in  one  example  of  our  system, 
chosen  from  the  statistical  assembly,  Nk  may  be  zero  or  unity.  If  it  is 
zero,  there  is  no  objection  to  another  particle  entering  it.  If  it  is  unity, 
there  is  no  possibility  that  another  particle  can  enter  it.  Averaging  over 
the  assembly,  the  probability  of  having  a  collision  in  which  a  particle 
is  knocked  into  the  fcth  state  must  clearly  have  an  additional  factor  equal 
to  the  fraction  of  all  examples  of  the  assembly  in  which  the  kth  state  is 
unoccupied.  Now  Nk  is  the  mean  number  of  particles  in  the  kth  state. 
Since  the  number  of  particles  is  always  zero  or  one,  this  means  that  Nk 
is  just  the  fraction  of  examples  in  which  the  fcth  state  is  occupied.  Then 
1  —  Nk  is  the  fraction  of  examples  in  which  it  is  unoccupied,  and  this  is 
just  the  factor  we  were  looking  for.  Similarly  we  want  a  factor  1  —  Nt 
to  represent  the  probability  that  the  Zth  state  will  bo  unoccupied  and 
available  for  a  particle  to  enter  it.  Then,  finally,  we  have  for  the  number 
of  collisions  per  second  in  which  particles  in  the  ith  and  jth  cells  are 
knocked  into  the  fcth  and  Ith  cells,  the  formula 

AllftJt,(l  -  #*)(!  -  tf,).  (4.1) 

In  the  Einstein-Bose  statistics,  there  is  no  such  clear  physical  way 
to  find  the  revised  law  of  collisions  as  in  the  Fermi-Dirac  statistics.  The 
law  can  be  derived  from  the  quantum  theory  but  not  in  a  simple  enough 
way  to  describe  here.  In  contrast  to  the  Fermi-Dirac  statistics,  in  which 
the  presence  of  one  molecule  in  a  cell  prevents  another  from  entering  the 
same  cell,  the  situation  with  the  Einstein-Bose  statistics  is  that  the 
presence  of  a  molecule  in  a  cell  increases  the  probability  that  another 
one  should  enter  the  same  cell.  In  fact,  the  number  of  molecules  going 
into  the  fcth  cell  per  second  turns  out  to  have  a  factor  (1  +  $*),  increasing 
linearly  with  the  mean  number  ffk  of  molecules  in  that  cell.  Thus,  the 
law  of  collisions  for  the  Einstein-Bose  statistics  is  just  like  Eq.  (4.1),  only 
with  +  signs  replacing  the  —  signs.  In  fact,  we  may  write  the  law  of 
collisions  for  both  forms  of  statistics  in  the  form 

ffi),  (4.2) 


98  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

where  the  upper  sign  refers  to  the  Einstein-Bose  statistics,  the  lower  to 
the  Fermi-Dirac. 

Next,  we  must  consider  the  change  of  the  mean  number  of  molecules 
in  the  fth  state,  with  time.  Using  the  law  of  collisions  (4.2)  and  proceed- 
ing as  in  the  derivation  of  Eq.  (1.6),  we  have  at  once 


±  Ni}(l  ±  Nj}  -  M(1  ±  ^)(1  ±  Ni)]-  (4-3) 


Having  found  the  number  of  collisions,  we  can  find  the  change  in  entropy 
per  unit  time.  Using  the  formulas  (2.7)  and  (2.11)  of  Chap.  V  for  the 
entropy  in  the  case  of  Fermi-Dirac  and  Einstein-Bose  statistics,  we  find  at, 
once  that 


where  again  the  upper  sign  refers  to  Einstein-Bose  statistics,  the  lower 
one  to  Fermi-Dirac  statistics.     Substituting  from  Eq.  (4.3),  we  have 


AHlhi  B*  -  ln 
ijkl 


±  #*)(!  ±  Ni)].     (4.5) 


As  in  Sec.  2,  we  can  write  four  expressions  equivalent  to  Eq.  (4.5),  by 
interchanging  the  various  indices  t,  jf,  fc,  I.  Adding  these  four  and  divid- 
ing by  four,  we  obtain 

2?  =  £§^H{  In  [#*#i(i  ±  #0(1  ±  #0]  -  ln[ftjt,(l  ±  #*) 

Hid 
(1  ±  Ni)]}[N*Nl(\  ±  #.)(!  ±  #0  -  ftjt,(l  ±  N.)(l  ±  Ni)].     (4.6) 


But,  as  in  Sec.  2,  this  expression  cannot  be  zero  as  it  must  be  for  a  steady 
state,  unless 


NkNi(l  ±  #0(1  ±  Ni)  =  NiNj(l  ±  Nk)(l  ±  #0, 

and  if  it  is  not  zero,  it  must  necessarily  be  positive.  Thus  we  have  demon- 
strated that  the  entropy  increases  in  an  irreversible  process,  and  have 
found  the  condition  for  thermal  equilibrium. 

From  Eq.  (4.7)  we  can  find  the  distribution  functions  for  the  Einstein- 
Bose  and  Fermi-Dirac  statistics.     We  rewrite  the  equation  in  the  form 


±  #0 


Sue.  41  THE  KINETIC  METHOD  99 

or 

N 
As  in  Sec.  2,  we  may  now  conclude  that  the  quantity  In —  must  be 

(1  ±  N) 

a  function  which  is  conserved  on  collision,  since  the  sum  for  the  two  par- 
ticles before  and  after  collision  is  constant.  And  as  in  that  section,  this 
quantity  must  be  a  linear  combination  of  the  kinetic  energy  and  the 
momentum,  the  coefficients  in  general  depending  on  position.  Also,  as 
in  that  section,  the  momentum  really  contributes  nothing  to  the  result, 
implying  merely  the  possibility  of  choosing  an  arbitrary  average  velocity 
for  the  particles.  Neglecting  this,  we  then  have 

In  -^~  =  a  +  6ekin,  (4.9) 

where  a  and  6  are  constants,  €km  is  the  kinetic  energy  of  a  molecule  in  the 
ith  cell.  That  is,  we  have 

ft* 


As  we  see  by  comparison  with  the  formulas  (3.5)  and  (6.4)  of  Chap.  V, 
the  quantity  b  is  to  be  identified  with  —1/kT.     Thus  we  have 

*'  =  — i-  •—  (4'U) 

e  aekT   +  1 

In  formula  (4.11),  as  in  (2.15),  there  are  certain  quantities  a  and  T 
which  are  constant  as  far  as  the  momenta  are  concerned,  but  which  might 
vary  from  point  to  point  of  space.  We  can  investigate  their  variation 
just  as  we  did  for  the  Boltzmann  statistics  in  Sec.  3.  The  formula  (3.3) 
for  the  change  of  the  distribution  function  with  time  on  account  of  the 
action  of  external  forces  holds  for  the  Einstein-Bose  and  Fermi-Dirac 
statistics  just  as  for  the  Boltzmann  statistics,  and  leads  to  a  formula  very 
similar  to  Eq.  (3.6)  which  must  be  satisfied  for  equilibrium.  The  only 
difference  comes  on  account  of  the  different  form  in  which  we  have 
expressed  the  constants  in  Eq.  (4.11).  Demanding  as  before  that  the 
relation  like  (3.6)  must  hold  independent  of  momenta,  we  find  that  the 
temperature  must  be  independent  of  position,  and  that  the  constant  a 
of  Eq.  (4.11)  must  be  given  by 


100  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VI 

a  -  ^t  (4.12) 

where  eo  is  a  constant,  </>  is  the  potential  energy  of  molecule.     Thus, 
finally,  we  have 

N*  =  "3        '  (4.13) 


where  c»  is  the  total  energy,  kinetic  and  potential,  of  a  molecule  in  the 
ith  cell,  in  agreement  with  Eqs.  (3.5)  and  (6.4)  of  Chap.  V. 


CHAPTER  VII 

FLUCTUATIONS 

A  statistical  assembly  contains  many  replicas  of  the  same  system, 
agreeing  in  large-scale  properties  but  varying  in  small-scale  properties. 
Sometimes  these  variations,  or  fluctuations,  are  important.  Thus,  two 
repetitions  of  the  same  experiment  may  disclose  different  densities  of  a 
gas  at  a  given  point,  though  the  average  density  through  a  large  volume 
may  be  the  same  in  each  case.  Such  fluctuations  of  density  can  be  of 
experimental  importance  in  such  problems  as  the  scattering  of  light,  which 
is  produced  by  irregularities  of  density.  Again,  in  the  emission  of  elec- 
trons from  a  hot  filament,  there  are  fluctuations  of  current,  which  are 
observable  as  the  shot  effect  and  which  are  of  great  practical  importance 
in  the  design  of  amplifying  circuits.  We  shall  take  up  some  of  the  simpler 
sorts  of  fluctuations  in  this  chapter.  We  begin  by  considering  the  fluc- 
tuations of  energy  in  a  canonical  assembly.  We  recall,  from  the  argu- 
ments of  Chap.  Ill,  Sec.  5,  that  an  assembly  of  systems  in  equilibrium 
with  a  temperature  bath  must  be  assumed  to  have  a  variety  of  energies, 
since  they  can  interchange  energy  with  the  bath.  We  can  now  show, 
however,  that  actually  the  great  majority  of  the  systems  have  an  energy 
extremely  close  to  a  certain  mean  value  and  that  deviations  from  this 
mean  are  extremely  small  in  comparison  with  the  total  energy.  This  can 
be  shown  by  a  perfectly  straightforward  application  of  the  distribution 
function  for  the  canonical  assembly,  in  the  case  w^here  our  system  is  a 
sample  of  perfect  gas  obeying  the  Boltzmarm  statistics,  and  we  start  with 
that  example. 

1.  Energy  Fluctuations  in  the  Canonical  Assembly.  —  Let  E  be  the 
energy  of  a  particular  system  in  the  canonical  assembly,  U  being  the 
average  energy  over  the  assembly,  or  the  internal  energy.  We  are  now 
interested  in  finding  how  much  the  energies  E  of  the  individual  systems 
fluctuate  from  their  average  value.  The  easiest  way  to  find  this  is  to 
compute  the  mean  square  deviation  of  the  energy  from  its  mean,  or 
(E  —  £7)2.  This  can  be  found  by  elementary  methods  from  the  Maxwell- 
Boltzmann  distribution  law.  Referring  to  Eq.  (1.1)  of  Chap.  IV,  we 
can  write  the  energy  as 

N 

E  =    ««,  (i.D 


1-1 
101 


102  INTRODUCTION  TO  CHEMICAL  PHYSICS          (CHAP.  VII 

where  «(i)  is  the  energy  of  the  ith  molecule,  and 

U  =  2««  (1.2) 


where  c(t)  is  the  average  energy  of  the  ith  molecule  over  the  assembly. 
Thus  we  have 


N 


(E-U)  =  ]£(€<•>  -€<«),  (1.3) 

t-i 
and 


N     N 

(?»  -  e(i))(7»  -  €<»).  (1.4) 


We  must  now  perform  the  averaging  in  Eq.  (1.4).  We  note  that 
there  are  two  sorts  of  terms:  first,  those  for  which  i  =  j;  secondly,  those 
for  which  i  7^  j.  We  shall  now  show  that  the  terms  of  the  second  sort 
average  to  zero.  The  reason  is  the  statistical  independence  of  two  mole- 
cules i  and  j  in  the  Boltzmann  distribution.  To  find  the  average  of  such 
a  term,  we  multiply  by  the  fraction  of  all  systems  of  the  assembly  in  which 
the  fth  and  jth  molecules  have  tho  particular  energies  c(£  and  ej^,  where 
kl  and  &,  are  indices  referring  to  particular  cells  in  phase  space,  and  sum 
over  all  states  of  the  assembly.  From  Eq.  (1.2)  of  Chap.  IV,  giving  the 
fraction  of  all  systems  of  the  assembly  in  which  each  particular  molecule, 
as  the  ith,  is  in  a  particular  state,  as  the  &»th,  we  see  that  this  average  is 


(€«    -   €<«)(€<*>    -    €<'>) 


k, 


J*  hi 

=   (^>""-niW)(?jrirgO>)  =  (gW  _  e<»>)(eC»)  -  g(0)  =  0.      (1.5) 

Having  eliminated  the  terms  of  Eq.  (1.4)  for  which  i  ^  j,  we  have  left  only 


N 


(E  -  C/)2"  =         (e<'>  -!<•">)*.  (1.6) 


That  is,  the  mean  square  deviation  of  the  energy  from  its  mean  equals 
the  sum  of  the  mean  square  deviation  of  the  energies  of  the  separate 
molecules  from  their  means.  Each  molecule  on  the  average  is  like  every 
other,  so  that  the  terms  in  the  summation  (1.6)  are  all  equal,  and  we  may 


SEC.  1]  FLUCTUATIONS  103 

write 


(E  -  £7)2  =  tflc  -  O2,  (1.7) 

where  «  represents  the  energy  of  a  single  molecule,  i  its  mean  value. 

We  can  understand  Eq.  (1 .7)  better  by  putting  it  in  a  slightly  different 
form.  We  divide  the  equation  by  f72,  so  that  it  represents  the  fractional 
deviation  of  the  energy  from  the  mean,  squared,  and  averaged.  In 
computing  this,  we  use  Eq.  (1.2)  but  note  that  the  mean  energy  of  each 
molecule  is  equal,  so  that  Eq.  (1.2)  becomes 

V  =  Nl.  (1.8) 

Using  Eq.  (1.8),  we  then  have 


Equation  (1.9)  is  a  very  significant  result.  It  states  that  the  fractional 
mean  square  deviation  of  energy  for  N  molecules  is  1/JVth  of  that  for  a 
single  molecule,  in  Boltzmann  statistics.  The  greater  the  number  of 
molecules,  in  othor  words,  the  less  in  proportion  are  the  fluctuations  of 
energy  from  the  mean  value.  The  fractional  deviation  of  energy  of  a 
single  molecule  from  its  mean  is  of  the  order  of  magnitude  of  its  total 
energy,  as  we  can  see  from  the  wide  divergence  of  energies  of  different 
molecules  to  be  observed  in  the  Maxwell  distribution  of  velocities  and  a.s 
we  shall  prove  in  the  next  paragraph.  Thus  the  right  side  of  Eq.  (1.9) 
is  of  the  order  of  magnitude  of  1/N.  If  N  is  of  the  order  of  magnitude  of 
1024,  as  with  a  large  scale  sample  of  gas,  this  means  that  practically  all 
systems  of  the  assembly  have  energies  departing  from  the  mean  by  some- 
thing whose  square  is  of  the  order  of  10~24  of  the  total  energy,  so  that  the 
average  deviation  is  of  the  order  of  10~12  of  the  total  energy.  In  other 
words,  the  fluctuations  of  energy  in  a  canonical  assembly  are  so  small 
as  to  be  completely  negligible,  so  long  as  we  are  dealing  with  a  sample  of 
macroscopic  size. 

To  evaluate  the  fluctuations  of  Eq.  (1.9)  exactly,  we  must  find  the 
fluctuations  of  energy  of  a  single  molecule.     We  have 


(6  -  g)2  =  5  _  2c€  +  (e)2 

=  T*  -  (i)2,  (1.10) 

a  relation  which  is  often  useful.  We  must  find  the  mean  square  energy  of 
a  single  molecule.  Using  the  distribution  function  (2.4)  or  (2.6)  of  Chap. 
IV,  we  find  easily  that 

(1.11) 


104  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VII 

Remembering  that 

i  =  ikT,  (1.12) 

as  shown  in  Eq.  (2.7)  of  Chap.  IV,  this  gives 


(e  -  €-)2  =  (v  -  |)(fcr)«  =  |(*r)',  (1-13) 

from  which  Eq.  (1.9)  becomes 


(E-u\_2i_  a  14) 

\~U~J    ~3AT'  (L14) 

fixing  the  numerical  value  of  the  relative  mean  square  deviation  of  the 
energy. 

2.  Distribution  Functions  for  Fluctuations. — In  the  preceding  section 
we  have  given  an  elementary  derivation  of  the  energy  fluctuations  in 
Boltzmann  statistics.  The  derivation  we  have  used  is  not  applicable  to 
many  interesting  fluctuation  problems,  and  in  the  present  section  we  shall 
develop  a  much  more  general  method.  Suppose  we  have  a  quantity 
x,  in  whose  fluctuations  from  the  mean  we  are  interested.  This  may  be 
the  energy,  as  in  the  last  section,  or  many  other  possible  quantities.  Our 
method  will  be  to  set  up  a  distribution  function /(x),  such  that  f(x)djc 
gives  the  fraction  of  all  systems  of  the  assembly  for  which  x  lies  in  the 
range  dx.  Then  it  is  a  simple  matter  of  integration  to  find  the  mean  of 
any  function  of  #,  and  in  particular  to  find  the  mean  square  deviation,  for 
which  the  formula  is 


(x  -  z)2  =  f(x  -  xYf(x)dx.  (2.1) 

We  shall  assume  that  the  energy  levels  of  the  problem  are  so  closely 
spaced  that  they  can  be  treated  as  a  continuous  distribution.  For  each 
of  the  energy  levels,  or  states  of  the  system,  there  will  be  a  certain  value 
of  our  quantity  x.  We  shall  now  arrange  the  energy  levels  according  to 
the  values  of  x  and  shall  set  up  a  density  function,  which  we  shall  write 


in  the  form  e  k  ,  such  that 


dx  (2.2) 


is  the  number  of  energy  levels  for  which  x  is  in  the  range  dx.  We  shall 
see  later  why  it  is  convenient  to  write  our  density  function  in  the  form 
(2.2).  Now  we  know,  from  the  canonical  assembly,  as  given  in  formula 
(5.15)  of  Chap.  Ill,  that  the  fraction  of  all  systems  of  the  assembly  in  a 

_  E 
given  energy  level  is  proportional  to  e  **.    Thus,  multiplying  by  the 

number  (2.2)  of  levels  in  dx,  we  find  that  the  fraction  of  systems  in  the 


SBC.  2]  FLUCTUATIONS  105 

range  dx  is  given  by 


f(x)dx  =  const,  e          kT        dx,  (2.3) 

where  E(x)  is  the  energy  corresponding  to  the  levels  in  dx.  We  may 
immediately  evaluate  the  constant,  from  the  condition  that  the  integral 
of  f(x)  over  all  values  of  x  must  be  unity,  and  have 

(E(r)-Ts(x)] 

(2.4) 


In  the  problems  we  shall  be  considering,  f(x)  has  a  very  sharp  and 
narrow  maximum  at  a  certain  value  .TO,  rapidly  falling  practically  to  zero 
on  both  sides  of  the  maximum.  This  corresponds  to  the  fact  that  jc 
fluctuates  only  slightly  from  XQ  in  the  systems  of  the  assembly.  The 
reason  for  this  is  simple.  The  function  E(x)  —  Ts(x)  must  have  a  mini- 
mum at  £0,  in  order  that/(x)  may  have  a  maximum  therc\  The  function 
f(x)  will  then  be  reduced  to  l/e  of  its  maximum  value  when  E(x)  —  Ts(x) 
is  greater  than  its  minimum  by  only  kT.  But  E  is  the  energy  of  the  whole 
system,  of  the  order  of  magnitude  of  NkT,  if  there  arc  N  atoms  or  mole- 
cules in  the  system,  and  wo  shall  find  likewise  that  TS(JC)  is  of  this  same 
order  of  magnitude.  Thus  an  exceedingly  small  percentage  change  in  x 
will  be  enough  to  increase  the  function  E(x)  —  Ts(x)  by  kT  or  much 
more. 

We  can  get  a  very  useful  expression  for/(r)  by  assuming  that 

E(x)  -  7'sCr) 

can  be  approximated  by  a  parabola  through  the  very  narrow  range  in 
which  f(x)  is  appreciable.     Let  us  expand  E(x)  —  Ts(x)  in  Taylor's 
series  about  x0.     Remembering  that  the  function  has  a  minimum  at 
z0,  so  that  its  first  derivative  is  zero  there,  we  have 
E(x)  -  Ts(x)  =  E(x»)  -  TS(XO) 

*  x.   .  /0  -. 

x-^  2+  "•  (2'5) 

The  second  derivatives  in  Eq.  (2.5)  are  to  be  computed  at  x  =  XQ.     Then 
the  numerator  of  Eq.  (2.4)  becomes 


where 

i 

a  ~  2icT\dx* 


106  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VII 

The  function  e~€(*~*g)l  is  called  a  Gauss  error  curve,  having  been  used 
by  Gauss  to  describe  distribution  functions  similar  to  our  f(x)  in  the 
theory  of  errors.  It  equals  unity  when  x  =  z0,  and  falls  off  on  both  sides 
of  this  point  symmetrically,  being  reduced  to  l/e  when 

x  —  Xo  =  ±l/\/a. 

Using  formulas  (2.6)  and  (2.7),  and  the  integrals  (2.3)  of  Chap.  IV,  we 
can  at  once  compute  the  denominator  of  Eq.  (2.4)  and  find 


f(x)  =     /JV«(*-*0)>.  (2.8) 

Formula  (2.8)  is  an  obviously  convenient  expression  for  the  distribution 
function.  From  it  one  can  find  the  mean  square  deviation  of  XQ.  Obvi- 
ously the  mean  value  of  x  is  x0l  from  the  symmetry  of  Eq.  (2.8).  Then, 
using  Eq.  (2.1),  we  have 


1 


(x  -  *o)2  =  ~  (2.9) 

In  Eq.  (2.9),  we  have  a  general  expression  for  mean  square  fluctua- 
tions, if  only  we  can  express  E  —  Ts  as  a  function  of  x.  This  ordinarily 
can  be  done  conveniently  for  the  internal  energy  E.  We  shall  now  show 
that,  to  a  very  good  approximation,  s  equals  the  entropy  S,  so  that  it 
also  can  be  expressed  in  terms  of  the  parameter  x,  by  ordinary  thermo- 
dynamic  means.  To  do  this,  we  shall  compute  the  partition  function  Z 
of  our  assembly  and  from  it  the  entropy.  To  find  the  partition  function, 

_    V 

as  in  Eq.  (5.17)  of  Chap.  Ill,  we  must  sum  ekT  over  all  stationary  states. 
Converting  this  into  an  integral  over  x  and  remembering  that  Eq.  (2.2) 
gives  the  number  of  stationary  states  in  dx,  we  have 

-(/?—  -} 
Z  =  je      kT     dx 


kT         fe-*(f-f^dx 

Ts(ro)}      r 

=  e   '        kT    '~JZ-  (2.11) 

Then,  using  Eq.  (5.16)  of  Chap.  Ill,  we  have 
A  =  U  -  TS  =  -kTlnZ 

=  E(x<>)  -  Ts(xQ)  -  kT  In  J^-  (2.12) 

Now  if  the  peak  of  f(x)  is  narrow,  E(XO)  will  be  practically  equal  to  C7,  the 
mean  value  of  E,  which  is  used  in  thermodynamic  expressions  like  Eq. 


SBC.  3]  FLUCTUATIONS  107 

(2.12).  It  and  TS  are  proportional  to  the  number  of  molecules  in  the 
system,  as  we  have  mentioned  before.  But  a,  as  we  see  from  Eq.  (2.7), 
is  of  the  order  of  magnitude  of  the  number  of  molecules  in  the  system,  so 
that  the  last  term  in  Eq.  (2.12)  is  of  the  order  of  the  logarithm  of  the 
number  of  molecules,  a  quantity  of  enormously  smaller  magnitude  than 
the  number  of  molecules  itself  (In  1023  =  23  In  10  =  53).  Hence  we  can 
perfectly  legitimately  neglect  the  last  term  in  PJq.  (2.12)  entirely.  We 
then  have  at  once 

*(x0)  =  S.  (2.13) 

This  expression,  relating  the  entropy  to  the  density  of  energy  levels  by 
use  of  Eq.  (2.2),  is  a  slight  generalization  of  what  is  ordinarily  called 
Gibbs's  third  analogy  to  entropy  [his  first  analogy  was  the  expression 
-  k  S/t  In  /t,  his  second  was  closely  related  to  Eq.  (2.13)].  UsingEq.  (2.13), 
we  can  then  write  the  highly  useful  formula 


(/*•    /*.  V2    — 
•*•       •*(>;    —  y-,rt~, 


3.  Fluctuations  of  Energy  and  Density.-  Using  the  general  formula 
(2.14),  we  can  find  fluctuations  in  many  quantities.  Let  us  first  find  the 
fluctuation  in  the  total  energy  of  the  system,  getting  a  general  result  whose 
special  caso  for  the  perfect  gas  in  Boltzmann  statistics  was  discussed  in 
Sec.  1.  In  this  case  x  equals  E,  so  that  d2E/dx2  =  0.  The  derivative 
of  S  with  respect  to  E  is  to  be  taken  at  constant  volume,  for  all  the  states 
represented  in  the  canonical  assembly  are  computed  for  the  same  volume 
of  the  system.  Then  we  have,  using  the  thermodynamic  formulas  of 
Chap.  II,  Sec.  5, 


(as\      i       (svs\  ^i 

\dU/v       T'         \dU2Jv  T 


___ 

TdU/v  T2CV' 

Substituting  in  Eq.  (2.14),  we  find  for  the  fluctuation  of  energy 


(E  -  U)2  =  kT*Cv.  (3.2) 

We  can  immediately  see  that  this  leads  to  the  value  we  have  already 
found  for  the  perfect  gas  in  the  Boltzmann  statistics.  For  the  perfect 
gas,  we  have  Cv  =  %Nk,  so  that 


(E  -  C7)2  =  $AT(A;T)2,  (3.3) 


agreeing  with  the  value  found  from  Eqs.  (1.6)  and  (1.13).  The  formula 
(3.2),  however,  is  quite  general,  holding  for  any  type  of  system.  Since 
Cv  is  of  the  same  order  of  magnitude  for  any  system  containing  N  atoms 


108  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VII 

that  it  is  for  a  perfect  gas  of  the  same  number  of  atoms,  we  see  that  the 
energy  fluctuations  of  any  type  of  system,  in  the  canonical  assembly,  are 
negligibly  small.  The  heat  capacity  CV  is  proportional  to  the  number 
of  atoms  in  the  system,  so  that  the  mean  square  deviation  of  energy  from 
the  mean  is  proportional  to  the  number  of  atoms,  and  the  fractional  mean 
square  deviation  of  energy  for  N  atoms  is  proportional  to  1/JV,  as  in 
Eq.  (1.9). 

As  a  second  illustration  of  the  use  of  the  general  formula  (2.14),  we 
take  a  perfect  gas  and  consider  the  fluctuations  of  the  number  of  molecules 
in  a  group  of  G  cells  in  the  molecular  phase  space.  Two  important 
physical  problems  are  special  cases  of  this.  In  the  first  place,  the  G 
cells  may  include  all  those,  irrespective  of  momentum,  which  lie  in  a 
certain  region  of  coordinate  space.  Then  the  fluctuation  is  that  of  tho 
number  of  molecules  in  a  certain  volume,  leading  immediately  to  tho 
fluctuation  in  density.  Or  in  the  second  place,  we  may  be  considering 
tho  number  of  molecules  striking  a  certain  surface  per  second  and  the 
fluctuation  of  this  number.  In  this  case,  the  G  colls  include  all  those 
whoso  molecules  will  strike  the  surface  in  a  second,  as  for  example  the  colls 
contained  in  prisms  similar  to  those  shown  in  Fig.  IV-2.  Such  a  fluctua- 
tion is  important  in  the  theory  of  the  shot  effect,  or  the  fluctuation  of  tho 
number  of  electrons  omitted  thcrmionically  from  an  olement  of  surface 
of  a  heated  conductor,  per  second;  we  assumo  that  the  number  omitted 
can  be  computed  from  the  number  striking  tho  surface  from  inside  tho 
metal. 

To  take  up  this  problem  mathematically,  wo  oxpross  tho  energy  and 
the  entropy  in  terms  of  the  2VYs,  the  average  numbers  of  molecules  in 
the  various  cells.  The  energy  is 

U  =      #,ft,  (3.4) 


where  ct  is  the  energy  of  a  molecule  in  the  tth  cell  of  tho  molecular  phaso 
space.  For  the  entropy,  combining  Eqs.  (2.7)  and  (2.11)  of  Chap.  V,  we 
have 

8  =  -kfti  In  Ni  +  (1  ±  #.)  In  (1  ±  #,)],  (3.5) 


where  the  upper  sign  refers  to  Einstein-Bose  statistics,  the  lower  to  Fermi- 
Dirac,  and  where  we  shall  handle  tho  Boltzmanu  statistics  as  the  limiting 
case  of  low  density.  In  this  case,  the  quantity  x  is  the  number  of  mole- 
cules in  the  particular  0  cells  we  have  chosen,  which  we  shall  call  No,  so 
that 

#t  =  x/G  a*  No/G,  (3.6) 


SBC.  3)  FLUCTUATIONS  109 

where  #»  is  the  average  number  of  molecules  in  one  of  the  cells  of  our 
group  of  G,  which  we  assume  are  so  close  together  that  the  numbers  #t 
and  energies  et-  are  practically  the  same  for  all  G  cells.  In  terms  of  this 
notation,  we  have 


U  =  Noc*  +  terms  independent  of  No,  (3.7) 

and 

8  =  -kN0  In  §?  ±  k(G  ±  NG)  In  (l  ± 

+  terms  independent  of  No.  (3.8) 

Then  we  have 


and 


Substituting  in  Eq.  (2.14),  we  have 


AT      +     "    >  A7  \         ~~  ^          /^  \7  \  '  (3.10) 

Mo         .J+     ,    MG\  \r  I  i     ,    fVo 


±  ~- 


and 

' 

in  which  we  remember  that  the  upper  sign  refers  to  the  Einstein-Bose 
statistics,  the  lower  to  the  Fermi-Dirac,  and  where  we  find  the  Boltzmaim 
statistics  in  the  limit  where  Noo/G  approaches  zero,  so  that  the  right  side 
becomes  merely  l/NaQ.  Thus,  we  see  that  in  the  Boltzmann  statistics 
the  absolute  mean  square  fluctuation  nf  the  ni]TnKfir  nf  mnWnlog  in  Q 
volume  of  phase  space  equals  the  mean  number  in  the  volume,  and  the 
relative  fluctuation  is  the  reciprocal  of  the  mean  number,  becoming  very 
small  if  the  number  of  molecules  is  large.  These  are  important  results, 
often  used  in  many  applications.  We  also  see  that  the  fluctuations  in 
Einstein-Bose  statistics  are  greater  than  in  Boltzmann  statistics,  while  in 
Fermi-Dirac  statistics  they  are  less,  becoming  zero  in  the  limit 

N*/G  -  ffi  -  1, 


110  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  VII 

since  in  that  limit  all  cells  in  the  group  of  G  are  filled  and  no  fluctuation  is 
possible. 

As  a  third  and  final  illustration  of  Eq.  (2.14),  we  consider  the  fluctua- 
tion of  the  density  of  an  arbitrary  substance,  really  a  generalization  of  the 
result  we  have  just  obtained.  Instead  of  the  fluctuation  of  the  density 
itself,  we  find  that  of  the  volume  occupied  by  a  certain  group  of  molecules; 
the  relative  fluctuations  will  be  the  same  in  either  case,  since  a  given 
proportional  increase  in  volume  will  give  an  equal  proportional  decrease 
in  density.  In  this  case,  then,  the  quantity  x  is  the  volume  V  of  a  small 
mass  of  material  selected  from  the  whole  mass.  The  derivatives  in  Eq. 
(2.14)  are  those  of  the  whole  internal  energy  and  entropy  of  the  system,  as 
the  volume  of  the  small  mass  is  changed.  Since  the  part  of  the  system 
exterior  to  the  small  mass  is  hardly  changed  by  a  change  of  its  volume,  we 
can  assume  that  only  the  internal  energy  and  entropy  of  the  small  mass 
itself  are  concerned  in  Eq.  (2.14).  Furthermore,  we  are  interested  merely 
in  the  fluctuation  in  density,  neglecting  any  corresponding  fluctuation  in 
temperature,  so  that  the  derivatives  of  Eq.  (2.14)  are  to  be  computed  at 
constant  temperature.  Then  we  can  rewrite  Eq.  (2.14)  as 


where  A  =  U  —  TS.     Using  the  thcrmodynainic  formulas  of  Chap.  II, 
we  have 


so  that 


_  _P 

9Vr  ~  ' 


and 

,    /*T/\     1  ^8) 


The  quantity  in  brackets  is  the  isothermal  compressibility,  which  is  inde- 
pendent of  V0.  We  see,  then,  that  the  relative  mean  square  fluctuation 
of  the  volume  is  inversely  as  the  volume  itself,  becoming  small  for  large 
volumes.  This  is  in  accordance  with  the  behavior  of  the  other  fluctua- 
tions we  have  found. 

Let  us  check  Eq.  (3.13)  by  application  to  the  perfect  gas  in  the  Boltz- 
mann  statistics.    Using  PV  -  NkT,  we  have  (~l/Vd(dV/dP)T 


SBC.  3]  FL  UCT  UA  TIONS  1 1 1 

Thus 

TrTZTvA2       /.T         i 

(3.14) 

where  No  is  the  mean  number  of  molecules  in  F0.  This  value  checks 
Eq.  (3.11),  for  the  case  of  the  Boltzmann  statistics,  giving  the  same 
value  for  the  relative  fluctuation  of  volume  which  wo  have  already  found 
for  the  fluctuation  of  the  number  of  molecules  in  a  given  volume,  as  we 
have  seen  should  be  the  case.  For  substances  other  than  perfect  gases, 
the  compressibility  is  ordinarily  less  than  for  a  perfect  gas,  so  that  Eq. 
(3.13)  predicts  smaller  relative  fluctuations  of  density;  a  perfectly  incom- 
pressible solid  would  have  no  density  fluctuations.  (  )M  the  other  hand,  in 
some  cases  the  compressibility  can  be  greater  than  for  a  perfect  gas.  An 
example  is  an  imperfect  gas  near  the  critical  point,  where  the  compressi- 
bility approaches  infinity,  a  finite  change  in  volume  being  associated  with 
no  change  of  pressure;  here  the  density  fluctuations  are  abnormally  great, 
being  visible  in  the  phenomenon  of  opalescense,  the  irregular  scattering 
of  light,  giving  the  material  a  milky  appearance.  Below  the  critical 
point,  in  the  region  where  liquid  and  gas  can  coexist,  it  is  well  known  that 
the  material  maintains  the  same  pressure,  the  vapor  pressure,  through  the 
whole  range  of  volume  from  the  volume  of  the  liquid  to  that  of  the  gas. 
Thus  here  again  the  compressibility  is  infinite.  Formula  (3.13)  cannot  be 
strictly  applied  in  this  case,  but  the  fluctuations  of  density  which  it  would 
indicate  are  easily  understood  physically.  A  given  volume  in  this  case 
can  happen  to  contain  vapor,  or  liquid  drops,  or  both,  and  the  fluctuation 
of  density  is  such  that  the  density  can  be  anywhere  between  that  of  the 
liquid  and  the  vapor.  Such  problems  arc  hardly  suitable  for  a  fluctuation 
theory,  however;  we  shall  be  able  to  handle  them  better  when  we  take  up 
equilibrium  between  phases  of  the  same  substance. 


PART  II 
GASES,  LIQUIDS,  AND  SOLIDS 


CHAPTER  VIII 

THERMODYNAMIC   AND   STATISTICAL  TREATMENT   OF  THE 
PERFECT  GAS  AND  MIXTURES  OF  GASES 

In  Chap.  IV,  we  learned  some  of  the  simpler  properties  of  perfect  gases 
obeying  the  Boltzinann  statistics,  using  simple  kinetic  methods.  We 
can  go  a  good  deal  further,  however,  and  in  the  present  chapter  we  apply 
thermodynamics  and  statistical  mechanics  to  the  problem,  seeing  how  far 
each  can  carry  us.  The  results  may  seem  rather  formal  and  uninteresting 
to  the  reader.  But  we  are  laying  the  groundwork  for  a  great  many 
applications  later  on,  and  it  will  be  found  very  much  worth  while  to  under- 
stand the  fundamentals  thoroughly  before  we  begin  to  apply  them  to  such 
problems  as  the  specific  heats  of  gases,  the*  nature  of  imperfect  gases, 
vapor  pressure,  chemical  equilibrium,  thermionic  emission,  electronic 
phenomena,  and  many  other  subjects  depending  direct  1  y  on  the  properties 
of  gases.  For  generality,  we  shall  include  a  treatment  of  mixtures  of 
perfect  gases,  a  subject  needed  particularly  in  discussing  chemical  equilib- 
rium. We  begin  by  seeing  how  much  information  thermodynamics  alone, 
plus  the  definition  of  a  perfect  gas,  will  give  us,  and  later  introduce  a 
model  of  the  gas  and  statistical  methods,  obtaining  by  statistical 
mechanics  some  of  the  results  found  by  kinetic  theory  in  Chap.  IV. 

1.  Thermodynamics  of  a  Perfect  Gas. — By  definition,  a  perfect  gas 
in  the  Boltzmann  statistics  is  one  whose  equation  of  state  is 

PV  =  nRT,  (l.l) 

which  has  already  been  discussed  in  Sec.  3  of  Chap.  IV.  Furthermore, 
from  the  perfect  gas  law,  using  Eq.  (6.2)  of  Chap.  II,  we  can  prove  that  a 
perfect  gas  obeys  Joule's  law  that  the  internal  energy  is  independent  of 
the  volume  at  constant  temperature.  For  we  have 

(6U\    _     (dP\ 

WA       Wr  (    ' 

This  is  a  reversal  of  the  argument  of  Chap.  II,  Sec.  6,  where  we  used 
Joule's  law  as  an  experimental  fact  to  prove  that  the  gas  scale  of  tempera- 
ture was  identical  with  the  thermodynamic  temperature.  Here  instead 
we  assume  the  temperature  T  in  Eq.  (1.1)  to  be  the  thermodynamic 
temperature,  and  then  Joule's  law  follows  as  a  thermodynamic  conse- 
quence of  the  equation  of  state. 

115 


116 


INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  VIII 


No  assumption  is  made  thermodynamically  about  the  specific  heat  of 
a  perfect  gas.  Some  results  concerning  it  follow  from  thermodynamics 
and  the  equation  of  state,  however.  Wo  have  already  seen  that 


CV  -  Cv 


(1.3) 


where  Cp  and  Cv  are  the  heat  capacities  of  n  moles  of  gas,  at  constant 
pressure  and  volume  respectively.  Furthermore,  we  can  find  thermo- 
dynamically how  CP  changes  with  pressure*,  or  C\  with  volume,  at  con- 
stant temperature.  We  have 


jLjf*a\ 

dP\dT}P 

^OF), 


dPOT       dTdP 


(1.4) 


But  from  the  Table  of  Thermodynamic  Relations  in  Chap.  II  we  have 

(dH^ 


Thus,  we  find 


0. 


dT 


(1.5) 


(1.6) 
By  an  exactly  analogous  proof,  substituting  (7  for  H,  V  for  P,  wo  can  prove 

f*a\  =  T(^I  .  (i.7) 


Substituting  the  perfect  gas  law  in  Eqs.  (1.6)  and  (1.7),  we  find  at  once 


for  a  perfect  gas.  That  is,  both  specific  heats  are  independent  of  pressure, 
or  volume,  at  constant  temperature,  meaning  that  they  are  functions  of 
the  temperature  only. 

Thermodynamics  can  state  nothing  regarding  the  variation  of  specific 
heat  with  temperature.  It  actually  happens,  however,  that  the  heat 
capacities  of  all  gases  approach  the  values  that  we  have  found  theoret- 
ically for  monatomic  gases  in  Eq.  (3.18),  Chap.  IV,  namely, 


f  nR,         CP  =  f  nRt 


(1.9) 


SBC.  1]  THE  PERFECT  GAS  117 

at  low  enough  temperatures.  This  part  of  the  heat  capacity,  as  we  know 
from  Chap.  IV,  arises  from  the  translational  motion  of  the  molecules. 
The  remaining  heat  capacity  arises  from  rotations  and  vibrations  of  the 
molecules,  electronic  excitation,  and  in  general  from  internal  motions,  and 
it  falls  to  zero  as  the  tomperaturo  approaches  the  absolute  zero,  on  account 
of  applications  of  the  quantum  theory  which  wo  shall  make  in  later 
chapters.  Sometimes  it  is  useful  to  indicate  this  internal  heat  capacity 
per  mole  as  Ct,  so  that  we  write 

Cv  =  %nR  +  nCi,         CV  -  $nR,  +  nC,,  (1.  10) 

where  experimentally  C\  goes  to  zero  at  the  absolute  zero,  liquation 
(1.10)  may  be  taken  as  a  definition  of  C,. 

Next  we  take  up  the  internal  energy,  entropy,  Holmholtz  free  energy, 
and  Gibbs  free  energy  of  a  perfect  gas  From  Joule',1-  law,  the  internal 
energy  is  a  function  of  the  temperature  alone,  independent  of  volume. 
We  let  (/n  be  the  internal  energy  per  mole*  at  the  absolute  zero,  a  quantity 
which  cannot  be  determined  uniquely  since  there  is  always  an  arbitrary 
additive  constant  in  the  energy,  as  we  have  pointed  out  in  ('hap.  I,  Sec.  1. 
Then  the  change  of  internal  energy  from  the  absolute  zero  to  temperature 
T  is  determined  from  the  specific  heat,  and  we  have 

r  =  wr»  +  (TcvdT 

a/0 

(1.11) 


We  find  the  entropy  first  as  a  function  of  temperature  and  pressure, 
using  the  relations,  following  at  once  from  the  Table  of  Thermodynamic 
Relations  of  Chap.  II,  and  the  equation  of  state, 

(es\   =  Cj.       (*&\   =  Jsv\   _  _UR 

\dTjf        T'         \dPjT          \dTjP  P  (IAZ) 

Substituting  for  CP  from  Eq.  (1.10)  and  integrating,  we  have 

r  fT       7m 

S  =  ±nR  In  T  -  ntt  In  P  +  n\    C^  +  const.  (1.13) 

^  Jo        ' 

The  constant  of  integration  in  Kq.  (1.13)  cannot  bo  determined  by  thermo- 
dynamics. It  is  of  no  practical  importance  when  we  are  considering  tho 
gas  by  itself,  for  in  all  cases  we  have  to  differentiate  the  entropy,  or  take 
differences,  in  our  applications.  But  when  we  come  to  the  equilibrium  of 
different  phases,  as  in  the  problem  of  vapor  pressure,  and  to  chemical 
equilibrium,  we  shall  find  that  the  constant  in  the  entropy  is  of  great 
importance.  Thus  it  is  worth  while  devoting  a  little  attention  to  it  here. 
There  is  one  piece  of  information  which  we  can  find  about  it  from  thormo- 


118  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  VIII 

dynamics:  we  may  assume  that  it  is  proportional  to  the  number  of  moles 
of  gas.  To  see  this,  consider  first  two  separate  masses  of  gas,  one  of  n\ 
moles,  the  other  of  H*  moles,  both  at  the  same  pressure  and  temperature, 
with  a  partition  between  their  containers.  The  total  entropy  of  the  two 
masses  is  certainly  the  sum  of  the  separate  entropies  of  the  two.  Now 
remove  the  partition  between  them.  This  is  a  reversible  process,  involv- 
ing no  heat  flow,  and  hence  no  change  of  entropy,  so  long  as  the  gases  on 
the  two  sides  of  the  partition  are  made  of  identical  molecules;  if  there 
had  been  different  gases  on  the  two  sides,  of  course  diffusion  would  have 
occurred  when  the  partition  was  removed,  resulting  in  irreversibility. 
Thus  the  entropy  of  the  combined  mass  of  (n\  +  n%)  moles  is  the  sum  of 
the  separate  entropies  of  the  masses  of  HI  and  n2  moles.  But  this  will  be 
true  only  if  the  entropy  is  proportional  to  the  number  of  moles  of  gas. 
We  know  that  this  is  true  with  every  term  of  Eq.  (1.13)  except  the  con- 
stant, and  we  see  therefore  that  it  must  be  true  of  the  constant  as  well. 
Thus  the  constant  is  n  times  a  quantity  independent  of  n,  P,  77,  and 
hence  depending  only  on  the  type  of  gas.  This  constant  must  have  the 
dimensions  of  /£,  so  that  it  must  be  nil  times  a  numerical  factor.  For 
reasons  which  we  shall  understand  shortly,  it  is  convenient  to  write  it  in 
the  form  n(i  +  %)R,  where  i  is  called  the  chemical  constant  of  the  gas. 
Thus  we  have 


S  =  ^nR  In  T  -  nR  In  P  +  n  I    C^-  +  nRli  +  ~|       (1 .14) 

It  is  often  useful  as  well  to  have  the  entropy  as  a  function  of  tempera- 
ture and  volume.     We  can  find  this  by  integrating  the  equations 

f^\    =  ^        f^\   =  f^\   =  ^ 

V  .nT*  I  W  *  V  Z±  \7  I  V   UHl  I  17"  '  N  LtHjJ 

\O ./   /  V  •*•  \O  w    /  T  \U  J.   /  V  V 

or  by  substituting  for  P  in  terms  of  T  and  V  from  the  perfect  gas  law  in 
Eq.  (1.14).  The  latter  has  the  advantage  of  showing  the  connection 
between  the  arbitrary  constants  in  the  two  equations  for  entropy,  in 
terms  of  T  and  P,  and  in  terms  of  T  and  V.  Using  this  method,  we  have 
at  once 

3  CT    dT  f        5 

S  =  ~nR  In  T  +  nR  In  V  +  n  I    d-^-  +  nR\  i  +  ^ •  —  In  (nl 


n  L  C^Y 


From  Eq.  (1.16)  we  note  that  the  additive  constant  in  the  entropy,  in 
the  form  involving  the  temperature  and  volume,  has  a  term  —  nR  In  n, 
which  is  not  proportional  to  the  number  of  moles.  This  is  as  we  should 
expect,  however,  as  we  can  see  from  rewriting  Eq.  (1.16)  in  the  form 


S  = 


ln  T  -  nR\n~  +  n    (  C.~  +  nfiti  +  |  -  In  R\     (1.17) 


SBC.  1]  THE  PERFECT  GAS  119 

In  the  form  (1.17),  each  term  is  proportional  to  n,  except  the  one 


This  involves  the  ratio  n/F,  the  number  of  moles  per  unit  volume,  which 
is  proportional  to  the  density.  We  thus  see  from  Eq.  (1.17)  that  if  two 
masses  of  gas  of  the  same  temperature  and  the  same  density  are  put  in 
contact,  the  total  entropy  is  independent  of  whether  they  have  a  partition 
between  them  or  not.  This  statement  is  entirely  analogous  to  the  previ- 
ous one  about  two  masses  at  the  same  temperature  and  pressure. 

Next  we  find  the  Helm  holt  z  free  energy  A  =  U  —  •  TS  as  a  function 
of  temperature  and  volume.  We  can  find  this  directly  from  Eqs.  (1.11) 
and  (1.16)  or  (1.17).  Wo  havo  at  once 


=  n\  I/o  -  |«r  In  T  -  RT  In  - 
L  &  n 

.  dT  -  TJ*CJjr\  -  KT(i  +  1  -  In  «)]•     (1.18) 


A 


/o 

The  two  terms  depending  on  C,  can  bo  written  in  two  other  forms  by 
integration  by  parts.     Those  aro 

rT      __ ,  c1  dT  _  _  rv  CT  dT\  ,, 

Jo  Jo        T  J0  \Jo     l  T  / 

CT  IT  CT 

--rj.#j,(?"ir    o-20' 

To  prove  Eq.  (1.19),  we  apply  the  formula  Jw  dv  —  uv  —  Jv  du  to  the 
right  side,  setting 
»r 


iajL          j»           i^iui  ,-          ,        Arr 

,%-Y>        du  = Y~>        v  =  *•>        dv  =  dl, 

and  Eq.  (1.19)  follows  at  once.  To  prove  Eq.  (1.20),  we  integrate  the 
expression  fC%/T  dT  on  the  left  side  of  Eq.  (1.19)  by  parts,  setting 
u  =  1/r,  du  =  -l/T2dT,  v  =  fCidT,  dv  =  Ct  dT,  from  which  Eq. 
(1.20)  follows.  Thus  we  have  the  alternative  formulas 

To  y 

A  =  n   t/o  -  ^RT  In  T  -  /2'f  In  ~ 

+  1  -  In  /Z)       (1.21) 


-    fi77  In  T  -  B5T  In  - 


C*  dT  -  RW  +  1  -  In  «)   •     d-22) 


-  In  «)  ]•     d- 


120  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  VIII 

Finally,  we  wish  the  Gibbs  free  energy  G  =  U  +  PV  —  TS  as  a 
function  of  pressure  and  temperature.  Using  Eqs.  (1.11),  (1.14),  (1.19), 
and  (1.20),  this  has  the  alternative  forms 

G  =  n(u«  -  ^RT  In  T  +  RT  In  P 

7t  dT  -  T  f  C\  ~  -  flW)     (1.23) 
Jo         V  / 


=  n\  f/0  -  ^RT  In 

-  f  (f  c^)"1  -  ««]  <i- 


T7  +  /M'  In  P 

-24> 

-    #7'  Jn  T  +  K77  In  P 


We  see  that  the  term  proportional  to  T  in  Eqs.  (1.23),  (1.24),  and  (1.25), 
—  RTi,  has  a  particularly  simple  form.  It  is  for  this  reason  that  the 
additive  constant  in  the  entropy,  nR(i  +  f),  in  Eq.  (1.20),  is  chosen  in 
the  particular  form  it  is.  For  practical  purposes,  the  appearance  of  this 
quantity  in  the  Gibbs  free  energy  is  mon*  important  than  it  is  in  the 
entropy. 

2.  Thermodynamics  of  a  Mixture  of  Perfect  Gases.  —  Suppose  we 
have  a  mixture  of  HI  moles  of  a  gas  1,  n2  moles  of  gas  2,  and  so  on,  all  in  a 
container  of  volume  V  at  temperature  7\  First  we  define  the  fractional 
concentration  c*7  of  the  ith  substance  as  the  ratio  of  the  number  of  moles 
of  this  substance  to  the  total  number  of  moles  of  all  substances  present: 

£t     r=     --         -      —         -  .  V^-  U 

n\  +  nz  +  -  •  • 

We  also  define  the  partial  pressure  Pt  of  UK;  ?th  substance  as  the  pressure 
which  it  would  exert  if  it  alone  occupied  the  volume  V.  That  is,  since  all 
gases  are  assumed  perfect, 

r>//T 

^  =  n^-  (2.2) 

Then  the  equation  of  state  of  the  mixture  of  gases  proves  experimentally 
to  be  just  what  we  should  calculate  by  the  perfect  gas  law,  using  the 
total  number  of  moles,  (n\  +  W&  +•••)>  that  is,  it  is 

P  -  fa  +  n2  +  -  •  •  )~  (2.3) 


SEC.  2]  THE  PERFECT  GAS  121 

Equation  (2.3)  may  be  considered  as  an  experimental  fact;  it  follows,  how- 
ever, at  once  from  our  kinetic  derivation  of  the  equation  of  state  in  Chap. 
IV,  for  that  goes  through  without  essential  change  if  we  have  a  mixture 
instead  of  a  single  gas.  Then  from  Eqs.  (2.1),  (2.2),  and  (2.3),  we  have 


(2.4) 

From  Eq.  (2.4),  in  other  words,  the  fractional  concentration  of  one  gas 
equals  the  ratio  of  its  partial  pressure  to  tho  total  pressure.  Plainly  as  a 
corollary  of  Eq.  (2.4)  we  have 

Pi  +  /J*  +  •  •  •  -  ir^r-i—  ->  -  p>          (2-5) 

ni  -f-  HZ  •+•   •  •   • 
and 

ci  +  c,  +   •  •  •    =  I.  (2.6) 

Equation  (2.5)  expresses  the  fact  that  the  sum  of  the  partial  pressures 
equals  the  total  pressure. 

We  next  consider  the  entropy,  Helmholtz  free  energy,  and  Gibbs  free 
energy  of  the  mixture  of  gases.  We  start  with  the  expression  (1.14)  for 
the  entropy  of  a  single  gas.  In  a  mixture  of  gases,  it  is  now  reasonable  to 
suppose  that  the  total  entropy  is  the  sum  of  the  partial  entropies  of  each 
gas,  each  one  being  given  by  Eq.  (1.14)  in  terms  of  the  partial  pressure  of 
the  gas.  If  we  have  a  mixture  of  n\  moles  of  the  first  gas,  n^  of  the  second, 
and  so  on,  the  total  entropy  is  then 


s  =      n'fi  ln  T  +      c'~r~  ~  R  ln  P]  +  5R  +  i}R'    (2'7) 

3 

where  Cj  is  the  internal  heat  capacity  per  mole  of  the  jth  gas,  ij  its  chem- 
ical constant.  We  can  express  Eq.  (2.7)  in  terms  of  the  total  pressure. 
Then  we  have 


s  =     n/j?  ln  T  +     C;    ~  R  ln  p  +  /e  +  * 

jlnc,.     (2.8) 

In  Eq.  (2.8),  the  first  summation  is  the  sum  of  the  entropies  of  the  various 
gases,  if  each  one  were  at  the  same  pressure  P.  The  second  summation 
is  an  additional  term,  sometimes  called  the  entropy  of  mixing.  Since  the 
c/s  are  necessarily  fractional,  the  logarithms  are  negative,  and  the  entropy 


122 


INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  VIII 


of  mixing  is  positive.     It  is  such  an  important  quantity  that  we  shall 
examine  it  in  more  detail. 

Suppose  the  volume  V  were  divided  into  compartments,  one  of  size 
CiF,  another  c2F,  otc.,  and  all  the  gas  of  the  first  sort  were  in  the  first 
compartment,  that  of  the  second  sort  in  the  second,  and  so  on.  Then 
each  gus  would  have  the  pressure  P,  and  the  entropy  of  the  whole  system, 
being  surely  the*  sum  of  tho  entropies  of  the  separate  samples  of  gas,  would 
be  given  by  the  first  summation  of  Eq.  (2.8).  Now  imagine  tho  partitions 
between  the  compartments  to  be  removed,  so  that  the,  gases  can^irrcversi- 
bly  diffuse  into  o.'irlj  nthi*r.  This  diffusion,  being  an  irreversible  process, 
muafr  result  in  rm  increase  uf  entropy,  and  the  second -term  of  Eg.  (2.8).  the 
entropy  of  mixing,  represents  just  this  increase.  To  verify  its  correctness, 
we  must  find  an  alternative  reversible  path  for  getting  from  the  initial 
state  to  the  final  one,  and  find  /  dQ/T  for  this  reversible  path.  That  will 


II 

i  Gas 
1+2 

Gas  2 

_J    Gasl+2      p 


(b)  (c) 

Flu    VIII-1. — Reversible  mixing  oi  two  gases 

give  the  change  of  entropy,  whether  the  actual  path  is  reversible  or  not. 
We  shall  set  up  this  reversible  process  by  moans  of  semi-permeable  mem- 
branes, membranes  allowing  molecules  of  one  gas  to  pass  through  them, 
but  impervious  to  the  other  gas.  Such  membranes  actually  exist  in  a  few 
cases,  as  for  instance  heated  palladium,  which  allows  hydrogen  to  pass 
through  it  freely,  but  holds  back  all  other  gases.  There  is  no  error 
involved  in  imagining  such  membranes  in  other  cases  as  well. 

We  simplify  by  considering  only  two  gases.  Originally  let  the  parti- 
tion separating  the  compartments  c\V  and  c2V  be  two  semipermeable 
membranes  in  contact,  one  permeable  to  molecules  of  type  1  but  not  of 
type  2  [membrane  (1)],  the  other  permeable  to  type  2  but  not  type  1 
[membrane  (2)].  The  two  together  will  not  allow  any  molecules  to  pass. 
Each  of  the  membranes  will  be  subjected  to  a  one-sided  pressure  from  the 
molecules  that  cannot  pass  through  it.  Thus,  in  Fig.  VIII-1  (a),  mem- 
brane (1)  is  pushed  to  the  left  by  gas  2,  membrane  (2)  to  the  right  by  gas 
1.  Each  of  the  membranes  then  really  forms  a  piston,  and  if  rods  are 
attached  to  them  as  in  Fig.  VIII-1,  they  are  capable  of  transmitting  force 
and  doing  work  outside  the  cylinder.  Now  let  membrane  (1)  move  slowly 
and  reversibly  to  the  left,  as  in  (b),  doing  work  on  some  outside  device. 
If  the  expansion  is  isothermal,  we  know  that  the  internal  energy  of  the 
perfect  gases  is  independent  of  volume,  so  that  heat  must  flow  in  just  equal 


SEC.  21 


THE  PERFECT  GAS 


123 


to  the  work  done.  We  can  then  find  tho  boat  flowing  in  in  the  process  by 
integrating  P  <1V  for  membrane  (1).  The  pressure  exerted  on  it,  when 
the  volume  to  the  right  of  it  is  T,  is  n-£RT/V9  sinee  only  the  molecules  2 
exert  a  pressure  on  it.  Thus  the  work  done  when  the  volume  increases 
from  c%V  to  V  is 


Cv     RT  I  V  \ 

ri2-±dV  =  n2RT  In  (-      )  =  -RTni  In  c2. 

JctV  '  \^2r  / 


(2.9) 


This  equals  the  heat  flowing  in.     Since  the  corresponding  increase  of 
entropy  is  the  heat  divided  by  the  temperature,  it  is 


-/en2lnc2. 


(2.10) 


Now  in  a  similar  way  we  draw  the  membrane  (2)  to  the  right,  extracting 
external  work  reversibly  and  letting  heat  flow  in  to  keep  the  temperature 
constant.  By  similar  arguments,  the 
increase  of  entropy  in  this  process 
is  —Rni  In  Ci.  And  the  total  change  of 
entropy  in  this  reversible  mixing  is 


ICal  - 


AS  = 


In  ci  +  H2  In  r2),      (2.11) 


just  the  value  given  foi   the  entropy  of 
mixing  in  Eq.  (2.8). 

It  is  interest  ing  to  see  hon  the  entropy 
of  mixing  of  two  gases  depends  on  the 
concentrations.  Let  n  =  /M  +  "2  ~  the 

t_      i       i  1  p  1  /•  rm 

total  number  ol   moles  oi    gas.      J  hen, 
remembering  that  c\  +  c2  =  1,  [Eq.  (2.G)],  we  have 

AS  =  ~nfl[ci  In  ci  +  (1  -  ci)  In  (1  -  ci)]. 


o  o  10 

Ct 

vni-2.    Entropy  of  mixing  of 

tWO   KU*U*S. 


(2.12) 


In  Fig.  VIII-2  we  plot  AS  from  Eq.  (2.12),  as  a  function  of  ci,  which 
can  range  from  zero  to  unity.  We  see  that  the  entropy  of  mixing  has  its 
maximum  value  when  c±  =  ^,  or  with  equal  numbers  of  the  two  types  of 
molecules.  At  this  concentration  its  value  is  given  by 

-n#(i  In  4  +  i  In  i)  =  nR  In  2 

=  0.6931n/2  =  1.375  cal.  per  mole  per  degree. 

Having  found  the  entropy  of  a  mixture  of  gases  in  Eq.  (2.8),  it  is  a  sim- 
ple thing  to  find  the  Gibbs  free  energy,  from  the  relation 

G  =  U  +  PV  -  TS. 
We  have 

+  f  RT 


124  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  VIII 

PV  =  ty-RT,  (2.14) 

j 

so  that 

G  =  jJnXZ,  +  flrjn,  In  c,-,  (2.15) 

y  y 

wliere 


K  rr  rr  ,rr 

=  U,  -  5/22'  In  71  +        C/dr  -  27       C^  +  RT  In  P  -  t,*!? 
2  Jo  Jo        J 

P  CTf?T  CT 

=  Uj-^RTlnT  -  T\    %±\    C,  dT  +  RT  In  P  -  {,-RT.     (2. 
*  Jo  1    Jo 


16) 


In  Eqs.  (2.15)  and  (2.16),  [//  represents  the  arbitrary  constant  in  the 
energy  for  the  jth  type  of  molecule,  corresponding  to  f/o  of  Eq.  (1.11), 
and  the  last  step  in  Eq.  (2.16)  is  made  by  the  same  integration  by  parts 
used  in  Eqs.  (1.19)  and  (1.20).  The  quantity  (?/  represents  the  Gibbs 
free  energy  per  mole  of  the  jth  gas  at  temperature  T  and  pressure  P. 
Thus  Eq.  (2.15)  indicates  that  the  Gibbs  free  energy  of  the  mixture  is  the 
sum  of  the  free  energies  of  the  constituents,  at  the  final  pressure  arid 
temperature,  plus  a  mixing  term  which  is  always  negative. 

3.  Statistical  Mechanics  of  a  Perfect  Gas  in  Boltzmann  Statistics.  ~ 
Since  the  internal  energy  of  a  perfect  gas  is  independent  of  volume,  by 
Joule's  law,  it  is  obvious  that  there  can  be  no  forces  acting  between  the 
molecules,  for  if  there  were,  they  would  result  in  an  internal  energy 
depending  on  the  volume.  Thus  the  molecular  model  of  a  perfect  gas, 
which  we  make  the  basis  of  our  statistical  treatment,  is  a  collection  of  N 
molecules,  each  of  mass  m,  exerting  no  forces  on  each  other.  If  the  gas  is 
monatomic,  each  molecule  requires  only  three  coordinates,  the  rectangular 
coordinates  of  its  center  of  gravity,  and  the  three  conjugate  momenta,  to 
describe  it  completely,  so  that  the  phase  space  contains  QN  dimensions. 
When  the  gas  is  polyatomic,  additional  coordinates  are  necessary  to 
describe  the  orientation  and  relative  distances  of  separation  of  the  atoms 
in  the  molecules.  We  assume  there  are  s  such  coordinates,  5  momenta,  so 
that  in  all  there  are  (3  +  s)  coordinates,  (3  +  s)  momenta,  for  each 
molecule,  or  (6  +  2s)N  dimensions  in  the  general  phase  space.  We  shall 
call  the  coordinates  of  the  jth  molecule 


and  the  momenta 


Here  Xjyfif  are  the  coordinates  of  the  center  of  gravity,  pxj,  pui,  pzj  the 
components  of  total  momentum,  of  the  molecule. 


SBC.  3]  THE  PERFECT  GAS  125 

The  first  step  in  applying  statistical  mechanics  to  our  gas  is  to  compute 
the  partition  function  Z,  given  by  Eq.  (5.17)  or  (5.22)  of  Chap.  III.  To 
do  this,  we  must  first  know  the  energy  of  the  gas,  E,  as  a  function  of  the 
coordinates  and  momenta.  Since  there  are  no  forces  between  the  mole- 
cules, this  is  a  sum  of  separate  terms,  one  for  each  molecule.  Now  it  is  a 
general  theorem  of  mechanics  that  the  energy  of  a  structure  like  a  mole- 
cule, composed  of  particles  exerting  forces  on  each  other  but  not  actod  on 
by  an  external  force  field,  is  the  sum  of  tho  kinetic  energy  of  the  structure 
as  a  whole,  determined  by  the  velocity  of  tho  center  of  gravity,  and  an 
additional  term  representing  the  energy  of  the  internal  motions.  Thus. 
for  the  energy  of  the  gas,  we  have 

(3..) 


In  Eq.  (3.1),  e'  represents  the  energy  of  internal  motions  of  a  molecule. 
In  evaluating  the  partition  function,  we  must  take  exp  (  —  E/kT), 
whore  E  is  given  in  Kq.  (3.1),  and  integrate  over  all  coordinates  and 
momenta.  We  observe  in  the  first  place  that,  since?  E  is  a  sum  of  terms 
for  each  molecule,  exp  (  —  E/kT)  will  be  a  product  of  such  terms,  and  the 
whole  partition  function  will  be  a  product  of  N  factors,  one  from  each 
molecule,  each  giving  an  identical  integral,  which  we  can  refer  to  as  the 
partition  function  of  a  single  molecule.  We  next  observe  that  the  parti- 
tion function  of  a  single  molecule  factors  into  terms  depending  on  the 
center  of  gravity  of  the  molecule  and  terms  depending  on  the  internal 
motion.  Thus  we  have 


[i 


/* 
j 


(3.2) 


The  integration  over  x,  y,  z  is  to  be  carried  over  the  volume  of  the  con- 
tainer and  gives  simply  a  factor  V.  The  integrations  over  px,  py,  p~  are 
carried  from  —  <*>  to  <*>  ,  and  can  be  found  by  Eqs.  (2.3)  of  Chap.  IV.  The 
integral  depending  on  the  internal  coordinates  and  momenta  will  not  be 
further  discussed  at  present;  we  shall  abbreviate  it 


^«  =  i^  I    '   I  e  kT  dqi-  dp, 

_LL 

T.  (3.3) 


126  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CiiAP.  VIII 

In  Eq.  (3.3),  Z»  is  the  internal  partition  function  of  a  single  molecule. 
The  second  way  of  writing  it,  in  terms  of  a  summation,  by  analogy  with 
Eq.  (5.17)  of  Chap.  Ill,  refers  to  a  summation  over  all  cells  in  a  2s-dimen- 
sional  phase  space  in  which  q\  •  p8  are  the  dimensions.  We  note,  for 
future  reference,  that  the  quantity  Z>  depends  on  the  temperature,  but 
not  on  the  volume  of  the  gas. 

Using  the  methods  just  described,  Eq.  (3.2)  becomes 


Z,  I 


Z  =    -^wmkT^Z,  (3.4) 

There  is  one  thing,  however,  which  wo  have  neglected  in  our  derivation, 
and  that  is  the  fact  that  the  gas  really  is  governed  by  Fermi-  Dirac  or 
Kinstcin-Bose  statistics,  in  the  limit  in  which  they  load  to  tho  Boltzrnann 
statistics.  As  wo  havo  soon  in  Chap.  V,  Sec.  1,  on  account  of  the  identity 
of  the  molecules,  thoro  aro  roally  AM  different  colls  of  tho  general  phase 
space  corresponding  to  one  state  or  complexion  of  the  system.  The 
reason  is  that  there  are  A"!  different  permutations  of  the  molecules,  each 
of  which  would  lead  to  the  same  number  of  molecules  in  each  cell  of  the 
molecular  phase  space,  and  each  of  which  therefore  would  correspond  to 
the  same  complexion.  In  other  words,  by  integrating  or  summing  over 
all  values  of  the  coordinates  and  momenta  of  each  molecule,  we  have 
counted  each  complexion  AH  times,  so  that  the  expression  (3.4)  is  Nl 
times  as  great  as  it  should  be,  and  we  must  divide  by  AH  to  get  the  correct 
formula.  Using  Stirling's  formula,  I/AM  =  (e/N)N  approximately,  and 
multiplying  by  this  factor,  our  amended  partition  function  is 

(3.5) 

We  shall  use  Eq.  (3.5)  as  the  basis  for  our  future  work. 

From  the  partition  function  (3.5),  we  can  now  find  the  Helmholtz  free 
energy,  entropy,  and  Gibbs  free*  energy  of  our  gas.     Using  the  equation 

A  =  -kT  In  Z,  we  have 

A  =  -\NkT  In  T  -  NkT  In  V  -  NkT  In  Z< 


_  NkThl         +  !  _  lu 

L  A8 

From  A  we  can  find  the  pressure  by  the  equation  P  —  —  (dA/dV^T*     We 
have  at  once 

NIcT 
p  =  iHi,        or        PV  =  NkT.  (3.7) 


SBC.  3]  THE  PERFECT  GAS  127 

Thus  we  derive  the  perfect  gas  law  directly  from  statistical  mechanics. 
We  can  also  find  the  entropy,  by  the  equation  S  =  —(dA/dT)r.  Using 
the  relation  Nk  =  n/Z,  we  have 

S  =  |nfl  In  T  +  nR  In  V  +  nR^(T  In  Z%) 

,      _J.     (ZwmW*   ,5       ,    x   Dxl      ,0  ^ 
+  nfll  In        YS  ------  h  2  -  in  (n/Z)  I-     (3.8) 

From  S  we  can  find  the  specific  heat  GY  by  the  equation  CV  =  T(dS/dT)v. 
We  have 

CF  =  j*nfi  +  nRTJ~(T  In  Zt).  (3.9) 

The  specific  heat  given  in  Eq.  (3.9)  is  of  the  form  given  in  Eq.  (1.10);  by 
comparison  we  see  that  the  internal  heat  capacity  per  mole,  Ct,  is  given  by 

,72 

Ct  =  RT~(TlnZJ,  (3.10) 

similar  to  Eq.  (5.21)  of  Chap.  III.  We  shall  use  Eq.  (3.10)  in  the  next 
chapter  to  compute  the  specific  heat  of  polyatomic  gases.  For  mon- 
atomic  gases,  for  which  there  are  no  internal  coordinates,  of  course  Ct 
is  zero. 

Using  Eq.  (3.10),  we  can  rewrite  T  In  Zl  and  its  temperature  deriva- 
tives in  terms  of  Ct.  First,  we  consider  the  behavior  of  T  In  Zz  and  its 
derivatives  at  the  absolute  zero.  Let  there  be  go  cells  of  the  lowest  energy 
in  the  molecular  phase  space,  gi  of  the  next  higher,  and  so  on.  It  is  cus- 
tomary to  call  these  g's  the  a  priori  probabilities  of  the  various  energy 
levels,  meaning  merely  the  number  of  elementary  eells  which  happen  to 
have  the  same  energy.  Then  we  have,  from  Eq.  (3.3), 

_/?  -Jl1  __!l° 

Zi=g0e   kT  +  ffle   **•+•••    =  ^   w        as         y^0>       (3<uj 

Hence  T  In  Z%  approaches  T  In  go  —  ~  as  T  approaches  zero.     In  the 

derivative  of  T  In  Zt  with  respect  to  temperature,  the  only  term  which 
does  not  approach  zero  with  an  exponential  variation  is  the  term  In  0o. 
Using  these  values,  then,  we  have 


p(T  In  Zt)  =\ng,  +        ct?  (3.12) 

T  In  Z<  =         +  T  In  ,.  4-  cT.  (3.13) 


128  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  VIII 

Using  Eq.  (3.13),  we  can  rewrite  Eq.  (3.6)  as 


A  =  n\  f/o  -  l/Zr  In  T  -  RT  In  f  H 


whore 

and 

i  =  / 

AT' 

No  being  Avogadro's  number,  the  number  of  molecules  in  a  mole,  given 
in  Eq.  (3.10)  of  Chap.  IV.  We  observe  that  Eq.  (3.14)  is  exactly  the 
same  as  (1.21),  determined  by  thermodynamics,  except  that  now  we  have 
found  the  quantities  UQ,  tho  arbitrary  constant  in  the  energy,  and  z,  the 
chemical  constant,  in  terms' of  atomic  constants.  Similarly,  wo  can  show 
that  all  the  other  formulas  of  Sec.  1  follow  from  our  statistical  mechanical 
methods,  using  Eqs.  (3.15)  and  (3.16)  for  the  constants  which  could  not  be 
evaluated  from  thermodynamics. 

If  we  have  a  mixture  of  NI  molecules  of  one  gas,  N2  of  another,  and 
so  on,  the  general  phase  space  will  first  contain  a  group  of  coordinates 
and  momenta  for  the  molecules  of  the  first  gas,  then  a  group  for  the 
second,  and  so  on.  The  partition  function  will  then  be  a  product  of  terms 
like  Eq.  (3.5),  one  for  each  type  of  gas.  The  entropy  will  be  a  sum  of 
terms  like  Eq.  (1.14),  with  n»  in  place  of  n,  and  Pt,  the  partial  pressure,  in 
place  of  P.  But  this  is  just  the  same  expression  for  entropy  in  a  mixture 
of  gases  which  we  have  assumed  thcrmodynamically  in  Eq.  (2.7).  Thus 
the  results  of  Sec.  2  regarding  the  thermodynamic  functions  of  a  mixture 
of  gases  follow  also  from  statistical  mechanics. 

It  is  worth  noting  that  if  we  had  not  made  the  correction  to  our  parti- 
tion function  on  account  of  the  identity  of  particles  and  had  used  the 
incorrect  function  (3.4)  instead  of  the  correct  one  (3.5),  we  should  not 
have  found  the  entropy  to  be  proportional  to  tho  number  of  molecules. 
We  should  then  have  found  an  entropy  of  mixing  for  two  samples  of  the 
same  kind  of  gas:  the  entropy  of  (n\  +  n2)  moles  would  be  greater  than 
the  sum  of  the  entropies  of  HI  moles  and  n2  moles.  It  is  not  hard  to  show 
that  the  resulting  entropy  of  mixing  would  be  just  the  value  found  in 
Eq.  (2.8)  for  the  mixing  of  unlike  gases.  This  is  natural;  if  we  forgot 
that  the  molecules  were  really  alike,  we  should  think  that  the  diffusion 
of  one  sample  of  gas  into  another  was  really  irreversible,  since  surely  we 
cannot  separate  the  gas  again  into  two  samples  containing  the  identical 


SEC-  31  THE  PERFECT  GAS  129 

molecules  with  which  we  started.  But  the  molecules  really  are  identical, 
and  it  is  meaningless  to  ask  whether  the  molecules  we  find  in  the  final 
samples  are  the  same  ones  we  started  with  or  not.  Thus  mixing  two 
samples  of  unlike  gases  increases  the  entropy,  while  mixing  two  samples  of 
like  gases  does  not.  It  might  seem  paradoxical  that  these  two  results 
could  he  simultaneously  true.  For  consider  the  mixing  of  two  unlike 
gases,  with  increase  of  entropy,  and  then  let  the  molecules  of  the  two  kinds 
of  gas  gradually  approach  each  other  in  properties.  When  do  they 
become  sufficiently  similar  so  that  the  process  is  no  longer  irreversible,  and 
there  no  longer  is  an  increase  of  entropy  on  mixing?  This  paradox  is 
known  as  Gibbs's  paradox,  and  it  is  removed  by  modern  ideo.s  of  the  struc- 
ture of  atoms  and  molecules,  based  on  the  quantum  theory.  Jn  the 
quantum  thenrvr  ^re  is  a.  perfectly  clear-cut  distinction:  cither  two 
particles  are  identical  or  they  are  not.  There  is  no  such  thing  as  a  gradual 
change  from  one  to  the  other,  for  identical  particles  are  things  like  elec- 
trons, of  fixed  properties,  which  we  cannot  change  gradually  at  will. 
With  this  clear-cut  distinction,  it  is  no  longer  paradoxical  that  identical 
particles  are  to  be  handled  differently  in  statistics  from  unlike  particles. 


CHAPTER  IX 

THE  MOLECULAR  STRUCTURE  AND  SPECIFIC  HEAT 
OF  POLYATOMIC  GASES 

We  have  seen  in  the  preceding  chapter  that  the  equation  of  stale,  of  a 
perfect  gas  is  independent  of  the  nature  of  the  molecule.  This  is  not 
true,  however,  of  the  specific  heat;  the  quantity  G\,  which  we  called  the 
internal  specific  heat,  results  from  molecular  rotations  and  vibrations  and 
is  different  for  different  gases.  For  monatomic  gases,  where  Ct  is  zero, 
we  have  found  CV  =  fn/2,  CP  =  §nR.  Using  the  value  R  =  1.987  cal. 
per  degree  per  mole,  from  Eq.  (3.9)  of  Chap.  IV,  we  have  found  the 
numerical  values  to  be  Cv  —  2.980  cal.  per  degree  per  mole,  and 
Cp  =  4.968  cal.,  values  which  are  correct  within  the  limits  of  experimental 
error  for  the  specific  heats  of  He,  Ne,  A,  Kr,  Xe,  and  of  monatomic  vapors 
of  metals,  when  extrapolated  to  zero  pressure,  so  that  they  obey  the 
perfect  gas  law.  But  for  gases  which  are  not  monatomic,  the  additional 
term  C»  in  the  specific  heat  can  only  bo  found  from  a  rather  careful  study 
of  the  structure  of  the  molecule.  This  study,  which  we  shall  make  in 
the  present  chapter,  is  useful  in  two  ways,  as  many  topics  in  this  book  will 
be.  In  the  first  place,  it  throws  light  on  the  specific  problem  of  the  heat, 
capacity  of  gases.  But  in  the  second  place,  it  leads  to  general  and  valu- 
able information  about  molecular  structure  and  to  theories  which  can  be 
checked  from  the  experimentally  determined  heat  capacities. 

1.  The  Structure  of  Diatomic  Molecules.  —Many  of  the,  most  impor- 
tant molecules  are  diatomic  and  furnish  a  natural  beginning  for  our  study. 
The  atoms  of  a  molecule  are  acted  on  by  two  types  of  forces,  fundamen- 
tally electrical  in  origin,  though  too  complicated  for  us  to  understand  in 
detail  without  a  wide  knowledge  of  quantum  theory.  First,  there  are 
forces  of  attraction,  the  forces  which  are  concerned  in  chemical  binding, 
often  called  valence  forces.  We  shall  look  into  their  nature  much  more 
closely  in  later  chapters.  These  forces  fall  off  rapidly  as  the  distance  r 
between  the  atoms  increases,  increase;  rapidly  with  decreasing  r.  Being 
attractions,  they  are  negative  forces,  as  shown  in  Fig.  IX-1  (a),  curve  I. 
Secondly,  there  are  repulsive  forces,  quite  negligible  at  large  distances,  but 
increasing  even  more  rapidly  than  the  attraction  at  small  distances. 
These  repulsions  are  just  the  mathematical  formulation  of  the  impene- 
trability of  matter.  If  two  atoms  are  pushed  too  closely  into  contact, 
they  resist  the  push.  The  repulsion,  a  force  of  positive  sign,  is  shown  in 

130 


SEC.  1] 


POLYATOMIC  GASES 


131 


Curve  II,  Fig.  IX-1  (a).  If  the  atoms  were  rigid  spheres,  this  repulsion 
would  be  zero  if  r  were  greater  than  the  sum  of  the  radii  of  the  spheres, 
and  would  become  infinite  as  r  became  less  than  this  sum  of  radii.  The 
fact  that  it  rises  smoothly,  not  discontinuously,  shows  that  the  atoms  do 
not  really  have  sharp,  hard  boundaries;  they  begin  to  bump  into  each 
other  gradually,  though  quite  rapidly.  Now  when  we  add  the  attractive 
and  repulsive  forces,  wo  got  a  curve  liko  III  of  our  figure.  This  represents 
a  negative,  attractive  force  at  largo  distances,  changing  .sign  and  becoming 
positive  at  small  distances,  ««•  the  repulsion  begins  to  outweigh  tho  attrac- 
tion. At  the  distance  rf,  where  the  force  changes  sign,  there  is  a  position 
of  equilibrium.  The  attraction  and  repulsion  just  balance,  and  the  atoms 
can  remain  at  that  distance  apart  indefinitely.  This,  then,  is  the  normal 
distance  of  separation  of  the  atoms  in  the  molecule.  For  small  deviations 


(a)  (b) 

Fia.    IX-1  — Force    (a)    and    energy    (6)    of    interaction    at    two    atoms   in    a    molecule 
I,  attractive  term,  II,  repulsive  term,  III,  resultant  curve. 

of  r  from  rt,  the  curve  of  force  against  distance  can  be  approximated  by  a 
straight  line:  the  force  is  given  by  —  constant  (r  —  /v),  a  force  propor- 
tional to  the  displacement,  the  sort  found  in  elastic  distortion,  and  leading 
to  simple  harmonic  motion.  Under  some  circumstances,  the  atoms 
vibrate  back  and  forth  through  this  position  of  equilibrium,  the  amplitude 
increasing  with  temperature.  At  the  same  time,  the  molecule  as  a  whole 
rotates  about  its  center  of  gravity,  with  an  angular  velocity  increasing 
with  temperature,  and  of  course  finally  it  moves  as  a  whole,  the  motion 
of  the  center  of  gravity  being  just  as  with  a  single  particle. 

Rather  than  using  the  force,  as  shown  in  Fig.  IX-1  (a),  we  more 
often  need  the  potential  energy  of  interaction,  as  shown  in  (6)  of  the  same 
figure.  Here  we  have  shown  the  potential  energy  of  the  attractive  force 
by  1,  that  of  the  repulsive  force  by  II,  and  the  total  potential  energy  by 
III.  At  the  distance  rf,  where  the  force  is  zero,  the  potential  energy  has  a 
minimum;  for  we  remember  that  the  slope  of  the  potential  energy  curve 
equals  the  negative  of  the  force.  The  potential  energy  rises  like  a  parah- 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  IX 


ola  on  both  sides  of  the  minimum;  if  the  force  is 
*(r- 


potential  energy  is 


where  k  is  a  constant. 


-k(r  —  re),  then  the 
It  continues  to  rise 


indefinitely  as  /•  decreases  toward  zero,  since  it  requires  infinite  work  to 
force  the  atoms  into  contact.  At  large  values  of  r,  however,  it  approaches 
an  asymptotic  value;  it  requires  only  a  finite  amount  of  work  to  pull  the 
atoms  entirely  apart  from  each  other.  This  amount  of  work,  indicated 
by  D  on  the  figure,  is  the  work  required  to  dissociate  the  molecule,  and  is 
important  in  thermodynamic  applications. 

In  Table  IX-1  we  list  values  of  re  and  D  for  a  number  of  important 
diatomic  molecules.     A  few  of  these,  as  the  hydrides  of  carbon,  nitrogen, 

TABLE  IX-1. — CONSTANTS  OF  DIATOMIC  MOLECULES 


Substance 

7>                 r,  A 
(electron  volts))       ° 

a,  A-1 

H2              .      .                                                  103 

4  454 

0.75 

1.94 

CH 

81 

3  5 

1   12 

1.99 

NH.                                                       :          97 

1  2 

1.08 

1.96 

OH 

102         '' 

4   4 

0.96 

2.34 

HC1          

102 

4.40 

1.27 

1.91 

NO 

123 

5.3 

15     i     3.06 

02                                            .    . 

117 

5  09 

20         2.68 

N* 

170 

7  35 

09         3.11 

CO                                                                     223 

9  fi 

13         2.48 

C2                                                                        128 

5  (i 

31          2.32 

cu 

57 

2  47 

98         2  05 

Br2 

10 

1.96 

2  28     1     1.97 

I2. 

36 

1   53 

2  66 

1.86 

Li2 

26 

1    14 

2  67 

0.83 

Na2                                                                      18 

0  7(i 

3  07 

0.84 

K,       .                                                      ;           12 

0  51 

3.91 

0  78 

The  data  are  taken  from  Spom-r,  "  Molekulspekticri  und 

ihie  AnwendutiKcn  auf  chemische  Piob- 

lerne,"  Spnnger,  Berlin,  1933,  which  tabulates  Din  electron  volts,  /*,,  and  vihiational  frequencies.     The 

values  of  a  in  the  table  above  are  computed  u-mg  Eq.  (4.5)  of  the  pit-sent  chapter,  solved  for  a  in  terms 

of  the  vibrational  frequency  and  D,  as  tabulated  by  Sponer 

Thus  a  calculation  of  the  vibrational 

frequency  fiorn  data  of  the  present  table,  using  Kq.  (4  o), 

\vill  iiiitornutically  give  the  right  value 

Sponer'a  data  are  taken  from  band  spectra. 

and  oxygen,  do  not  ordinarily  occur  in  chemistry,  but  they  are  formed  in 
discharge  tubes  and  are  stable  molecules.  The  values  of  re  are  given  in 
angstrom  units  (abbreviated  A),  equal  to  10  8  cm.  The  values  of  D  are 
given  in  kilogram-calories  per  gram  mole,  where  we  remember  that 
1  kg.-cal.  is  1000  cal.,  or  4.185  X  1010  ergs.  We  also  give  D  in  electron 
volts.  One  electron  volt  by  definition  is  the  energy  acquired  by  an 
electron  in  falling  through  a  difference  of  potential  of  one  volt.  This  is 
the  charge  on  the  electron,  4.80  X  10~l°  e.s.u.,  times  one  volt,  or  ?fa  e.s.u. 


SBC.  1]  POLYATOMIC  GASES  133 

Thus  one  electron  volt  is  4.80  X  10~10/300  =  1.60  X  10~12  erg.  To 
compare  with  the  other  unit,  we  note  that  the  value  of  D  in  kilogram- 
calories  is  computed  for  a  gram  mole,  that  in  electron  volts  for  a  single 
molecule.  Thus  we  have 

1  electron  volt  per  molecule  =  1.60  X  10~12  erg  per  molecule 
=  1.60  X  10-  ia  X  6.03  X  1023  ergs  per  mole 
1.60  X  10~12  X  6.03  X  1023  , 


=  23.05  kg.-cal.  per  mole.  (1.1) 

A  very  useful  empirical  approximation  to  the  curves  of  Fig.  IX-1  has 
been  given  by  Morse,  and  it  is  often  called  a  Morse  curve.  As  a  matter  of 
fact,  Fig.  IX-1  was  drawn  from  Morse's  equation.  This  approximation  is 


Force  = 
Energy  =  C  +  D(e~^r~-^  -  2r  n(>  ->•>).  (1.2) 

Here  C  is  a  constant  fixing  the  zero  on  the  scale  of  ordinates  and  therefore 
arbitrary,  since  there  is  always  an  arbitrary  additive  constant  in  the 
energy.  D  is  the  energy  of  dissociation  tabulated  in  Table  IX-1,  and 
finally  a  is  a  constant  determining  the  curvature  about  the  minimum  of 
the  curve,  given  in  the  last  column  of  Table  IX-1.  Thus  from  the  data 
given  in  Table  IX-1  and  the  function  (1.2),  calculations  can  be  made  for 
the  interatomic  energy  or  force.  In  Eq.  (1.2),  the  first  term,  the  positive 
one,  represents  the  repulsive  part  of  the  potential  energy  between  the 
two  particles,  important  at  small  distances,  while  the  second,  negative 
term  represents  the  attraction  at  larger  distances.  While  the  Morse 
curve  has  no  direct  theoretical  justification,  still  it  proves  to  represent 
fairly  accurately  the  curves  which  have  been  calculated  in  a  few  cases 
from  quantum  mechanics.  Such  calculations  have  shown  that  it  is 
possible  to  explain  in  detail  the  interatomic  energy  curves,  the  magni- 
tudes of  D,  r(,  etc.  Nevertheless,  the  explanations  are  so  complicated 
that  it  is  better  simply  to  treat  the  constants  of  Table  IX-1  as  empirical 
constants,  without  trying  to  understand  why  some  molecules  have  greater 
/)'s,  some  less,  etc.,  in  terms  of  any  model.  For  future  reference,  how- 
ever, it  is  worth  while  pointing  out  that  the  smaller  D  is,  the  less  energy 
is  required  to  dissociate  the  molecule,  and  therefore  the  lower  the  tem- 
perature needed  for  dissociation.  We  shall  later  talk  about  thermal 
dissociation  of  molecules;  from  the  table  it  is  clear  that  the  best  molecules 
to  use  as  examples,  the  ones  which  will  dissociate  at  lowest  temperatures, 
will  be  iodine  and  the  alkali  metals  lithium,  sodium,  and  potassium. 
Conversely,  N2  and  CO  require  such  a  high  energy  for  their  dissociation 
that  they  do  not  dissociate  under  ordinary  circumstances. 


134  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IX 

2.  The  Rotations  of  Diatomic  Molecules. — If  molecules  were  governed 
by  classical  mechanics,  the  motions  of  their  atoms  would  have  the  follow- 
ing nature.  First,  the  molecules  as  a  whole  would  have  a  uniform  motion 
of  translation,  the  mean  kinetic  energy  being  ffcT,  on  account  of  the 
equipartition  of  energy,  discussed  in  Chap.  IV,  Sec.  2.  Secondly,  the 
molecules  would  rotate  with  uniform  angular  momentum  about  an  arbi- 
trary axis  passing  through  the  center  of  gravity.  Two  coordinates  are 
necessary  to  specify  the  rotation  of  the  molecule;  for  instance,  the  latitude 
and  longitude  angles  of  the  line  joining  the  centers  of  the  atoms.  Thus, 
from  equipartition,  the  mean  kinetic  energy  of  rotation  would  be 
(%)kT  =  kT.  Finally,  the  atoms  would  vibrate  back  and  forth  along 
the  line  joining  them.  One  coordinate,  the  interatomic  distance  r,  deter- 
mines this  vibration.  Thus,  from  equipartition  the  mean  kinetic  energy 
of  vibration  would  be  %kT.  At  the  same  time,  in  simple  harmonic  motion, 
there  is  a  mean  potential  energy  equal  to  the  mean  kinetic  energy  and 
hence  equal  also  to  \kT,  so  that  the  oscillation  as  a  whole  would  con- 
tribute kT  to  the  energy.  We  should  then  find  a  mean  energy  of  rotation 
and  vibration  of  2kT,  with  a  contribution  to  the  heat  capacity  per  mole 
of  2Nok  =  2R  cal.  per  degree.  This  would  bo  the  value  of  C,,  the  heat 
capacity  of  internal  motions  mentioned  in  Chap.  VIII,  Sec.  1,  if  the  gas 
obeyed  classical  mechanics.  Actually,  the  observed  values  are  less  than 
this,  increasing  from  small  values  at  low  temperatures  to  something 
approaching  2R  at  very  high  temperatures,  and  the  discrepancies  come 
from  the  fact  that  the  quantum  theory,  rather  than  the  classical  theory, 
must  be  used. 

We  have  seen  in  Chap.  IV,  Sec.  2,  that  equipartition  of  energy  is  found 
only  when  a  distribution  of  energy  levels  is  so  closely  spaced  as  to  be 
practically  continuous.  The  translational  levels  of  a  gas  are  spaced  as 
closely  as  this,  as  we  have  seen  in  Chap.  IV,  Sec.  1,  so  that  we  are  per- 
fectly justified  in  assuming  equipartition  for  the  translational  motion, 
resulting  in  the  heat  capacity  Cv  =  InR.  But  the  rotational  and  vibra- 
tional  levels  are  not  so  closely  spaced,  and  we  must  use  the  quantum 
theory  to  get  evon  an  approximately  correct  value  for  this  part  of  the 
specific  heat.  Our  first  problem,  then,  is  to  find  what  the  energy  levels 
of  a  diatomic  molecule  really  are.  This  can  be  done  fairly  accurately  by 
quite  elementary  methods.  To  a  good  approximation  we  can  treat  the 
rotation  and  vibration  separately,  assuming  that  the  total  energy  is 
the  sum  of  a  rotational  and  a  vibrational  term.  We  can  treat  the  vibra- 
tion as  if  the  molecule  wore  not  rotating,  and  the  rotation  as  if  it  were  not 
vibrating,  but  as  if  the  atoms  were  fixed  at  the  interatomic  distance  re. 

Let  us  consider  the  rotation  first.  In  Chap.  Ill,  Sec.  3,  we  have 
found  that  the  energy  of  a  rotating  body  of  moment  of  inertia  /,  angular 
momentum  p$,  is  p02/27,  which  is  equal  to  the  familiar  expression  i/co2, 


SBC.  2]  POLYATOMIC  OASES  135 

where  co  is  the  angular  velocity.  Furthermore,  we  have  found  that  in 
the  quantum  theory  the  angular  momentum  is  quantized  :  that  is,  it  can 
take  on  only  certain  discrete  values,  pe  —  nh/2ir,  where  n  is  an  integer. 
Thus,  the  energy  according  to  elementary  methods  can  have  only  the 
values  En  =  n2/i2/87r2/,  as  in  Eq.  (3.7),  Chap.  III.  As  a  matter  of  fact, 
as  we  saw  in  Eq.  (3.9),  Chap.  Ill,  we  must  modify  this  formula  slightly. 
To  agree  with  the  usual  notation  used  for  molecular  energy  levels,  we 
shall  denote  the  quantum  number  by  K,  rather  than  n.  Then  it  turns 
out  that  the  energy,  instead  of  being  given  by  K2h~/8w*I,  is  given  by  the 
slightly  different  formula 

E^  =  K(K  +  1)^,  (2.1) 

where  K  can  take  on  the  values  0,  1,  2,  ...  To  evaluate  these  rotational 
energy  levels,  we  need  the  moment  of  inertia  /,  in  terms  of  quantities  that 
we  know.  This  is  the  moment  of  inertia  for  rotation  about  the  center  of 
gravity  of  the  molecule.  Let  the  masses  of  the  two  atoms  be  Wi,  nii,  and 
let  nil  be  at  a  distance  TI  from  the  center  of  gravity,  ra2  at  a  distance  r2. 
Then  we  have 


Using  Eq.  (2.2),  we  find  at  once 


I  =  -  r—*,          *  =  -  -p  —  e 
mi  +  mo  mi  +  m2 

But  we  have  7  =  Srar2  =  m\r\  +  //^l-     Thus 


=  rf, 

=  ra2r2.  (2.2) 


„>  0\ 

re.  (2.3) 


t  =  rf,  (2.4) 

l          *  e  e1  v      y 

where 

miniz  111  /0  Kv 

M  =  -  -  ,         or        -  =  --  ---  (2.5) 

nil  +  nit  »       mi       m2 

That  is,  the  moment  of  inertia  is  that  of  a  mass  /x  (sometimes  called  the 
reduced  mass)  at  a  distance  re  from  the  axis. 

Having  found  the  rotational  energy  levels,  we  wish  first  to  find  how 
closely  spaced  they  are,  to  see  whether  we  can  use  the  classical  theory  to 
compute  the  specific  heat.  The  thing  we  are  really  interested  in  is  the 
spacing  of  adjacent  levels  as  compared  with  kT]  if  the  spacing  is  small 
compared  with  fc!F,  the  summation  in  the  partition  function  can  be 


136  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IX 

replaced  by  an  integration  and  equipartition  will  hold.  Let  us  consider 
the  two  lowest  states,  for  K  =  1  and  0,  and  find  the  energy  difference 
between  them.  We  have 


-  #<,  =  (1  X  2  -  0  X 

=  -A2- 

47T2/' 


(2.6) 


We  wish  to  compare  this  quantity  with  kT;  it  is  more  convenient  to  define 
a  quantity  which  we  can  call  a  characteristic  temperature  6rot  by  the 
equation 

A2 


and  then  our  condition  for  the  applicability  of  the  integration  is  T  »  9rot. 
We  now  give,  in  Table  IX-2,  values  for  the  characteristic  temperatures 

TABLE  IX-2.  —  CHARACTERISTIC  TEMPERATURE  FOR  ROTATION,  DIATOMIC  MOLECULES 

Orot, 

Substance  °abs. 

H2           .  171 

CH  41  4 

NH.  44  1 

OH  .  55  0 

HC1  30  5 

NO  4.93 

O2  4  17 

N2  5  78 

CO  5  53 

C2  4  70 

C12                                                       .  .  .0  693 
Br2                                                                                                .           0  233 

h  0  108 

Li2  I  96 

Xa2  0  447 

K5  0  162 

By  EIJ  (2  7),  we  have  defined  Orot  by  the  relation  that  fcO,,,t  equals  the  eneigy  difference  between 
the  two  lowest  rotational  energy  levels  of  the  molecule  The  method  of  calculation  from  the  value  of 
re  in  Table  IX-1  is  illustrated  in  the  text. 

for  the  same  diatomic  molecules  listed  in  Table  IX-1.  These  values  are 
calculated,  using  Eq.  (2.7),  from  the  masses  of  the  atoms,  known  from 
the  atomic  weights  and  Avogadro's  number,  and  the  values  of  re  in  Table 
IX-1.  Thus,  for  instance,  for  H2  we  find 

ft2  (6.61  X  10-27)2(6.03  X  1023) 


I 

X  10-8)2(1.379  X 


171°  abs. 


SBC.  21  POLYATOMIC  GASES  137 

1.008 

2 
Here  /*  =  —  =  6  Q3  x  1023'  and  r°  =  °-75  X  10~*  cm- 

From  Table  IX-2,  we  see  that  the  gases  are  divided  distinctly  into 
three  types.  In  the  first  place,  hydrogen  stands  entirely  by  itself,  on 
account  of  its  small  mass.  The  characteristic  temperature  0rot,  having 
the  value  171°  abs.,  is  the  only  one  at  all  comparable  with  room  tempera- 
ture. Next  are  the  hydrides,  with  characteristic  temperatures  between 
20  and  60°  abs.  Finally,  the  characteristic  temperatures  of  all  gases 
not  containing  hydrogen  lie  below  6°  abs.  Now  it  is  not  easy  to  calculate 
the  specific  heat  of  a  rotating  molecule  in  the  quantum  theory  on  account 
of  mathematical  difficulties,  but  the  result  is  qualitatively  simple.  The 
specific  heat  rises  from  zero  at  low  temperatures,  comes  to  the  classical 
value  at  high  temperatures,  and  the  range  of  temperature  in  which  it  is 
rising  is  in  the  neighborhood  of  the  characteristic  temperature  6rot  which 
we  have  tabulated  in  Table  IX-2.  We  may  then  infer  from  Table  IX-2 
that  for  molecules  not  containing  hydrogen,  the  rotational  specific  heat  will 
have  attained  its  classical  value  at  a  very  low  temperature,  so  that  we 
are  entirely  justified  in  using  the  classical  value  in  our  calculations.  As 
an  illustration,  we  give  values  computed  for  the  specific  heat  of  NO  at  low 
temperatures.  We  remember  that  the  translational  part  of  (7/»  is 
f/2  =  4.97  caL,  whereas  if  the  rotational  specific  heat  is  added  we  have 
%R  =  6.96  cal.  The  specific  heat  is  actually  computed1  to  be  4.97  at 
0.5°  abs.,  5.12  at  1.0°,  6.91  at  5.0°,  and  6.95  at  10°.  Of  course,  NO  at 
atmospheric  pressure  liquefies  at  a  higher  temperature  than  this,  but  at 
sufficiently  reduced  pressure  the  boiling  point  can  be  reduced  as  far  as 
desired,  so  that  there  is  nothing  impossible  about  having  the  vapor  at  a 
temperature  as  low  as  desired. 

The  hydrides  have  a  decidedly  higher  temperature  range  in  which 
the  rotational  specific  heat  is  less  than  the  classical  value.  And  for  hydro- 
gen, the  quantum  theory  value  is  appreciably  less  than  the  classical  value 
even  at  room  temperature.  Thus,  at  92°  abs.,  we  have  the  value  5.28  for 
CP;  at  197°,  6.30;  at  288°,  6.78.  The  specific  heat  of  hydrogen  presents 
complications  not  occurring  with  any  other  substance.  It  turns  out,  for 
reasons  which  are  too  complicated  to  go  into  here,  that  in  the  energy  levels 
of  hydrogen  and  of  other  diatomic  molecules  made  of  two  like  atoms,  we 
can  make  a  rather  sharp  separation  between  the  energy  levels 


1  For  these  values,  and  much  other  data  relating  to  thermal  properties  of  gases,  see 
Landolt-Bornstein,  "Physikalisch-chernische  Tabellen,"  Dritter  Erganzungsband, 
Dritter  Teil,  pp.  2315-2364,  Springer,  1936. 


138  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IX 

in  which  K  is  even,  and  those  in  which  K  is  odd.  As  a  matter  of  fact,  if  a 
molecule  is  in  a  state  with  K  even,  for  instance,  almost  no  physical  agency, 
such  as  collisions  with  other  molecules,  seems  to  have  any  tendency  to 
transfer  it  to  a  state  with  K  odd.  It  is  almost  as  if  the  gas  wore  a  mixture 
of  two  gases,  one  with  K  even,  the  other  with  K  odd.  Actually  names 
have  been  given  to  these  gases,  the  case  of  K  even  being  called  parahydro- 
gen,  that  of  K  odd  being  called  orthohydrogen.  At  high  temperatures, 
such  as  we  ordinarily  have,  we  have  molecules  of  both  types  in  thermal 
equilibrium.  At  first  sight  we  should  expect  that  there  would  be  about 
equal  numbers  of  molecules  of  both  sorts,  but  an  additionally  complicating 
feature  concerning  the  a  priori  probabilities  of  the  states  results  in  there 
being  three  times  as  many  molecules  of  orthohydrogen  as  of  parahydrogcn 
at  high  temperatures.  When  specific  heat  measurements  are  made  at 
low  temperatures,  it  has  always  boon  the  practice  to  start  with  hydrogen 
at  room  temperature,  and  cool  it  down.  On  account  of  the  slow  rate  of 
conversion  of  one  type  of  hydrogen  into  the  other,  the  two  types  of  hydro- 
gen appear  in  the  same  ratio  of  three  to  one  \\hen  the  low  temperature 
measurements  are  made.  This  is  not  the  equilibrium  distribution  cor- 
responding to  low  temperature.  At  the  very  lowest  temperature,  we 
should  expect  all  the  molecules  to  be  in  the  lowest  possible  state,  that  of 
K  —  0,  a  state  of  parahydrogen.  Thus  to  compute  the  observed  specific 
heat,  we  must  assume  a  mixture  of  ortho-  and  parahydrogen  in  the  ratio 
of  three  to  one,  find  the  specific  heat  of  each  separately,  and  add.  When 
this  is  clone,  the  result  agrees  with  experiment.  To  get  the  true  equilib- 
rium mixture  at  low  temperature,  we  must  either  wait  a  period  of  a 
number  of  days,  or  employ  certain  catalysts,  which  speed  up  the  trans- 
formation from  one  form  of  hydrogen  to  the  other. 

3.  The  Partition  Function  for  Rotation. — Though  we  shall  not  be 
able  to  find  the  rotational  specific  heat  on  account  of  mathematical 
difficulties,  still  it  is  worth  while  setting  up  the  partition  function  for 
rotation  and  showing  the  limiting  value  which  it  approaches  at  high 
temperatures.  To  do  this,  we  must  sum  exp  (  —  Elot/kT\  where  Elot  is 
given  in  Eq.  (2.1),  for  all  values  of  K.  There  is  one  point,  however,  which 
we  have  not  yet  considered.  That  is  the  fact  that  the  energy  levels  are 
what  is  called  degenerate:  each  level  really  consists  of  several  stationary 
states  and  several  cells  in  the  phase  space.  The  reason  for  this  is  what  is 
called  space  quantization.  We  merely  describe  it,  without  giving  the 
justification  in  terms  of  the  quantum  theory.  It  is  natural  that  the 
angular  momentum,  Kh/2ir,  of  the  rotating  molecule  can  be  oriented  in 
different  directions  in  space.  As  a  matter  of  fact,  it  turns  out  that  in 
quantum  theory  there  are  just  (2K  +  1)  allowed  orientations,  each 
corresponding  to  a  different  stationary  state  and  a  different  cell.  One 
simple  way  of  describing  these  orientations  is  in  terms  of  a  vector  model, 


SEC.  3] 


POLYATOMIC  GASES 


139 


as  shown  in  Fig.  IX-2.  Here  we  have  a  vector  of  length  Kh/2ir.  Then 
it  can  be  shown  that  the  projection  of  this  vector  along  a  fixed  direction 
is  allowed  to  have  just  the  values  Mh/2ir,  where  M  is  an  integer,  cor- 
responding to  the  various  orientations  shown  in  the  figure.  Obviously, 
the  maximum  value  of  M  is  K,  coming  when  the  angular  momentum  is 
oriented  along  the  fixed  direction,  and  the  minimum  is  —  M  when  it  is 
opposite.  But  there  are  just  (2K  +  I)  integers  in  the,  group  K,  K  —  1, 
K  —  2,  •  •  •  —(K  —  1),  —  K,  justifying  us  in  our  statement  that  there 
are  (2K  +  1)  allowed  orientations.  One  says  thnt  the 
state  is  (2K  +  l)-fold  degenerate. 

Considering  this  degeneracy,  we  see  that  the  term  in 
the  partition  function  corresponding  to  a  given  K  must 
really  be  counted  (2K  +  1)  times,  since  all  these  stationary 
states,  corresponding  merely  to  different  orientations  in 
space,  obviously  have  the  same  energy.  Thus,  we  have 


(3.1) 


z. 


*'u;.  IX-2.— 


where  Zrot  is  the  factor  in  the  partition  function  of  a 

single  molecule,  Z,  of  Eq.  (3.3)  of  Chap.  VIII,  coming  from 

rotation.     It  is  this  summation  which  unfortunately  can-  quantization 

not  be  evaluated  analytically.     But  we  can  handle  it  in 

the  limit  of  high  temperature,  for  then  the  terms  corre- 

spending  to  successive  K's  will  differ  so  little  that  the 

summation  can  be  replaced  by  an  integration.     We  then  have 


of 


2 
1,  o,    -i,    -2, 

""*•"• 


/•oo  -(+l\h' 

lim    Zrot  =    I     (2K  +  })c~**'ikT-  <1K. 
•/o 


(3.2) 


T  »0rot 


The  bulk  of  the  integral  in  Eq.  (3.2)  will  come  from  high  quantum  num- 
bers or  high  values  of  K.  For  these,  we  can  neglect  unity  compared  to 
K,  obtaining 


lim 

T  »0rot 


(3.3) 
(3.4) 


using  the  integrals  (2.3)  of  Chap.  IV. 

From  the  expression  (3.4),  we  can  in  the  first  place  find  the  rotational 
heat  capacity,  using  Eq.   (5.21)  of  Chap.  III.     This  may  be  written 


C 

tr 


140  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IX 

where  we  must  multiply  by  NQ  because  our  quantity  Zrot  refers  to  a  single 
molecule.  Thus  we  have,  substituting  Eq.  (3.4)  in  Eq.  (3.5), 

Crot  =  Nok,  (3.6) 

in  accordance  with  equipartition.  It  is  also  interesting  to  compute  tho 
contribution  of  tho  rotation  to  the  entropy,  as  given  in  Eq.  (3.8)  of  Chap. 
VIII.  From  that  equation,  the  contribution  is 

*R*T(T  In  /„,)  =  wft(ln  ~-  +  1  +  In  T\  (3.7) 

Thus,  using  Eq.  (3.8)  of  Chap.  VIII,  the  entropy  in  the  temperature  range 
where  the  rotation  can  be  treated  classically,  but  where  the  vibration  is 
not  excited  enough  to  contribute  appreciably  to  the  entropy,  is 

S  =  %nR  In  T  +  nR  In  V  +  nR[i'  +  1  -  In  (n/Z)],  (3.8) 

where 

.,  (2rm)*'k''*         .      tar2/ A'  ,Q     , 

i'  =  In  v~ — j/8 h  In  -  jp    -  (3.9) 

The  quantity  i1  of  Eq.  (3.9)  can  be  considered  as  tho  chemical  constant 
of  a  diaTf"Tni'p  f'"*r  *"  fionnontion  with  the  formula  (3.8)  tor  the  entropy. 
We  must  remember,  however,  that  Eq.  (3.8)  holds  only  in  a  restricted 
temperature  range,  as  stated  above;  with  some  gases,  the  vibration  begins 
to  contribute  to  the  entropy  even  at  room  temperature,  as  we  shall  see 
in  the  next  section.  It  is  sometimes  useful  to  have  the  formula  for  Gibbs 
free  energy  of  a  diatomic  gas  in  the  range  where  Kq.  (3.8)  is  correct.  This 
is  easily  found  to  be 

(f  =  n(U»  -  iRT  In  T  +  RT  in  P  -  RTi'),  (3.10) 

where  i'  is  given  in  Eq.  (3.9). 

4.  The  Vibration  of  Diatomic  Molecules. — In  addition  to  their 
rotation,  we  have  seen  that  diatomic  molecules  can  vibrate  with  simple 
harmonic  motion  if  the  amplitude  is  small  enough.  We  shall  use  only 
this  approximation  of  small  amplitude,  and  our  first  stop  will  be  to  calcu- 
late the  frequency  of  vibration.  To  do  this,  we  must  first  find  the  linear 
restoring  force  when  the  interatomic  distance  is  displaced  slightly  from  its 
equilibrium  value  rr.  We  can  get  this  from  Eq.  (1.2)  by  expanding  the 
force  in  Taylor's  series  in  (r  —  rt).  We  have 

Force  =  2aD[l  -  2a(r  -  rf)  •  •  •   -  I  -f  a(r  -  r.)  •  •  •  ] 

=  -2a*D(r  -  r.),  (4.1) 

neglecting  higher  terms.  Now  we  can  find  the  equations  of  motion  for 
the  two  particles  of  mass  m\  and  m^  at  distances  r\  and  r^  from  the  center 


SBC.  4]  POLYATOMIC  GASES  141 

of  gravity,  where  TI  +  r2  =  r,  under  the  action  of  the  force  (4.1).     These 
are 


r2  - 


~  i       r2  -  rc).  (4.2) 

We  divide  the  first  of  these  equations  by  MI,  the  second  by  W2,  and  add, 
obtaining 


or 

M~  =  -2a'D(r  -  r.),  (4,3) 

where  /x  is  given  by  Eq.  (2.5).  The  vibration,  then,  is  like  that  of  a 
particle  of  mass  /x,  with  a  force  constant  —  2a2D.  By  elementary  mechan- 
ics, we  know  that  a  particle  of  mass  M,  acted  on  by  a  linear  restoring  force 
—  kx,  vibrates  with  a  frequency 

1         /*  (AA\ 

v  =  2^\t  (4<4) 

Thus  the  frequency  of  oscillation  of  the  diatomic  molecule  is 

(4.5) 

We  have  found  in  Eq.  (3.8),  Chap.  Ill,  that  the  energy  levels  of  an 
oscillator  of  frequency  i>,  in  the  quantum  theory,  are  given  by 

fe\,b  =  (v  +%)hv,  (4.6) 

where  v  is  an  integer  (called  n  in  Chap.  III).  The  spacing  of  successive* 
levels  is  hv.  We  may  then  expect,  as  with  the  case  of  rotation,  that  for 
temperatures  T  for  which  hv/kT  is  small,  or  temperatures  large  compared 
with  a  characteristic  temperature 


the  classical  theory  of  specific  heats,  based  on  the  use  of  the  integration 
to  find  the  partition  function,  is  applicable,  while  for  temperatures  small 
compared  with  6vlb  we  must  use  the  quantum  theory.  To  investigate 
this,  we  give  in  Table  IX-3  the  characteristic  vibrational  temperatures 
of  the  molecules  we  have  been  considering.  The  values  of  Table  IX-3 


142  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IX 

can  be  found  from  D  and  a  as  tabulated  in  Table  IX-1.     Thus  for  H2  we 
have 


6.61  X_10^       /2(1.95  X  1Q8)2(103)(4.185  X  1010) 
*"       ar(1.379  X  10-")    /  1.008 

V  ~2~ 

=  6140°  abs.  (4.8) 


We  see  from  Tal?le  IX-3  tlmt  fnr  prn.p.f.inp.l]y  all  thft  molecules  the  charac- 
i«  Inrpro  pnmpfljpd  to  room  tempcratyrn,.  so  thftl;  at  all 


ordinary  tttnippfatdip^  wo  ™n^  11gr>  *-hf>  quantum  theory  of  specific  heat. 
We  also  note  that  in  every  case  the  characteristic  temperature  for  vibra- 

TABLE  IX-3.  —  CHARACTERISTIC  TEMPERATURE  FOR  VIBRATION,  DIATOMIC  MOLECULES 

Substnnoo  ^V1b,  °  a^>s. 

H2  6140 

CH  .      .                                  4100 

NH.    .  .                                 .   4400 

OH  ..   5360 

HC1  4300 

NO             .  .   2740 

O2  .  .   2260 

N2  .  .  3380 

CO             .  .   3120 

C2  ..   2370 

C12  ...       810 

Br2  470 

I2  .  .     310 

Li2              .               .  500 

Na,  230 

K2                             ....  140 

These  values  are  calculated  as  in  Eq.  (4.8). 


is:i£firv  large  compare^]  lift  i^n-*'  nf  rnt  itinn^  That  is.  the  rotational 

"^  ^"fh    m^rfi   filoafilv  spn.p.P<l   tlin.n    f.lio 


This  is  a  characteristic  feature  of  molecular  energy  levels,  which  is  of 
great  importance  in  the  study  of  band  spectra,  the*  spectra  of  molecules. 
5.  The  Partition  Function  for  Vibration.  —First,  wo  shall  calculate 
the  partition  function  and  specific  heat  of  our  vibrating  molecule  by  classi- 
cal theory,  though  we  know  that  this  is  not  correct  for  ordinary  tempera- 
tures. Using  the  expression  (4.1)  for  the  force,  we  have  the  potential 
energy  given  by 


-  r.)»,  (5.1) 

and  the  kinetic  energy  is 


SBC.  5]  POLYATOMIC  GASES  143 

where  pr  is  the  momentum  associated  with  r,  equal  to  /*  dr/dt.  Then,  by 
analogy  with  Eq.  (5.22)  of  Chap.  Ill,  the  vibrational  partition  function 
Zvib  can  be  computed  classically  as  an  integral, 

1    r  °°        <'*D(r  -  r,)«          /»  * 

z-  =  /J  '~"~*r  "*J_« 


In  the  integral  over  r,  we  can  approximately  replace  by  an  integral  from 
—  oc  to  oo ,  for  the  exponential  in  the  integrand  is  practically  zero  for 
negative  values  of  r.  If  we  do  this,  we  have 

kT  f      . 

=  TV'  (5'4) 

The  use  of  an  equation  analogous  to  Eq.  (3.5)  gives  the  value  R  for  the 
vibrational  contribution  to  the  specific  heat,  as  mentioned  at  the  begin- 
ning of  Sec.  2. 

Next  we  calculate  the  specific  heat  in  the  quantum  theory.     The 
partition  function  is 


=  Ye    \   ^2jkT 

_JLL/          _*i         _?*r  \ 

=  e  a"\i  +  ^  *r  +  ^  ^  +  •  •  • ; 

a  ~  V&T 

(5.5) 

using  the  formula  for  the  sum  of  a  geometric  series, 

1  +  x  +  z2  +.--=- -  i- -  (5.6) 

We  note  that  at  high  temperatures,  the  numerator  of  Eq.  (5.5)  can  be  set 

equal  to  unity,  the  denominator  becomes  I  1  —  f  1  —  TTTJ  •  •  •    )     =  r™' 

L          \         Kl  /J       Ki 

so  that  the  partition  function  reduces  to  kT/hv,  in  agreement  with  the 
classical  value  (5.4).  Using  Eqs.  (5.5)  and  (3.5),  we  have  for  the  vibra- 
tional specific  heat  per  mole  the  value 


144 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  IX 


This  result  was  first  obtained  by  Einstein  and  is  often  called  an  Einstein 
function.  Introducing  the  characteristic  temperature  from  Eq.  (4.7),  we 
have 

Ovlh 


Cr          
vib    ~~~ 


v  (;¥-_-, 


(5.8) 


It  is  also  interesting  to  find  the  internal  energy  associated  with  the  vibra- 
tion; proceeding  as  in  Eq.  (5.20)  of 
Chap.  Ill,  we  see  this  at  once  to  be 


Nhv 


~T~ 


Nh 


(5.9) 


JOL 
ht; 


The  average  energy  and  heat  capacity 
per  oscillator,  from  Eqs.  (5.9)  and  (5.8), 
are  plotted  as  functions  of  temperature 
in  Fig.  IX-3.  It  will  be  seen  that  the 
energy  is  ^hv  at  the  absolute  zero  and 
increases  from  this  value  quite  slowly. 
The  slow  rise,  with  horizontal  tangent 
of  the  energy  curve  at-  the  absolute  zero, 
is  what  leads  to  the  vanishing  specific 
hczit  at  the  absolute  zero.  At  higher 
temperatures,  however,  the  energy 
approaches  the  classical  equipartition 
value  kT,  and  the  heat  capacity  ap- 
proaches the  classical  value  k. 

As  an  example  of  the  application  of 
FIG.    ix-;*.    Average   energy   and   Eq.  (5.8),  we  compute  the  vibrational 

heat  capacity  of  an  oscillator,  according  spocific  heat  for  CO,  ill  Table  IX-4, 
to  the  quantum  theory.  *  ' 

then  find  the   total   specific   heat    by 

adding  the  vibrational  heat  to  %R  =  6.96  cal.,  which  is  the  sum  of 
f R  for  translation,  R  for  rotation,  and  compare  with  the  correct  value. 
The  agreement  between  our  calculations  and  the  "correct"  values  of 
CP  in  Table  IX-4  is  good  but  not  perfect.  More  accurate  calculation 
agrees  practically  perfectly  with  experiment;  in  fact,  calculation  is  in 
general  a  more  accurate  method  than  the  best  experiments  for  finding  the 
specific  heat  of  a  gas,  and  the  "correct"  values  of  Table  IX-4  are  really 
simply  the  results  of  more  exact  and  careful  calculation  than  we  have 
made.  It  is  worth  while  discussing  the  errors  in  our  calculation.  In  the 
first  place,  the  frequency  v  which  we  have  found  from  the  constants  of 
the  Morse  curve  is  correct,  for  as  a  matter  of  fact  the  constants  a  in 
Table  IX-1  were  computed  from  the  frequencies  and  values  of  D  observed 


SEC.  6]  POLYATOMIC  GASES  145 

from  band  spectra,  using  Eq.  (4.5)  solved  for  a.  But  the  Einstein  specific 
heat  formula  (5.7)  is  not  exactly  correct  in  this  case,  for  the  actual  inter- 
atomic potential  is  not  simply  a  linear  restoring  force,  as  we  have  assumed 
when  we  use  the  theory  of  the  linear  oscillator.  Not  only  that,  but  as  wo 
have  mentioned  there  is  interaction  between  the  vibration  and  the  rota- 
tion of  the  molecule.  These  effects  make  a  small  correction,  which  can 
be  calculated  and  which  accounts  for  part  of  the  discrepancy  between  the 
last  two  columns  in  Table  IX-4.  They  do  not  account,  however,  for  the 

TABLE  IX-4.--  COMPUTED  SPECIFIC  HEVT  OF  CO 

r«   o  ,  i  Vib rational         ;        Total  specific,  Cp 

J  i        specific  heat         ,  heat  Cp  correct 


500 

i 
0.18 

7  14 

7  12 

1000 

0.94 

7  90 

7  94 

2000 

1  63 

8  59 

8  67 

3000 

1.81 

8.77 

8  90 

4000 

1  81) 

8  8T> 

9  02 

5000 

1  92 

8  88 

9  10 

i 

1 

Cp  is  given  in  calorics  per  mole.  Tho  calculation  is  made  by  Eq.  (5  8).  Vtiluos  tabulated  in  last 
column,  "Cp  correct,"  are  from  Landolt-Boinstem,  "  Physikahach-chemische  Tabellen,"  Drifter 
Erganzungsband,  Dnttor  Teil,  p.  2324,  Springer,  1936. 

fact  that  the  correct  specific  heat  rises  above  the  value  %R  —  8.94  cal., 
for  the  quantum  vibrational  specific  heat  never  rises  above  the  classical 
value.  This  effect  comes  in  on  account  of  a  ne\v  feature,  electronic 
excitation,  which  enters  only  at  very  high  temperature.  We  can  explain 
itrbnefiy  by  stating  that  electrons,  as  well  as  linear  oscillators,  can  exist 
in  various  stationary  states,  as_a  result  of  which  they  contribute  to  the 
specific  heat.  Their  specific  heat  curves  are  somewhat  similar  to  an  Ein- 
stem  curve^  but  with  extremely  high  p.hflTRo.t^rist.if.  tempfimfrirraTso  that 
even  at  5000°  they  are  at  the  very  low  part  of  the  curve  and  contribute 
onlv  slightly  to  the  specific  heat.  When  these  small  contributions  are 
computed  and  added  to  the  values  found  from  rotation  and  vibration,  the 
final  results  agree  very  accurately  with  observation. 

6.  The  Specific  Heats  of  Polyatomic  Gases. — We  shall  now  discuss 
the  specific  heats  of  polyatomic  gases,  without  going  into  nearly  the  detail 
we  have  used  for  diatomic  molecules.  In  the  first  place,  the  rotational 
kinetic  enerpry  is  different,  from  thai,  in  Hi«.toinin  molecules.  It  requires 
three,  rather  than  two,  coordinates  to  describe  the  orientation  of  a.  poly- 
atomic molecule  in  space.  Thus,  imagine  an  axis  rigidly  fastened  to  the 
molecule.  Two  coordinates,  say  latitude  and  longitude  angles,  are 
enough  to  fix  the  orientation  of  this  axis  in  space.  But  the  molecule  can 
still  rotate  about  this  axis,  and  an  additional  angle  must  be  specified  to 
determine  the  amount  which  it  has  rotated  about  the  axis.  These  three 


146  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  IX 

coordinates  all  have  their  momenta  and  their  terms  in  the  kinetic  energy. 
And  when  we  find  the  mean  kinetic  energy  of  rotation,  the  new  variable 
contributes  its  $kT,  according  to  equipartition.  ThiiH  ||»A  ^r^y,  t,rQr>«- 
lational  and  rotational,  amounts  to  3kT  per  mnWuluffir  3nRT  for  n  moles, 
and  the  transnational  and  rotational  heat  capacity  is  CV  =  3nB  =  5.96 
cal.  per  degree  per  mole,  C/>  =  4nR  =  7.95  cal.  per  degree  per  mole.  In 
addition  to  translational  and  rotational  energy,  the  polyatomic  molecules 
like  the  diatomic  ones  can  have  vibrational  energy.  As  a  matter  of  fact, 
they  can  havo  considerably  more  vibrational  energy  than  a  diatomic 
molecule,  for  they  havo  more  vibrational  degrees  of  freedom.  A  diatomic 
molecule  has  only  one  mode  of  vibration,  but  a  triatomic  molecule;  has 
throe.  Thus  the  water  molecule  can  vibrate  in  the  three  ways  indicated 


(a)  5170°  Abs  (b)  5400°  Abs  (c)  2290°  Abs 

FIG.  IX-4.-  Modes  ol  vibration  of  the  1LO  molnculc  Tho  arro\vs  indicate  the  direc- 
tion of  vibration  ot  each  atom,  for  the  normal  mode  whose  characteristic  temperature  IP 
indicated.  For  similar  information  on  a  variety  of  molecules,  see  11.  Sponer,  "  Molekulspek- 
tren  und  ihre  Anwfndungen  aiif  chomisrhe  Problems,"  Springer,  Berlin,  1935. 

by  the  arrows  of  Fig.  IX-4.  The  arrowy  show  the  directions  of  displace- 
ment of  the  throe  atoms  in  the  vibration.  In  general,  to  find  the  number 
of  vibrational  degrees  of  freedom,  it  can  be  shown  that  one  takes  all  the 
degrees  of  freedom  of  the  atoms  of  the  molecule,  regarded  as  free.  This  is 
3N,  if  there  are  N  atoms,  each  having  three  rectangular  coordinates. 
Then  one  subtracts  from  this  total  the  number  of  other  degrees  of  free- 
dom: the  translational  degrees  of  freedom  of  the  molecule  as  a  whole, 
three,  and  the  rotational  degrees  of  freedom  (none  when  N  =  1,  two  when 
N  =  2,  three  when  N  ^  3).  Thus  the  number  of  vibratiomil  degrees  of 
freedom  is 

3(1)  —3  —  0=0  for  a  monatomie  gas, 

3(2)  —  3  —  2  =  1  for  a  diatomic  gas, 

3(3)  —  3  —  3  =  3  for  a  triatomic  gas,  and  in  general 

3N  -  Gfortf  ^  3.  (6.1) 

Each  of  the  vibrational  degrees  of  freedom  given  by  Eq.  (6.1)  would 
have  a  mean  kinetic  and  potential  energy  of  kT,  according  to  equiparti- 
tion, and  would  contribute  an  amount  R  to  the  specific  heat.  As  with 
diatomic  molecules,  however,  the  quantum  theory  tells  us,  and  we  find 


SEC.  6] 


POLYATOMIC  GASES 


147 


experimentally,  that  the  vibrational  specific  heat  is  practically  zero  at 
low  temperatures.  We  give  a  single  example,  the  specific  heat  of  water 
vapor,  which  will  show  what  actually  happens.  Since  this  is  a  triatomic 
molecule,  the  specific  heat  should  be  Cp  =  ±nR  =  7.95  cal.  per  degree  per 
mole  for  translation  and  rotation,  plus  three  Einstein  terms,  as  given  by 
Eq.  (5.8),  for  characteristic  temperatures  which  are  5170°  abs.,  5400°  abs., 
and  2290°  abs.,  for  the  modes  of  vibration  (a),  (b),  mid  (e)  respectively 
in  Fig.  IX-4.1  Calculations  are  given  in  Table  IX-5,  where  the  columns 

TABLE  IX-5. — COMPUTED  SPECIFIC  HEAT  OF  WATEK  VAI-OR 


T,  °  abs. 

(«) 

W 

(c) 

:  CP, 

calculated 

0-, 
correct 

300 

0  00 

0  00 

0  10 

8  05 

8  (X) 

400 

0  00 

0  00 

0  25 

8  20 

8  16 

500 

0  00 

0  00 

0.40 

8  35 

8  38 

600 

0  02 

0  02 

0  66 

8  63 

8  64 

800 

0  16 

0  15 

1  05 

9  31 

9  20 

1000 

0  31 

0  30 

1  30 

9.86 

9  80 

1500 

0  80 

0  78 

1  65 

11  18 

11  15 

2000 

1  18 

1  15 

1  79 

12  07 

12  09 

3000 

1  58 

1  52 

[  90 

12  90 

13  10 

See  fomments  under  Table  IX-4 

headed  (a),  (b),  (c)  are  the  vibrational  heat  capacities  for  the  three  modes 
of  vibration,  the  next  column  gives  the  calculated  CP,  and  the  last  one  the 
correct  (V,  found  from  more  accurate  calculation  and  agreeing  well  with 
experiment.  As  with  CO,  the  slight  discrepancies  remaining  between  our 
calculation  and  the  correct  values  can  be  removed  by  more  elaborate 
methods,  including  the  interaction  between  vibration  and  rotation,  and 
electronic  excitation. 

The  calculation  we  have  just  made  is  based  on  the  assumption  that 
the  vibrations  of  the  molecule  are  simple  harmonic,  the  force  being  prft- 
portional  to  the  displacement  and  the  potential  energy  to  the  square  of 
the  displacement.  Ordinarily  this  is  a  fairly  good  approximation  for  the 
amplitudes  of  vibration  met  at  ordinary  temperatures,  but  there  are  some 
important  cases  where  this  is  not  true.  An  example  is  found  in  the 
so-called  phenomenon  of  hindered  rotation.  There  are  some  molecules,  of 
which  ethane  CHa-CHa,  shown  in  Fig.  IX-5,  is  an  example,  in  which  one 
part  of  the  molecule  is  almost  free  to  rotate  with  respect  to  another  part. 
Thus,  in  this  case,  one  CHs  group  can  rotate  with  respect  to  the  other 
about  the  line  joining  the  carbons  as  an  axis.  The  rotation  would  be 

1  See  Sponer,  "  Molekiilspektren.  und  ihre  Anwendungen  auf  chcmische  Probleme," 
Vol.  I,  Springpr,  1935,  for  vibrational  frequencies  of  this  and  other  molecules. 


148  INTRODUCTION  TO  CHEMICAL  I'HYSICti  [CHAP.  IX 

perfectly  free  if  the  potential  energy  were  independent  of  the  angle  of 
rotation  0,  so  that  there  were  no  torques.  Then  as  far  as  this  degree  of 
freedom  was  concerned,  there  would  be  no  potential  energy,  so  that  the 
mean  energy  on  the  classical  theory  would  be  ?kT,  the  kinetic  energy, 
rather  than  kT,  the  sum  of  the  kinetic  and  potential  energies,  as  in  an 
oscillator.  Actually  in  such  cases,  however,  there  are  slight  torques,  with 
a  periodicity  in  the  case  of  ethane  of  120°  in  6.  These  arise  presumably 
from  repulsions  between  the  hydrogen  atoms  of  the  two  methyl  groups, 
suggesting  that  the  potential  energy  might  have  a  maximum  value  when 
the  hydrogens  in  the  two  groups  were  opposite  each  other,  and  a  minimum 
when  the  hydrogens  of  one  group  were  opposite  the  spaces  between  hydro- 
gens in  the  other.  In  such  a  case,  for  small  energies,  the  motion  would 


Fio.  IX-5      Tho  ethane  molecule,  OII3-CH., 

be  an  oscillation  about  one  of  the  minima  of  the  potential  energy  curve, 
while  for  larger  energies,  greater  than  the  maximum  of  the  potential 
energy  curve,  the  motion  would  be  a  rotation,  but  not  with  uniform  angu- 
lar velocity.  In  such  a  case,  in  the  classical  theory,  the  mean  kinetic 
energy  would  equal  \kT  in  any  case.  The  mean  potential  energy,  how- 
ever, would  increase  as  \kT  for  low  temperatures,  where  the  motion  was 
oscillatory,  but  would  approach  a  limiting  value,  equal  to  the  mean 
potential  energy  over  all  angles,  which  it  would  practically  reach  at  the 
temperatures  at  which  most  of  the  molecules  were  rotating  rather  than 
oscillating.  Thus  the  heat  capacity  per  molecule  connected  with  this 
degree  of  freedom  would  be  k  at  low  temperatures,  but  would  fall  to 
?k  at  higher  temperatures  where  the  rotation  became  more  nearly  free. 
With  the  quantum  theory,  of  course  the  heat  capacity  would  resemble 
that  of  an  oscillator  at  low  temperatures,  starting  from  zero  at  the  abso- 
lute zero,  then  rising  to  the  neighborhood  of  &,  but  falling  to  $k  at  high 
temperatures  as  in  the  classical  case.  Measurements  and  calculations  of 
specific  heat  and  entropy  of  molecules  which  might  be  expected  to  show 
free  rotation  of  one  part  with  respect  to  another  generally  seem  to  indicate 


SEC.  6]  POLYATOMIC  GASES  149 

that  at  ordinary  temperatures  the  rotations  are  really  hindered  by  periodic 
torques  in  this  way,  the  heat  capacity  being  more  like  that  of  an  oscillator 
than  that  of  a  rotator.  It  is  clear  that  from  measurements  of  the  specific 
heat  one  can  work  backward  and  find  useful  information  about  the  magni- 
tude of  the  torques  hindering  the  rotations,  and  hence  about  the  inter- 
atomic forces. 


CHAPTER  X 

CHEMICAL  EQUILIBRIUM  IN  GASES 

In  Chap.  VIII  we  treated  mixtures  of  gases  in  which  the  concentra- 
tions were  determined.  Now  we  take  up  chemical  equilibrium,  or  the 
problem  of  mixtures  of  gases  which  can  react  with  each  other,  so  that  the 
main  problem  is  to  determine  the  concentrations  of  the  various  gases  in 
equilibrium.  In  this  problem,  as  in  all  cases  of  chemical  reactions,  there 
are  two  types  of  question  that  we  may  ask.  In  the  first  place,  there  is  the 
rate  of  reaction.  Given  two  gases  capable  of  reacting  and  mixed  together, 
how  fast  will  the  reaction  occur  and  how  will  this  rate,  depend  on  pressure 
and  temperature?  In  the  second  place,  there  is  the  question  of  equilib- 
rium. To  every  reaction  there  is  a  reverse  reaction,  so  that  the  final 
state  of  equilibrium  will  represent  a  balance  between  the  direct  and  the 
reverse  reactions,  with  definite  proportions  of  all  the  substances  in  the 
equilibrium  mixture.  We  may  wish  to  know  what  these  proportions  are. 
The  first  type  of  problem,  the  rate  of  reaction,  can  be  answered  only  by 
kinetic  methods.  Gas  reactions  take,  place  only  when  the  reacting  mole- 
cules are  in  collision  with  each  other,  and  only  when  the  colliding  mole- 
cules happen  to  have  a  good  deal  more  than  the  average  energy.  Thus 
to  find  the  rate  of  reaction  we  must  investigate  collisions  in  detail  and 
must  know  a  great  deal  about  the  exact  properties  of  the  molecules.  In 
almost  no  case  do  we  know  enough  to  calculate  a  rate  of  reaction  directly 
from  theory.  We  can,  however,  find  how  the  rate  of  reaction  depends 
on  the  concentrations  of  the  various  substances  present  in  the  gas,  and 
even  this  small  amount  of  information  is  useful.  It  allows  us  to  use  the 
kinetic  method  to  find  the  concentration  of  substances  in  equilibrium,  for 
we  can  simply  apply  the  condition  that  the  concentrations  are  such  that 
they  do  not  change  with  time,  and  this  gives  us  equations  leading  to  the 
so-called  mass  action  law.  The  results  we  find  in  this  way,  however,  are 
incomplete.  They  do  not  tell  us  how  the  equilibrium  changes  with  tem- 
perature, a  very  important  part  of  the  problem.  Fortunately,  these* 
questions  of  equilibrium  can  be  answered  completely  by  the  method  of 
thermodynamics  and  statistical  mechanics.  For  in  equilibrium,  the 
Gibbs  free  energy  of  the  mixed  gas  must  have  the  minimum  value  possible, 
and  this  condition  leads  not  merely  to  the  mass  action  law  but  to  complete 
information  about  the  variation  of  the  equilibrium  with  temperature.  As 
usual,  thermodynamics  gives  us  more  complete  and  satisfactory  informa- 

150 


SBC.  1]  CHEMICAL  EQUILIBRIUM  IN  GASES  151 

tion,  but  about  a  more  restricted  problem,  that  of  thermal  equilibrium. 
In  our  discussion  to  follow,  we  shall  start  with  the  kinetic  method,  speak- 
ing about  the  mechanism  of  gas  reactions  and  carrying  the  method  as 
far  as  we  can.  Then  we  shall  take  up  the  thermodynamic  treatment, 
deriving  the  conditions  of  equilibrium,  and  finding  the  interesting  fact 
that  the  chemical  constants  of  gases,  introduced  previously  in  connection 
with  the  entropy,  are  fundamental  in  the  study  of  chemical  reactions. 
1.  Rates  of  Reaction  and  the  Mass  Action  Law.  —  Let  us  write  a  simple 
chemical  equation;  for  instance, 

2H2  +  02^2H20,  (1.1) 

describing  the  combination  of  hydrogen  and  oxygen  to  form  water,  and 
the  reverse,  the  dissociation  of  water  into  hydrogen  and  oxygen.  The 
equation  expresses  the  fact  that  when  two  molecules  of  H2  and  one  of  ()2 
disappear,  two  of  H2O  appeal-;  or  vice  versa.  Now  lot  us  form  the  sim- 
plest kinetic  picture  of  the  reaction  that  we  can.  For  the  combination  of 
two  hydrogens  and  an  oxygen  to  form  two  water  molecules,  we  suppose  in 
the  first  place  that  a  triple  collision  of  the  two  hydrogens  and  the  one 
oxygen  molecule  is  necessary;  we  suppose  further  that  in  a  certain  fraction 
of  such  collisions,  a  fraction  which  may  depend  on  the  temperature,  the 
three  molecules  react.  Thus  the  number  of  sets  of  molecules  reacting 
per  unit  time  will  be  proportional  to  the  number  of  triple  collisions  per 
unit  time.  This  number  of  collisions  in  turn  will  be  proportional  to  the 
number  of  oxygen  molecules  per  unit  volume  and  to  the  square  of 
the  number  of  hydrogens  per  unit  volume.  For  plainly  if  we  double  the 
number  of  oxygens,  we  double  the  chance  that  one  will  be  found  at  tho 
point  where  the  collision  will  take  place;  while  if  we  double  the  number 
of  hydrogens,  we  double  the  chance  that  one  hydrogen  will  be  found  at 
the  location  of  the  collision,  and  furthermore  we  double  the  chance  that, 
if  one  is  there,  another  will  also  be  found  on  hand.  Since,  at  a  given 
temperature,  the  number  of  molecules  per  unit  volume  is  proportional  to 
the  pressure,  we  find  for  the  number  of  sets  of  molecules  that  react  per 
second 

CCDPnVV  (1.2) 

Here  (K^m^a^CQefficient  depending  on  the  size  of  the  molecules,,  their 
velocities,  the  probability  that  if  they  collide  they  will  react,  etc.  The 
quantities  PU2  and  Pot  are  the  partial  pressures  of  H2  and  02;  that  is, 
they  are  the  pressures  which  these  gases  would  exert  by  themselves,  if 
their  molecules  only  were  occupying  the  volume.  It  is  the  evaluation  of 
C(T)  as  a  function  of  temperature  which,  as  we  have  previously  suggested, 
is  almost  prohibitively  difficult  by  purely  theoretical  methods. 


152  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  X 

At  the  same  time  that  direct  reactions  are  taking  place,  there  will  be 
reverse  reactions,  dissociations  of  water  molecules  to  produce  hydrogens 
and  oxygons.  From  the  chemical  equation  (1.1)  we  see  that  two  water 
molecules  must  be  present  in  ordor  to  furnish  the  necessary  atoms  to  break 
up  into  hydrogen  and  oxygen  molecules.  Thus,  by  the  type  of  argument 
we  have1  just  used,  the  rate  of  the4  reverse  reaction  must  be  proportional  to 
the  square  of  the  number  of  water  molecules  per  unit  volume  or  to  the 
square  of  the  partial  pressure  of  water;  we  may  write  it  as 

C'(T)P?M.  (1.3) 

Suppose  we  start  with  only  hydrogen  and  oxygen  in  the  container,  with 
no  water  vapor.  Reactions  will  occur  at  a  rate  given  by  Kq.  (1.2), 
producing  water.  As  this  happens,  the  oxygen  and  hydrogen  will  be 
gradually  used  up,  so  that  their  partial  pressures  will  decrease  and  the 
number  of  molecules  reacting  per  unit  time  will  diminish.  At  the  same 
time,  molecules  of  water  will  appear,  so  that  the  partial  pressure  of  water 
will  build  up  and  with  it  the  number  of  dissociation  reactions  given  by 
Kq.  (1.3),  in  which  water  dissociates  into  hydrogen  and  oxygen.  This 
will  tend  to  diminish  the  amount  of  water  and  increase  the  amount  of 
hydrogen  and  oxygen,  until  finally  an  equilibrium  will  occur,  with  sta- 
tionary amounts  of  the  various  gases,  though  individual  molecules  an; 
reacting,  changing  from  water  vapor  to  oxygen  and  hydrogen  and  back 
again  with  great  rapidity,  l,p  equilibrium,  frhp  ^ymhnr  ^f  reactions  of 
type  (1.2)  must  just  equal  the  number  of  type  (1.3)  per  unit  time.  Thus 
we  must  have 


or 


where  Kp(T)  is  a  function  of  temperature,  the  subscript  P  indicating  the 
fact  that  Eq.  (1.4)  is  stated  in  terms  of  partial  pressures  (we  shall  pres- 
ently state  it  in  a  slightly  different  way).  Eq.  (1.4)  expresses  the  law  of 
mass  fie ti on  for  tKB-Dnrtip.uliir  reaction  in  Question. 

From  Eq.  (1.4)  we  can  derive  information  about  the  effect  of  adding 
hydrogen  or  oxygen  on  the  equilibrium.  Thus  suppose  at  a  given  tem- 
perature there  is  a  certain  amount  of  water  vapor  in  equilibrium  with  a 
certain  amount  of  hydrogen  and  oxygen.  Now  we  add  more  hydrogen 
and  ask  what  happens.  In  spite  of  adding  hydrogen,  the  left  side  of 
Eq.  (1.4)  must  stay  constant.  If  the  hydrogen  did  not  combine  with 
oxygen  to  form  water,  JPH2  would  increase,  the  other  P's  would  stay  con- 
stant and  the  expression  (1.4)  would  increase.  The  only  way  to  prevent 
this  is  for  some  of  the  added  hydrogen  to  combine  with  some  of  the 


/    /  P&Po*       C'(T) 

'  -    ^ 


SEC.  1]  CHEMICAL  EQUILIBRIUM  IN  GASES  153 

oxygen  already  present  to  form  some  additional  water.  This  will  decrease 
both  terms  in  the  numerator  of  Eq.  (1.4),  increase  the  denominator,  and 
so  bring  back  the  expression  to  its  original  value.  Information  of  this 
type,  then,  can  be  found  directly  from  our  kinetic  derivation  of  the  mass 
action  law.  But  we  should  know  a  groat  doal  more  if  we  could  calculate 
KP(T),  for  thru  we  could  find  the  actual  amount  of  dissociation  and  its 
variation  with  pressure  and  temperature. 

It  is  easy  to  formulate  the  mass  action  law  in  the  general  case,  by 
analogy  with  what  we  have  done  for  our  illustrative  reaction.  In  the 
first  place,  let  us  write  our  chemical  equations  in  i\  standard  form. 
Instead  of  Eq.  (1.1),  we  write 

2H2  +  02  -  2H2O  =  0.  (1.5) 

We  understand  Eq.  (1.5)  to  mean  that  two  molecules  (or  moles)  of  hydro- 
gen, one  of  oxygen,  appear  in  the  reaction,  while  two  molecules  (or  moles) 
of  water  disappear.  The  reverse  reaction,  according  to  this  convention, 
would  be  written  with  opposite  sign.  We  write  our  general  chemical 
equation  by  analogy  with  Eq.  (1.5),  each  symbol  having  an  integral 
coefficient  v,  giving  the  number  of  molecules  (or  moles)  of  the  correspond- 
ing substance  appearing  in  the  reaction,  negative  r's  corresponding  to 
the  disappearance  of  a  substance.  Let  there  be  a  number  of  substances, 
denoted  by  1,  2,  .  .  .  (as  112,  O2,  H2O  in  the  example),  with  correspond- 
ing v\,  j>2,  .  .  .  ,  and  partial  pressures  PI,  P2,  .  .  .  Then  it  is  clear  by 
analogy  with  our  example  that  the  general  mass  action  law  can  be  stated 

7V'/Y«  '  '  '    =  Kp(T).  (1.6) 

Here  the  terms  with  negative  v'$  automatically  appear  in  the  denomina- 
tor, as  they  should  from  Eq.  (1.4). 

It  is  often  convenient  to  restate  Eq.  (1.6),  not  in  terms  of  partial 
pressures,  but  in  terms  of  the  number  of  moles  of  each  substance  present, 
or  in  terms  of  fractional  concentrations.  Thus  let  there  be  HI  moles 
of  the  first  substance,  ?i2  of  the  second,  etc.  Then  we  have 


by  the  perfect  gas  law,  where  V  is  the  volume  occupied  by  the  mixture  of 
gases.     Substituting  in  Eq.  (1.6),  we  have 


(1-8) 

EQ  nation  (1.8.)  is  con^ciiiQiilJJiiLfinding  the  effect  iif_a_£iiaiige  of  volume 
on  the  equilibrium.  For  example,  in  our  case  of  water,  from  Eq.  (1.5), 
v\  +  *>2  +  •  •  •  =  )>H2  +  ^02  +  ''Hao  =  2+1—2=1.  Thus  we  have 


154  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  X 


aftQa   _        py 

2      ~~~    ' 


a  quantity  proportional  to  the  volume.  Now  let  the  volume  be  changed 
at  constant  temperature.  If  the  volume  increases,  the  numerator  must 
increase,  showing  that  there  must  be  dissociation  of  water  into  hydrogen 
and  oxygen.  On  the  other  hand,  decrease  of  the  volume  produces  recom- 
bination. This  is  a  special  case  of  the  general  rule  which  is  seen  to  follow 
from  Eq.  (1.8)  :  decrease  of  volume  makes  the  reaction  run  in  the  direction 
to  reduce  the  total  number  of  moles  of  gas  of  all  sorts.  In  our  special 
case,  if  two  moles  of  hydrogen  and  one  of  oxygen  combine  to  give  two 
moles  of  water  vapor,  there  is  one  mole  of  gas  less  after  the  process  than 
before.  It  seems  reasonable  that  decrease  of  volume  should  force  the 
equilibrium  in  this  direction. 

It  is  also  useful  to  write  the  mass  action  law  in  terms  of  the  relative 
concentrations  of  the  gases.  From  Eq.  (1.6),  using  Eq.  (2.4)  of  Chap. 
VIII,  or  Pl  =  tftP,  where  c,  is  the  relative  concentration  of  the  M\  gas,  we 
have 

'i'1*'1  '  '  '    =  /£^T}     =  K(l\  T).  (1.10) 

Equation  (1.10)  is  convenient  for  finding  the  effect  of  pressure  on  the 
equilibrium,  as  Eq.  (1.8)  was  for  finding  the  effect  of  volume.  Thus  in 
the  case  of  the  dissociation  of  water  vapor,  we  have 


showing  that  increasing  pressure  increases  the  concentration  of  water 
vapor.  Of  course,  this  is  only  a  different  form  of  stating  the  result  (1.9), 
but  is  generally  more  useful. 

2.  The  Equilibrium  Constant,  and  Van't  Hoff's  Equation.  —  In  the 
preceding  section  we  derived  the  mass  action  law,  but  have  not  evaluated 
the  equilibrium  constant  KP(T)  or  K(P,  T).  Now  WTC  shall  carry  out 
our  thermodynamic  discussion,  leading  to  a  derivation  of  this  constant. 
The  method  is  clear:  we  remember  from  Chap.  II,  EW  3 


free  energy  is  a  minimum  for  a  system  at  constant  pressure  and  tempera- 
tureT  Then  we  find  the  Gibbs  free  energy  G  of  the  mixture  of  gases,  and 
vfljv  fhfTpnnnpnf  rations  to  rrmke  it  a  minimum.  From  Eq.  (2.15)  of 
Chap.  VIII,  we  have 


(J  -       w/r'/  +  KTn,  In  c,.  (2.1) 


For  equilibrium,  we  must  find  the  change  of  G  when  the  numbers  of  moles 
of  the  various  substances  change,  and  set  this~shange  equat'To  "zero. 


SEC.  21  CHEMICAL  EQUILIBRIUM  IN  GASES  155 

Using  Eq.  (2.1),  we  have 

dG  =  2J(G,  +  RT  In  a)dn,  +  Jn,  d(G]  +  RT  In  c,)  =  0.     (2.2) 
y  ; 

The  second  sum  is  zero.  In  the  first  place,  the  Or/s  do  not  depend  on  the 
n/s,  so  that  they  do  not  change  when  the  n/s  are  varied.  For  the  con- 
concentrations,  we  have  d  In  Cj  =  dcjcj.  Hence  the  last  term  becomes 

5jfir(n//c,)<fc,.     But  by  Eq.  (2.1),  Chap.  VIII,  n,/c,  =  ni  +  n,  +  •  •  •  , 

^y 

independent  of  j,  so  that  the  summation  is  really 


Furthermore,  by  Eq.  (2.6),  Chap.  VIII,  Sc/  =  1,  so  that 

dSc/  =  2  dc;  =  0, 
being  the  change  in  a  constant.     Hence,  we  have  finally 

dG  =       (G,  +  fiZ1  In  c,)dn/  =  0  (2.3) 


as  the  condition  of  equilibrium.  But  from  the  chemical  equation  wo 
know  that  the  piimi^ej:jiLmQle^iii[es  yf  thq  ^tJrJzyiKi_aiiDcai'iii^  in  an  actual 
reaction  must  be  proportional  to  vt*  the  coefficient  appearing  in  tho 
chemical  equation.  Hejp.fi  the  dn.'*  ipnst  ho  nroportional  to  tho  v/s,  and 
we  may  rewrite  Eq.  (2.3)  as 

(0,  +  RT  In  c,)  =  0.  (2.4) 


Taking  the  exponential  and  putting  all  terms  involving  the  c's  on  the  left, 
the  others  on  the  right,  we  have 


=  K(P,  T),  whew 

(2.5) 


\  J 


In  Eq.  (2.5)  weJiave  found  the  saine^na^a_actiQrL-IawLJia  in  Ea.  (1.1UL 
but^^ilJi^^complete  evaluatioiLJiLJJie  equilibrium  constant  K.  Using 
Eq.  (2.16),  Chap.  VIII,  for  G,,  we  verify  at  once  that  K(P,  T)  varies  with 
P  as  in  Eq.  (1.10),  and  we  find 

In  Kf(T)  -  -       >          -     In  T  -    T  ^  dr  -  ^.     (2.6) 


156  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  X 

In  Eq.  (2.6),  U}  is  the  arbitrary  additive  constant  jciying  the  energy  of 
the  /th  gas  per  mole  at  the  absolute  zero,  C,  is  the  heat  capacity  per  mole 
of  the  jth  gas  coming  from  rotations  and  vibrations,  and  ?';  is  tho  chemical 
constant  of  ihc  j't 


There  is  an  important  relation  connecting  the  change  of  either  K 
or  KP  with  temperature  and  a  quantity  called  the  heat  of  reaction.  By 
definition,  the  heat  of  reaction  is  the  heat  absorbed.  ™»  tihp  ""*™*QS^  of 
enthalpy  A/7,  when  the  reaction  proceeds  reversibly  so  that  v\  moles  of 
the_first  type  of  molecule  are  produced,  v*  of  tho  second,  etc.T  at  constant 
pressure?  and  temperature.  From  Eqs.  (2.13),  (2.14)  of  Chap.  VIII,  this 
is  at  once  seen  to  be 


c,  dT). 


(2.7) 


Now  let  us  find  the  change  of  In  K(P,  T)  or  In  Kr(T)  with  temperature. 
Differentiating  Eq.  (2.6),  we  have 


/a 
\ 


_ 
W~  ~       "       i  VRT* 


Equation  (2.8)  is  called  Van't  Hoff's  equation  and  is  a  very  imnortant 
one  in  physical  chemistry.  It  can  be  shown  at  once  that  the  same  equa- 
tion holds  for  K(P,  T). 

Van't  Hoff's  equation  can  be  used  in  cither  of  two  ways.  First,  we 
may  know  the  heat  of  reaction,  from  thermal  measurements,  and  we  may 
then  use  that  to  find  the  slope  of  the  curve  giving  K(P,  T)  against  tem- 
perature. Let  us  see  which  way  this  predicts  that  the  equilibrium  should 
be  displaced  by  increasing  temperature.  Suppose  that  heat  is  absorbed 
in  the  chemical  reaction,  so  that  A//  is  positive.  Then  the  constant 
K  (P,  T)  will  increase  with  temperature.  That  means  that  at  high 
temperatures  more  of  the  material  is  in  the  form  that  requires  heat  to 
produce  it.  For  instance,  to  dissociate  water  vapor  into  hydrogen  and 
oxygen  requires  heat.  Therefore  increase  of  temperature  increases  the 
amount  of  dissociation.  In  the  second  place,  we  may  use  Van't  HofFs 
equation  to  find  the  heat  of  reaction,  if  the  change  of  equilibrium  constant 
with  temperature  is  known.  This,  as  a  matter  of  fact,  is  one  of  the  com- 
monest ways  of  measuring  heats  of  reaction  in  physical  chemistry. 

The  heat  of  reaction  at  the  absolute  zero,  from  Eq.  (2.7),  is 

E/,.  (2.9) 


SBC.  2]  CHEMICAL  EQUILIBRIUM  IN  GASES  157 

It  is  interesting  to  see  that  this  can  be  calculated  from  the  quantities  D 
of  Sec.  1,  Chap.  IX.  From  Table  IX- 1  we  know  the  value  of  Z),  the 
energy  required  to  dissociate  various  diatomic  molecules,  and  similar 
values  can  be  given  for  polyatomic  molecules.  Thus  lot  us  consider  our 
case  of  the  dissociation  of  water  vapor.  To  remove  one  hydrogen  atom 
from  an  H^O  molecule  requires  118  kg.-cal.  per  mole  (not  given  in  the 
table),  and  to  remove  the  second  hydrogen  from  the  remaining  OH 
molecule  requires  102,  a  total  of  220  kg.-cal.  In  our  reaction,  there  are 
two  1120  molecules,  requiring  440  kg.-cal.  to  dissociate  them  into  atoms. 
That  is,  440  kg.-cal.  are  absorbed  in  this  process.  But  now  imagine  the 
four  resulting  hydrogen  atoms  to  combine  to  form  t\\o  Ha  molecules  and 
the  two  oxygens  to  combine  to  form  ( )2.  Each  pair  of  hydrogens  liberates 
103  kg.-cal.  in  recombining,  a  total  of  200  kg.-cal..  and  the  two  OXV^MIS 
liberate  117  kg.-cal.,  so  that  200  +  1 17  =  323  kg.-cal.  are  liberated  in  this 
part  of  the  process.  The  net  result  is  an  absorption,  of  440  —  323  ==  117 
kg.-cal.,  so  that  A//0  is  117  kg.-cal.  This  is  in  fairly  good  agreement 
with  the  experimental  value  of  about  113  kg.-cal.  It  is  interesting  to 
notice  that  the  final  result  is  the  difference  of  two  fairly  large1  quantities, 
so  that  relatively  small  errors  in  the  7)'s  can  result  in  a  rather  large  error 
in  the  heat  of  reaction. 

The  calculation  which  we  have  just  made  for  A//0  does  not  follow 
exactly  the  pattern  of  Eq.  (2.9).  To  see  just  how  that  equation  is  to  be 
interpreted,  we  must  give  values  to  the  various  ^/s.  In  general,  since 
there  is  an  undetermined  constant  in  any  potential  energy,  we  can  assign 
the  f//s  at  will.  But  there  is  a  single  relation  between  them,  on  account 
of  the  possibility  of  formation  of  water  from  hydrogen  and  oxygen.  Let 
Uut  be  the  energy  per  mole  of  hydrogen  at.  tin*  absolute  zero,  f/o_.  of 
oxygen,  UK&  of  water,  all  of  course  in  the  vapor  state1.  Then  from  the, 
last  paragraph  we  know  that  the  energy  of  two  moles  of  hydrogen,  plus 
that  of  one  mole  of  oxygen,  is  117  kg.-cal.  greater  than  the  energy  of  two 
moles  of  water  vapor.  That  is, 

2Um  +  l;o>  =  2tV,o  +  117  kg.-cal.  (2.10) 

Statements  like  Eq.  (2.10)  are  sometimes  written  in  combination  with 
the  chemical  equation,  in  a  form  like 

2H2  +  02  =  2H2O  +  117  kg.-cal.  (2.11) 

Aside  from  Eq.  (2.10),  the  U,'s  can  be  chosen  freely;  that  is,  any  two  of 
them  can  be  chosen  at  will  and  then  the  third  is  determined.  Now  let 
us  compute  A//0,  using  Eqs.  (2.9)  and  (2.10).  It  is 

A//o  =  217,1,  +  C7o2  -  2C7Hlo 

=  2C7H20  +  117  kg.-cal.  -  2Ulli0 

=  117  kg.-cal.,  (2.12) 


158  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  X 

in  agreement  with  our  previous  value.  From  this  example  it  is  clear 
that  all  the  undetermined  constants  among  the  t/o's  cancel  from  the  sum 
in  Eq.  (2.9),  leaving  a  uniquely  determined  value  of  A//0. 

We  have  just  seen  that  a  knowledge  of  the  heats  of  dissociation,  or 
constants  D,  of  the  various  molecules  concerned  in  a  reaction  allows  us  to 
find  A//o,  the  heat  of  reaction  at  the  absolute  zero.  A  further  knowledge 
of  the  specific  heats  of  the  molecules  gives  us  all  the  information  we  need 
to  find  the  equilibrium  constant  KP(T),  according  to  Kq.  (2.6),  except  for 
the  final  constant  Siv';.  In  other  words,  this  knowledge  is  enough  to  find 
the  rate  of  change  of  In  K /•('/')  with  temperature,  according  to  Eq.  (2.8), 
but  not  enough  to  determine  the  constant  of  integration  of  the  integrated 
equation  (2.6).  But  we  have  seen  in  Chap.  VIII  how  to  find  the  con- 
stants i  theoretically,  find  later  we  shall  see  how  to  find  them  experimen- 
tally from  vapor  pressure  measurements.  We  now  see  why  these 
constants  are  so  important  and  why  they  are  railed  chemical  constants: 
they  determine  the  constants  of  integration  for  problems  of  chemical 
equilibrium.  For  this  reason,  a  great  deal  of  attention  has  gone  to  finding 
accurate  values  for  them. 

3.  Energies  of  Activation  and  the  Kinetics  of  Reactions.  A  curious 
fact  may  have  struck  the  reader  in  connection  with  the  example  which 
we  have  used,  the  equilibrium  between  water  vapor  and  hydrogen  and 
oxygen.  Calculating  the  equilibrium,  we  find  that  at  room  temperature 
the  amount  of  hydrogen  and  oxygen  in  equilibrium  with  water  vapor  is 
entirely  negligible;  even  at  several  thousand  degrees  only  a  few  per  cent 
of  the  water  vapor  is  dissociated.  This  certainly  accords  with  our  usual 
experience  with  steam,  which  does  not  dissociate  into  hydrogen  and 
oxygen  in  steam  engines.  And  yet  if  hydrogen  and  oxygen  gases  are 
mixed  together  in  a  container  at  room  temperature,  they  will  remain 
indefinitely  without  anything  happening.  A  spark  or  other  such  dis- 
turbance is  required  to  ignite  them;  as  a  result  of  ignition,  of  course,  a 
violent  explosion  results,  the  hydrogen  and  oxygen  being  practically 
instantaneously  converted  into  water  vapor.  The  heat  of  combustion, 
which  is  the  sume  thing  as  the  heat  of  reaction  of  117  kg.-cal.  which  we 
have  just  computed,  is  a  very  large  one  (one  of  the  largest  for  any  common 
reaction);  since  an  explosion  is  an  adiabatic  process,  this  heat  cannot 
escape,  but  will  go  into  raising  the  temperature,  and  consequently  the 
pressure*,  of  the  resulting  water  vapor  enormously.  It  is  this  sudden 
rise  of  pressure  and  temperature  that  constitute  the  explosion.  But  now 
we  ask,  wrhy  was  the  spark  necessary?  Why  do  not  the  hydrogen  and 
oxygen  combine  immediately  when  they  are  placed  in  contact? 

Our  first  supposition  might  be  that  the  triple  collisions  of  two  hydro- 
gen molecules  and  one  oxygen,  which  we  have  postulated  as  being  neces- 
sary for  the  reaction,  were  rare  events.  But  this  is  not  the  case. 


SEC.  3]  CHEMICAL  EQUILIBRIUM  IN  GASES  159 

Calculation,  taking  into  account  the  cross  section  of  the  molecules,  shows 
that  at  ordinary  temperatures  and  pressures  there  will  be  a  tremendous 
number  of  such  collisions  per  unit  time.  The  only  remaining  hypothesis 
is  that  even  when  two  hydrogen  molecules  and  an  oxygen  are  in  the 
intimate  contact  of  a  collision,  still  it  is  such  a  rare  thing  for  their  atoms 
to  rearrange  themselves  to  form  two  water  molecules  that  for  all  practical 
purposes  it  never  happens.  This  is  indeed  the  case.  The  proportion  of 
all  such  triple  collisions  in  which  a  reaction  takes  place  is  excessively 
small,  at  ordinary  temperatures,  though  it  is  finite;  if  we  waited  long 
enough,  equilibrium  would  be  attained,  but  it  might  take  thousands  of 
years.  But  the  probability  of  a  reacting  collision  increases  enormously 
with  the  temperature,  which  is  the  reason  why  a  spark,  a  loeali/ed  region 
of  exceedingly  high  temperature,  can  start  the  reaction.  Once  it  is 
started,  the  heat  liberated  by  the  reaction  near  the  spark  raises  the  gas  in 
the  neighborhood  to  such  a  high  temperature  that  it  in  turn  can  react, 
liberating  more  heat  and  allowing  gas  still  further  away  to  react,  and  so 
on.  In  this  way  a  sort  of  wave  or  front  of  reaction  is  propagated  through 
the  gas,  with  a  very  high  velocity,  and  this  is  characteristic  of  explosion 
reactions. 

It  is  true  in  general  that,  given  a  collision  of  the  suitable  molecules 
for  a  reaction,  the  probability  of  reaction  increases  enormously  rapidly 
with  the  temperature.  When  measurements  of  rates  of  reaction  are 
made,  it  is  found  that  the  probability  of  reaction  can  be  expressed  approxi- 
mately by  the  formula 

Qi 

Probability  of  reaction  =  const.  X  c    k'*\  (3.1) 

where  Qi  is  a  constant  of  the  dimensions  of  an  energy.  Equation  (3.1) 
suggests  the  following  interpretation:  suppose  that  out  of  all  the  collisions, 
only  those  in  which  the  colliding  molecules  taken  together  have  an 
energy  (translational,  rotational,  and  vibrational)  of  Qi  or  more  can 
produce  a  reaction.  By  the  Maxwcll-Boltzmann  distribution,  the  frac- 

-5i 
tion  of  molecules  having  this  energy  will  contain  the  factor  e    kT.     (The 

fraction  having  an  energy  greater  than  Qi  can  be  shown  to  contain  this 
factor,  as  well  as  the  fraction  having  an  energy  between  Q\  and  Qi  +  dQi, 
which  is  what  we  usually  consider.)  Thus  we  can  understand  the  varia- 
tion of  rate  of  reaction  with  temperature.  We  must  next  ask,  why  do  the* 
molecules  need  the  extra  energy  Qi,  in  order  to  react?  This  energy  is 
ordinarily  called  the  energy  of  activation,  and  we  say  that  only  the 
activated  molecules,  those  which  have  an  energy  at  least  of  this  amount, 
can  react. 

To  understand  why  an  energy  of  activation  is  required  for  a  reaction, 
we  may  think  about  a  hypothetical  mechanism  for  the  reaction.  In  our 


160  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  X 

particular  case  of  hydrogen  and  oxygen  combining  to  form  water,  we 
imagine  a  collision  in  which  two  hydrogen  molecules  hit  an  oxygen 
molecule  (this  will  be  the  way  the  collision  will  appear,  for  on  account  of 
their  light  mass  the  hydrogens  will  be  traveling  much  faster  than  the 
oxygen  in  thermal  equilibrium).  During  the  collision,  the  atoms  rear- 
range themselves  to  form  two  water  molecules,  which  then  fly  apart  with 
very  groat  energy  (on  account  of  the  heat  of  reaction).  Now,  obviously, 
those  particular  collisions  will  be  favored  for  reaction  in  which  the  atoms 
need  the  minimum  rearrangement  in  the  reaction,  and  a  little  reflection 
shows  that  the  most  favorable  configuration  is  that  shown  in  Fig.  X-l 
(a),  in  which  the  velocities  of  the  various  atoms  are  shown  by  arrows.  As 
we  follow  the  successive  sketches  of  Fig.  X-l  showing  the  progress  of  the 
collision,  we  see  that  the  hydrogens  approach  the  oxygens,  attaining  in  (c) 
a  shape  very  much  like  two  water  molecules.  In  the  first  part  of  the 
collision,  (a)  arid  (6),  the  hydrogens  have  most  of  the  kinetic  energy.  For 
a  favorable  reaction,  however,  the  relations  between  the  velocities  of 
hydrogens  and  oxygons  on  collision  must  be  such  that  the  hydrogen  gives 
up  most  of  its  kinotic  energy  to  the  oxygon.  The  condition  for  this 
can  bo  found  from  elementary  considerations  of  conservation  of  momen- 
tum and  kinetic  energy  on  collision,  and  demands  that  the  oxygen  atoms 
be  moving  in  the  same  direction  as  the  hydrogens  on  collision  but  with 
considerably  smaller  velocity.  That  is,  the  oxygen  molecule  must  have 
had  considerable  vibnition.il  kinetic  energy  and  the  correct  phase  of 
vibration,  while  the  hydrogens  must  have  had  large  translational  kinetic 
energy.  Now  in  the  second  part  of  the  collision,  (W),  (c),  (/),  and  (g),  the 
oxygens  have  most  of  the  kinetic  energy.  They  fly  apart,  carrying  the 
hydrogens  with  them,  and  form  the  atoms  into  two  water  molecules.  The 
hydrogens  end  up  bound  to  the  oxygens,  but  with  some  vibrational 
kinetic  energy  in  the  mode  of  vibration  indicated  by  (g). 

We  can  now  follow  the  energy  relations  in  the  reaction  by  drawing  a 
suitable  potential  energy  curve.  The  potential  energy  of  the  whole 
system,  of  course,  depends  on  the  positions  of  all  the  atoms  and  would 
have  to  be  plotted  as  a  function  of  many  coordinates.  We  can  simplify, 
however,  by  considering  it  as  a  function  only  of  the  distance  r  between 
the  oxygen  atoms.  For  each  value  of  r,  there  will  be  a  particular  position 
for  the  hydrogen  atoms  that  will  correspond  to  a  minimum  of  energy. 
Thus  in  (a),  where  the  oxygen  atoms  are  forming  an  oxygen  molecule,  the 
hydrogen  molecules  and  the  oxygen  molecule  will  attract  each  other 
slightly,  provided  the  oxygen-hydrogen  distance  is  considerable,  but  will 
repel  provided  they  come  too  close  together,  as  we  shall  learn  later  when 
we  consider  intcrmolecular  forces  in  imperfect  gases.  There  will  be  a 
position  of  equilibrium,  with  the  hydrogens  a  considerable  distance — three 
or  four  angstroms — away  from  the  oxygen,  and  with  an  energy  of  perhaps 


SEC.  3] 


CHEMICAL  EQUILIBRIUM  IN  GASES 


161 


162 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  X 


a  fraction  of  a  kilogram  calorie  lower  than  the  energy  at  infinite  separation 
of  hydrogen  from  oxygen.  Similarly  in  (g)  the  atoms  are  formed  into  two 
water  molecules,  and  the  minimum  of  energy  of  the  hydrogens  comes 
when  they  are  at  the  distances  and  angles  with  respect  to  the  oxygen 
which  we  find  in  a  water  molecule.  We  now  show,  in  Fig.  X-2,  a  sketch 
of  the  potential  energy  of  the  whole  system,  when  the  oxygens  are  at 
distance  r,  and  the  hydrogens  are  in  their  positions  of  minimum  energy 
for  each  value  of  r. 

First,  we  ask  how  Fig.  X-2  was  constructed.  When  the  oxygens  are 
close  together  forming  an  oxygon  moloculo,  the  energy  of  the  hydrogens, 
being  only  interinolecular  attraction,  is  small  and  the  curve  is  practically 


10  20 

r  (Angstroms) 

Fr<;    X-2. — Potential  onorgy  of  4H  +  2O,  as  function  of  <  M)  distanro  r 

the  interatomic  energy  for  the  oxygen  molecule.  This  is  the  curve  (a) 
of  Fig.  X-2,  going  to  an  energy  at  infinite  separation  which  is  greater  by 
Z)(  =  117  kg.-cal.)  than  at  the  minimum,  which  comes  at  1.20  A.  On 
the  other  hand,  when  the  oxygens  are  far  apart  they  form  two  water 
molecules.  At  infinite  separation  of  these  two  molecules,  the  energy  of 
the  whole  system  is  less  by  117  kg.-cal.  (=  A//0,  which  only  happens  to 
be  equal  to  the  D  of  the  oxygen  molecule  by  a  coincidence),  than  when 
two  hydrogen  and  one  oxygen  molecule  are  formed.  The  curve  (6) 
shows  the  interaction  between  these  water  molecules.  Starting  with  the 
asymptotic  energy  just  mentioned,  the  curve  rises  with  decreasing  dis- 
tance, because  the  two  water  molecules,  set  with  their  negative  oxygen 
ions  facing  each  other,  repel  each  other  on  account  of  electrostatic  repul- 
sion of  like  charges.  As  the  distance  decreases,  to  something  of  the  order 
of  three  angstroms,  the  molecules  begin  to  hit  each  other,  causing  the 
curve  (6)  to  rise  steeply.  Curves  (a)  and  (6)  both  form  limiting  cases. 
For  small  r's,  curve  (a)  must  be  correct,  and  for  larger  r's  curve  (6).  The 


SEC.  3]  CHEMICAL  EQUILIBRIUM  IN  GASES  163 

full  line  in  Fig.  X-2  represents  a  sketch  of  the  way  the  actual  curve  may 
look,  reducing  to  these  two  limiting  curves.  It  will  be  noted  that  at 
intermediate  distances  the  actual  curve  lies  below  either  curve  (a)  or 
(b).  Essentially,  the  reason  is  as  follows:  When  we  have  quite  separated 
molecules,  as  the  two  water  molecules  at  large  distances,  each  atom  of  one 
molecule  repels  each  atom  of  the  other.  But  as  they  approach,  as  in 
configuration  (c)  of  Fig.  X-l,  there  is  a  little  uncertainty  as  to  whether 
they  form  two  water  molecules,  or  two  hydrogens  and  an  oxygen.  As  a 
consequence,  the  oxygen  atoms  make  a  compromise  between  repelling 
each  other,  as  they  would  in  two  water  molecules,  and  attracting,  as  they 
would  in  an  oxygen  molecule.  That  is,  the  repulsion  which  causes  the 
rise  in  curve  (6),  Fig.  X-2,  is  diminished  and  the  actual  curve  does  not 
continue  to  rise  as  curve  (6)  does. 

Now  that  we  have  the  curve  of  Fig.  X-2,  we  can  apply  it  to  the  reac- 
tion as  shown  in  Fig.  X-l.  The  first  part  of  the  reaction,  diagrams  (a), 
(6),  and  (c)  of  Fig.  X-l,  cannot  be  represented  directly  on  Fig.  X-2,  for 
in  it  the  hydrogen  molecules  have  a  great  deal  of  kinetic  energy  and  aro 
by  no  means  in  the  position  of  minimum  potential  energy.  But  by  (r) 
of  Fig.  X-l  the  hydrogen  atoms  have  given  up  most  of  their  kinetic 
energy  to  the  oxygens,  and  during  the  rest  of  the  process  the  curve  of 
Fig.  X-2  applies  fairly  accurately.  As  far  as  the,  first  part  of  the  process 
is  concerned,  we  can  interpret  it  as  a  process  in  which  the  oxygens  had 
their  vibrational  energy  increased  from  such  a  value  as  E}  in  Fig.  X-2  to 
Et,  symbolized  by  the  arrow  in  the  figure.  When  they  had  the  energy  K\ 
they  simply  vibrated  back  and  forth  for  a  short  range  about  the  distance 
re  of  minimum  energy.  But  with  the  energy  E2  the  motion  changes 
entirely:  the  oxygens  fly  apart,  carrying  the  hydrogens  with  them  to  form 
water  molecules  and  ending  up  with  infinite  separation  of  the  molecules 
and  a  very  high  kinetic  energy.  And  now  we  see  the  need  of  the  energy 
of  activation.  From  Fig.  X-2  wo  see  that  there  is  a  maximum  of  potential 
energy  between  the  minimum  at  i\  and  the  still  lower  value  at  infinite 
separation.  For  the  water  molecules  to  separate,  the  energy  E%  must 
lie  higher  than  this  maximum.  But  this  energy  is  supplied,  as  we  have 
seen,  by  the  combined  energy  of  all  the  colliding  molecules  before  collision. 
We  thus  sec  that  the  energy  of  activation  Qi  is  to  be  interpreted  as  the 
height  of  this  maximum  above  the  minimum  at  r€. 

A  minimum  of  potential  energy  such  as  that  at  re  in  Fig.  X-2,  separated 
by  a  maximum  from  a  still  lower  region,  is  often  met  in  atomic  and 
molecular  problems  and  is  called  a  position  of  metastablc  equilibrium,  or 
a  metastable  state.  It  is  stable  as  far  as  small  displacements  are  con- 
cerned, but  a  large  displacement  can  push  the  system  over  the  maximum, 
after  which  it  does  not  return  to  the  original  position  but  to  the  entirely 
different  configuration  of  really  lowest  potential  energy.  In  all  such 


164  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  X 

cases,  the  rate  of  transition  from  the  metastable  to  the  really  stable  con- 
figuration, at  temperature  T,  depends  on  a  factor  exp  (—  Qi/kT),  where 
Qi  is  the  energy  of  activation,  or  height  of  the  maximum  above  the  mini- 
mum, for  in  all  such  cases  it  is  only  the  molecules  with  energy  greater  than 
Qi  that  can  react.  Let  us  see  how  rapidly  such  a  factor  can  depend  on 
temperature.  From  Fig.  X-2  it  seems  reasonable  that  in  that  case  Q\ 
could  bo  of  the  order  of  magnitude  of  40  kg.-cal.  Then  the  factor 
exp  (-QiAT)  will  become  exp  (-40,000/tf  5T)  =  exp  (-  40,000/1.  98  T)  = 
exp  (-20,000/T)  approximately.  For  T  =  300°  abs.,  this  factor  is 
exp  (-66.7)  =  1Q-*29  approximately,  while  for  T  =  3000°  abs.  it  is 
exp  (  —  6.67)  =  10~3  approximately.  Thus  an  increase  of  temperature 
from  room  temperature  at  3000°  abs.,  which  could  easily  be  attained  in  a 
spark,  might  make  a  difference  of  1C26  in  the*  rate  of  reaction.  A  process 
that  would  take  10~16  sec.  at  the  high  temperature  might  take  1010  sec.,  or 
3  X  108  yrs.,  at  the  low  temperature,  and  would  for  all  practical  purposes 
never  happen  at  all.  This  is  an  extreme  but  by  no  means  an  unreasonable 
example. 

We  are  now  in  position  to  see  why,  though  the  energy  of  activation 
enters  into  the  rate  of  reaction  in  such  an  important  way,  it  does  not  affect 
the  final  equilibrium.  The  factor  C(T\  in  Eq.  (1.2),  determining  the 
rate  of  combination  of  hydrogen  and  oxygen  molecules  to  form  water,  will 
contain  a  factor  exp  (  —  Qi/kT),  as  we  have  seen.  But  a  glance  at  Fig. 
X-2  shows  that  in  the  reverse  reaction,  in  which  two  water  molecules 
combine  to  form  hydrogen  and  oxygen,  the  water  molecules  must  have 
an  energy  at  least  equal  to  Q\  +  A//0,  so  that  they  can  climb  over  the 
maximum  of  potential  energy  and  approach  closely  enough  to  form  an  oxy- 
gen molecule.  Thus  the  probability  of  the  reverse  collision,  given  by  Eq. 
(1.3),  contains  the  f  actor  C'(7T)  with  the  exponential  exp  [-«?i+A//0)/*7T]. 
Finally  in  the  coefficient  KP(T)j  given  by  Eq.  (1.4),  we  must  have 

C'  (T} 

with   the  factor  exp   {[-(Qi  +  A//0)  +  Q\]/kT\,  which  equals 


exp  [  —  (£Ho/kT)],  in  agreement  with  Eqs.  (2.6)  and  (2.9),  the  energy  of 
activation  canceling  out.  Of  course,  in  this  simple  argument  we  have 
neglected  such  things  as  specific  heats,  so  that  we  have  not  reproduced 
the  whole  form  of  Eq.  (2.6)  from  a  kinetic  point  of  view,  but  this  could 
be  done  if  sufficient  care  were  taken. 

There  is  one  point  about  a  reaction  like  the  combination  of  two  water 
molecules  to  form  oxygen  and  hydrogen,  which  we  have  just  mentioned, 
that  is  worth  discussion.  From  Fig.  X-2,  if  the  molecules  approach  with 
energy  E%  sufficient  to  pass  over  the  maximum  of  potential,  they  will  not 
be  trapped  to  form  oxygen  and  hydrogen  molecules  unless  the  energy  of 
the  oxygens  drops  from  J?2  to  some  value  like  EI  during  the  collisions. 
This  of  course  can  happen  by  giving  the  excess  energy  to  the  hydrogen 


SEC.  3] 


CHEMICAL  EQUILIBRIUM  IN  (JASES 


165 


atoms,  sending  them  shooting  off  as  hydrogen  molecules.  But  there  are 
sometimes  other  reactions  in  which  this  cannot  happen.  For  instance, 
consider  the  simple  recombination  reaction  of  two  oxygen  atoms  to  form 
an  oxygen  molecule,  shown  in  Fig.  X-3.  Here  if  the  atoms  approach  with 


FIG.  X-.3. — Kerombmation  of  atoms  to  form  a  molecule. 


energy  E%,  there  is  nothing  within  the  system  itself  able  to  absorb  the 
necessary  energy  to  make  them  fall  down  to  the  energy  E\,  and  be  bound 
to  form  a  molecule.  Such  a  recombination  of  two  atoms  can  only  occur 
if  they  happen  to  be  in  collision  with  a  third  body,  atom  or  molecule,  at 
the  same  time,  which  can  absorb  the  excess  energy  and  leave  the  scene  of 
collision  with  high  velocity. 


CHAPTER  XI 
THE  EQUILIBRIUM  OF  SOLIDS,  LIQUIDS,  AND  GASES 

We  have  ho  far  studied  only  perfect  gases  and  have  not  taken  up 
imperfect  gases,  liquids,  and  solids.  Before  we  treat  them,  it  is  really 
necessary  to  understand  what  happens  when  two  or  more  phases  are  in 
equilibrium  with  each  other,  and  the  familiar  phenomena  of  melting,  boil- 
ing, and  the  critical  point  and  the  continuity  of  the  liquid  and  gaseous 
states.  We  shall  now  proceed  to  find  the  thermodynamic  condition  for 
the  coexistence  of  two  phases  and  shall  apply  it  to  a  general  discussion' 
of  the  forms  of  the  various  thermodynamic  functions  for  matter  in  all 
three  states. 

1.  The  Coexistence  of  Phases. — It  is  a  matter  of  common  knowledge 
that  at  the  melting  point,  a  solid  and  a  liquid  can  exist  in  equilibrium  with 
each  other  in  any  proportions,  as  can  a  liquid  and  vapor  at  the  boiling 
point.  There  is  no  tendency  for  the  relative  proportions  of  the  two 
phases,  as  they  are  called,  to  change  with  time.  On  the  other  hand,  if 
we  are  not  at  the  melting  or  boiling  point,  there  is  no  such  equilibrium. 
At  100°C.,  for  instance,  water  vapor  above  atmospheric  pressure  will 
immediately  start  to  condense,  enough  liquid  forming  so  that  the  remain- 
ing vapor  and  the  liquid  will  come  to  atmospheric  pressure;  while  if  water 
at  this  temperature  is  below  atmospheric  pressure,  enough  liquid  will 
evaporate  or  boil  away  to  raise  the  pressure  to  one  atmosphere,  so  that 
only  at  atmospheric  pressure  can  the  two  coexist  at  100°C.  in  arbitrary 
proportions.  For  each  temperature  the  equilibrium  takes  place  at  a 
definite  pressure;  that  is,  we  can  give  a  curve,  called  the  vapor  pressure 
curve  or  in  general  the  equilibrium  curve,  in  the  P-T  plane,  along  which 
equilibrium  occurs.  This  curve  separates  those  parts  of  the  P-T  plane 
where  just  one,  or  just  the  other,  of  the  phases  can  exist.  Thus  in  gen- 
eral, where  a  number  of  phases  occur  in  different  regions  of  the  P-T  plane, 
equilibrium  lines  separate  the  regions  where  each  phase  occurs  separately. 
Along  a  line  two  phases  exist;  where  three  lines  join  at  a  point,  three 
phases  can  exist,  and  such  a  point  is  called  a  triple  point. 

The  resulting  diagram  is  called  a  phase  diagram.  In  the  figures 
below,  such  a  diagram  is  drawn  for  water,  a  familiar  and  in  some  ways  a 
remarkable  example.  Figure  XI-1  shows  the  diagram  for  a  scale  of 
pressures  on  which  the  critical  point  is  represented  by  a  reasonable  value. 
The  ordinary  melting  and  boiling  points,  at  1  atm.  and  at  0°C.,  and 
100°C.,  respectively,  are  easily  found.  We  see  that  the  boiling  point 

166 


SBC.  1] 


THE  EQUILIBRIUM  OF  SOLIDS 


167 


rises  rapidly  to  higher  temperatures  as  the  pressure  is  raised,  until  finally 
the  critical  point  is  reached,  above  which  there  is  no  longer  discontinuity 
between  the  phases.  The  melting  point,  on  the  other  hand,  is  almost 
independent  of  pressure,  decreasing  as  a  matter  of  fact  very  slightly  with 
increasing  pressure. 

Figure  XI-2  gives  a  different  pressure  scale,  on  which  small  fractions 
of  an  atmosphere  can  be  noted.  The  triple  point  is  immediately  observed, 
corresponding  to  a  low  pressure  and  a  temperature  almost  at  0°C.,  at 
which  ice,  water,  and  water  vapor  can  exist  at  the  same  time,  so  that  if  a 
dish  of  water  is  cooled  to  this  temperature  in  a  suitable  vacuum,  a  coating 
of  ice  will  form  and  steam  will  bubble  up  from  below  the  ice.  Below  this 
temperature,  liquid  water  does  not  occur,  but  as  we  can  sec  an  equilibrium 
is  possible  between  solid  and  gas.  If  the  solid  is  reduced  below  the 


£200 

JS 

a 

8 
Jioo 


Ice 


Water 


rpor 


/ce 


Wafer! 


Vapor 


100    200    300  400 
Deg.C 

FIG.  XI-1. — Phase  diagram  of 
water.  Critical  point  *  PC  =  218 
atm.,  Tc  =  374°O 


-20  0  20 

Deg.C 

FIG.  XI-2. — Low-pressure 
phase  diagram  of  water.  Triple 
point:  P  =  4..5S  mm.,  T  = 
().0075°0. 


pressure  corresponding  to  equilibrium,  it  will  evaporate  directly  into 
water  vapor.  This  is  the  way  snow  and  ice  disappear  in  weather  below 
freezing;  and  it  is  a  familiar  fact  that  solid  carbon  dioxide,  whose  triple 
point  lies  at  a  pressure  greater  than  atmospheric;,  evaporates  by  this 
method  without  passing  through  the  liquid  phase. 

In  Fig.  XI-3,  the  pressure  scale  is  changed  in  the  other  direction,  so 
that  we  show  up  to  12,000  atm.  Here  the  gaseous  phase,  which  exists 
for  pressures  only  up  to  a  few  hundred  atmospheres,  cannot  be  shown  on 
account  of  the  scale.  On  the  other  hand,  a  great  deal  of  detail  has 
appeared  in  the  region  of  the  solid.  It  appears  that,  in  addition  to  the 
familiar  form  of  ice,  there  are  at  least  five  other  forms  (the  fifth  exists  at 
higher  pressures  than  those  shown  in  the  figure).  These  forms,  called 
polymorphic  forms,  presumably  differ  in  crystal  structure  and  in  all 
their  physical  properties,  as  density,  specific  heat,  etc.  The  regions 
where  these  phases  exist  separately  are  divided  by  equilibrium  lines,  on 


168 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  XI 


10,000 


-40        -20 


40 


0 
Deg.C. 

FIG.  XI-3. — HiKh-pressuro  phase  diagram  of  water. 


200- 


0) 

a 
</> 
o 

E 


100- 


10  20 

cc/gm 

FIG.  XI-4. — P-V-T  surface  for  water. 


SEC.  2] 


THE  EQUILIBRIUM  OF  SOLIDS 


169 


which  two  of  them  can  coexist  in  equilibrium,  and  a  number  of  triple 
points  are  shown.  Transitions  from  one  phase  to  another  along  an 
equilibrium  line  are  called  polymorphic  transitions.  There  has  never 
been  found  any  suggestion  of  a  critical  point,  or  termination  of  an  equilib- 
rium line  with  gradual  coalescence  of  the  two  phases  in  properties,  for 
any  equilibrium  between  liquid  and  solid  or  between  two  solid  phases. 
Critical  points  appear  to  exist  only  in  the  liquid-vapor  transition. 

2.  The  Equation  of  State. — The  three  figures  that  we  have  drawn 
give  only  part  of  the  information  about  the  phase  equilibrium;  for  greater 


10,000- 


s 

<b 
a 

CO 

o 

i 


5,000- 


'/ce  anef  vapor 
Icc/gm 

Fi«.  XI-5. — P-V-T  Hiirftu'o  for  water,  high  pressure. 

completeness,  we  should  show  the  whole  equation  of  state1,  the  relation 
between  pressure,  temperature,  and  volume.  This  is  done  in  Figs.  XI-4 
and  XI-5,  where  P-V-T  surfaces  arc  shown  in  perspective,  for  the  case  of 
the  liquid-vapor  equilibrium  and  for  the  polymorphic  forms  in  equilibrium 
with  the  liquid.  A  number  of  isothermals,  lines  of  constant  temperature, 
are  drawn  on  each  surface  to  make  them  easier  to  interpret.  Some  simple 
facts  arc  immediately  obvious  from  these  surfaces.  For  instance,  water 
is  exceptional  in  that  the  solid,  ice,  has  a  larger  volume  than  the  liquid. 
As  we  see,  this  is  true  of  ice  I,  the  phase  existing  at  low  pressure,  but  it  is 
not  true  of  the  other  phases  II,  III,  V,  VI,  all  of  which  have  smaller 
volumes  than  water  at  the  corresponding  pressure.  Furthermore,  we  see 
that  though  water  seems  quite  incompressible  as  far  as  the  low  pressure 


170  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XI 

surface  in  the  first  figure  is  concerned,  the  second  figure  shows  that  a 
pressure  of  12,000  atm.  produces  a  diminution  of  volume  of  about  20  per 
cent.  Again,  the  melting  point  of  ice  I  is  hardly  affected  by  the  pressures 
indicated  in  the  low  pressure  surface,  but  the  other  surface  shows  that  a 
pressure  of  about  2000  atm.  lowers  the  melting  point  by  more  than  20°C. 
One  interesting  fact  to  notice  is  that  a  vertical  lino  cutting  either  of  tho 
surfaces  will  cut  it  in  just  one  point;  that  is,  for  a  given  volume  and  tem- 
perature, the  pressure  is  uniquely  determined.  We  shall  see  shortly 
that  this  can  be  shown  to  be  true  quite  generally. 

As  we  see,  the  P-V-T  surfaces  are  divided  into  a  number  of  different 
regions  with  sharp  edges  separating  them.  In  some  of  these  regions  one 
phase  alone  can  exist,  while  in  others  two  phases  can  coexist.  The  regions 
of  the  second  type,  when  projected  onto  the  P-T  surface,  become  the 
equilibrium  lines  that  we  have  previously  mentioned;  thus  they  are  ruled 
surfaces,  the  rulings  being  parallel  to  the  volume  axis.  At  a  given  pres- 
sure and  temperature  on  an  equilibrium  line,  in  other  words,  tho  volume 
ran  have  any  value  between  two  limiting  values,  tho  volumes  of  tho  two 
phases  in  question.  The  meaning  of  this  is  that  there  is  a  mixture  of  the 
two  phases,  so  that  the  volume  of  the  mixture  depends  on  tin;  relative 
concentrations  of  the  two  arid  is  not  really  a  property  of  either  phase,  but 
is  a  measure  of  the  relative  concentrations. 

3.  Entropy  and  Gibbs  Free  Energy.  Tho  equation  of  state  does  not 
alone  determine  the  thermodynamic  behavior  of  a  substance;  wo  must 
also  know  its  specific  heat,  or  its  entropy  or  Gibbs  free  energy.  Wo  shall 
first  give  the  entropy  as  a  function  of  pressure  and  temperature.  This 
can  of  course  be  determined  entirely  by  experiment.  We  start  with  the 
solid  at  the  absolute  zero.  There,  according  to  tho  quantum  theory,  as 
we  have  seen  in  Chap.  Ill,  the  entropy  is  zero.  The  entropy  of  tho  solid 
at  a  higher  temperature  can  bo  found  from  tho  specific  heat,  for  at  con- 
stant pressure  we  have 


Since,  according  to  the  quantum  theory,  the  specific  heat  goes  to  zero  at 
the  absolute  zero,  tho  integral  in  Eq.  (3.1)  behaves  properly  at  the  abso- 
lute zero.  By  means  of  Eq.  (3.1),  we  find  the  entropy  of  the  solid  at  any 
temperature1,  at  a  given  pressure;  since  tho  specific  heat  depends  only 
slightly  on  pressure,  this  means  practically  that  the  entropy  of  the  solid 
is  a  function  only  of  temperature,  not  of  pressure,  on  the  scales  used  in 
Figs.  XI-1  and  XI-2,  though  not  in  Fig.  XI-3.  Next,  we  wish  the 
entropy  of  the  liquid  and  vapor.  If  the  pressure  is  below  that  at  the 
triple  point,  a  horizontal  line,  or  line  of  constant  pressure,  in  the  phase 
diagram  will  carry  us  from  the  region  of  solid  into  that  of  vapor.  There 


SEC.  3]  THE  EQUILIBRIUM  OF  SOLIDS  171 

is  a  discontinuous  change  of  entropy  as  we  cross  the  line,  equal  to  the  heat 
absorbed  (the  latent  heat  of  vaporization)  divided  by  the  temperature 
(the  temperature  of  sublimation  for  the  pressure  in  question).  This 
change  of  entropy,  which  we  may  call  the  entropy  of  vaporization  and 
denote  by  A/SV,  is 

AS.  -  Y:  (3-2) 

where  Lv,  Tv  are  the  latent  heat  and  temperature  of  vaporization  at  the 
given  pressure.  Adding  this  change  of  entropy  to  the  entropy  of  the  solid 
(3.1)  just  below  the  sublimation  point,  we  have  the  entropy  of  the  vapor 
just  above  this  point.  Then,  applying  Eq.  (3.1)  to  the  vapor  rather  than 
the  solid,  we  can  follow  to  higher  temperatures  at  constant  pressure  and 
find  the  entropy  of  the  gas,  as 


CTvn  i          CT 

=         --'-dT  +  J'°  + 

JO       ^  1  v          J  T, 


(3.3) 


In  Eq.  (3.3),  the  first  term  represents  the  entropy  of  the  solid  at  the 
sublimation  point,  the  second  the  increase  of  entropy  on  vaporizing,  and 
the  third  the  further  increase  of  entropy  from  the  sublimation  point  up  to 
the  desired  temperatures. 

If  the  pressure  is  above  the  triple  point,  the  solid  will  first  melt,  then 
vaporize.  In  this  case,  we  can  proceed  in  a  similar  way.  On  melting, 
the  entropy  increases  by  the  entropy  of  fusion,  determined  from  the  latent 
heat  of  fusion  and  temperature  of  fusion  by  the  relation 

AS/  =  '-£,  (3.4) 

analogous  to  Eq.  (3.2).  Then  the  entropy  of  the  liquid  at  any  tempera- 
ture is 


(*Tic  T  l*Tr 

Si-         <£ur+^  +        L*'dT. 

JO        *  If          JT,    1 


(3.5) 


As  the  temperature  rises  further,  the  liquid  will  vaporize  and  the  entropy 
will  increase  by  the  entropy  of  vaporization.  The  gas  above  this  tem- 
perature will  then  have  the  entropy 

8.  =    r^dT  +  %  +    F'CfdT  +  (7  +   f^f'dT.        (3.6) 
Jo     :/  //       jTf    I  I*       JT,  / 

It  is  interesting  to  note  that  a  relation  between  the  latent  heat  of  vapori- 
zation of  the  solid,  the  heat  of  fusion,  and  the  heat  of  vaporization  of  the 
liquid,  at  the  triple  point,  arises  from  the  fact  that  Eqs.  (3.3)  and  (3.6) 
must  give  identical  values  for  the  entropy  of  the  gas  at  the  triple  point. 


172 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  XI 


Since  in  this  case  Tf  =  Tv  =  T,  the  integrals  involving  the  specific  heats 
of  the  liquid  and  gas  drop  out,  and  wo  have  at  once  the  relation 

Lv  of  solid  =  Lf  +  Lv  of  liquid,  at  the  triple  point.  (3.7) 

In  Fig.  XI-6  we  show  the  entropy  of  water  in  its  three  phases,  as  a 
function  of  pressure  and  temperature,  computed  as  we  have  described 
above.  We  are  struck  by  the  resemblance  of  this  figure  to  that  giving 
the  volume,  Fig.  XI-4;  the  entropy,  like  the  volume,  increases  with 
increase  of  temperature  or  decrease*  of  pressure.  Lines  of  constant  pres- 


10     20    30    40     50 
Ceil  per  Mole  Degree 

FIG.  XI-0.    -Kntropy  of  Wiiter  iia  function  ot  pressure  und  temperature. 

sure  are  drawn  in  Fig.  XI-6.  The  regions  of  coexistence  of  phases  are 
shown  in  Fig.  XI-6  as  in  Fig.  XI-4,  and  the  latent  heat  is  given  by  the 
length  of  the  horizontal  line  lying  in  the  region  of  coexistence,  multiplied 
by  the  temperature.  Graphs  of  the  form  of  Fig.  XI-6  (generally  pro- 
jected onto  the  T-S  plane)  are  of  considerable1  practical  importance  in 
problems  involving  thcrmodynamic  cycles,  as  heat  engines  and  refrigera- 
tors, on  account  of  the  fact  that  the  isothermals  are  represented  by  lines 
of  constant  T  and  adiabatics  by  lines  of  constant  S,  so  that  the  diagram 
of  a  Carnot  cycle  in  such  a  plot  is  simply  a  rectangle.  Furthermore,  the 
area  of  a  closed  curve  representing  a  cycle  in  the  T-S  diagram  gives 
directly  the  work  done  in  the  cycle,  just  as  it  does  in  the  P-V  diagram. 


SEC.  3]  THE  EQUILIBRIUM  OF  SOLIDS  173 

This  is  seen  at  once  from  the  first  and  second  laws  in  the  form 

Tdft  =  dU  +  PdV. 

Integrating  around  a  closed  cycle,  we  must  have  u)  dV  —  0,  since  (r  is  a 
function  of  the  state  of  the  system.     Hence 

<j>T  dS  =  fP  dV,  (3.8) 

where  CD  indicates  an  integral  taken  about  a  complete  cycle,  and  since 
(foP  dV  equals  the  work  done,  u)T  dS  must  equal  it  also. 


I     23456789   10   II 

-G  (Kg. coil/mole) 
FIG.  XI-7. — Gihbs  froo  energy  of  water,  as  function  of  pressure  and  temperature 

The  Gibhs  free  energy  G  as  a  fund  ion  of  pressure  and  temperature  is 
sketched  in  Fig.  XI-7.  It  can  also  be  found  directly  from  experiment. 
At  constant  pressure,  we  have  ft  =  —  (d(?/d77)p,  or  G  =  —  J/SrfT7,  and 
since  V  =  (dG/dP)r,  we  have  G  =  J  V  dP  at  constant  temperature,  from 
a  combination  of  which  the  Gibbs  free  energy  can  be  found  from  equation 
of  state  and  specific  heat.  The  surface  of  Fig.  XI-7  looks  quite  different 
from  those  for  volume  and  entropy;  for  while  the  volume  and  entropy 
change  discontinuously  with  a  change  of  phase,  resulting  in  the  ruled 
surfaces  indicating  coexistence  of  phases  wrhich  are  so  characteristic  of 
Figs.  XI-4,  XI-5,  and  XI-6,  the  Gibbs  free  energy  must  be  equal  for  the 
two  phases  in  equilibrium.  This  has  already  been  discussed  in  Sec.  3, 
Chap.  II  and  follows  from  the  fundamental  property  of  the  Gibbs  free 
energy,  that  its  value  must  be  a  minimum  for  equilibrium  at  given  pres- 


174  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XI 

sure  and  temperature.  Thus,  if  there  is  equilibrium  between  two  phases 
at  a  given  pressure  and  temperature,  a  transfer  of  some  material  from 
one  phase  to  the  other  cannot  change  the  Gibbs  free  energy,  so  that  the 
value  of  G  must  be  the  same  for  both  phases.  As  we  observe  from  Fig. 
XI-7,  each  phase  has  a  different  surface  for  G  as  a  function  of  P  and  T, 
and  the  intersection  of  two  of  these  surfaces  gives  the  condition  for 
equilibrium.  It  is  interesting  to  notice  the  behavior  of  the  surface  near 
the  critical  point:  the  lines  of  constant  pressure,  which  are  drawn  on  the 
surface,  have  discontinuities  of  slope  below  the  critical  point  but  merely 
continuous  changes  of  slope  above  this  point.  We  shall  see  in  a  later 
chapter  how  such  lines  can  come  about  mathematically. 

4.  The  Latent  Heats  and  Clapeyron's  Equation. — There  is  a  very 
important  thermodynamie  relation  concerning  the  equilibrium  between 
phases,  called  Clapeyron's  equation,  or  sometimes  the  (1ln,peyron-(Mausius 
equation.  By  way  of  illustration,  let  us  consider  the  vaporization  of 
water  at  constant  temperature  and  pressure.  On  our  P-V-T  surface,  the 
process  we  consider  is  that  in  which  the  system  is  carried  along  an  iso- 
thermal on  the  ruled  part  of  the  surface,  from  the  .state  \\here  it  is  all 
liquid,  with  volume  Vi,  to  the  state  \\here  it  is  all  gas,  with  volume  Va. 
As  we  go  along  this  path,  we  wish  to  find  ihe  amount  of  heat  absorbed. 
We  can  find  this  from  one  of  Maxwell's  relations,  Kq.  (4.12),  Chap.  II: 


dVjr  ~  \dT/  (4'1} 

The  path  is  one  of  constant  temperature,  so  that  if  we  multiply  by  T 
this  relation  gives  the  amount  of  heat  absorbed  per  unit  increase  of 
.volume.  >But  on  account  of  the  nature  of  the  surface,  (dP/dT)v  is  the 
same  for  any  point  corresponding  to  the  same  temperature,  no  matter 
what  the  volume  is;  it  is  simply  the  slope  of  the  equilibrium  curve  on  the 
P-T  diagram,  which  is  often  denoted  simply  by  rtP/dT  (since  in  the  P-T 
diagram  there  is  only  one  independent  variable,  and  we  do  not  need 
partial  derivatives).  Then  we  can  integrate  and  have  the  latent  heat  L 
given  by 


or 

L  =  T^(Va  -  Vi),  '          (4.2) 

which  is  Clapeyron's  equation.    It  is  often  written 

dP  L 


_ 
dT  ~  T(Vt  -  Vi)' 


SBC.  4]  THE  EQUILIBRIUM  OF  SOLIDJS  175 

or 

W  -  T(V.  -  7Q.  .     . 

dP  -  I  (4<3) 

Clapeyron's  equation  holds,  as  we  can  see  from  its  method  of  derivation . 
for  any  equilibrium  between  phases.  In  the  general  case,  the  difference 
of  volumes  on  the  right  side  of  the  equation  is  the  volume  after  absorbing 
the  latent  heat  L,  minus  the  volume  before  absorbing  it. 

There  is  another  derivation  of  Clapeyron's  equation  which  is  very 
instructive.  This  is  based  on  the  use  of  the  Gibbs  free  energy  G.  In  the 
last  section  we  have  seen  that  this  quantity  must  be  equal  for  two  phases 
in  equilibrium  at  the  same  pressure  and  temperature,  and  that  if  one 
phase  has  a  lower  value  of  G  than  another  at  given  pressure  and  tempera- 
ture, it  is  the  stable  phase  and  the  other  one  is  unstable.  We  can  verify 
these  results  in  an  elementary  way.  We  know  that  in  going  from  liquid 
to  vapor,  the  latent  heat  L  is  the  difference  in  enthalpy  between  gas  and 
liquid,  or  L  =  //„  —  HI.  Bui  if  the  change  is  carried  out  in  equilibrium, 
the  heat  absorbed  will  also  equal  T  (1$,  so  that  the  latent  heat  will  be 
T(SU  —  Si).  Equating  these  values  of  HK  latent  heat,  we  have 

//.  -  Hi  -  L  =  T(Sa  -  ,SU 
or 

Hg  -  TKtt  =  Hi  -  TStj         Gt,  =  <;,,  (4.4) 

or  our  previous  condition  that  the  Gibbs  free  energy  should  be  the  same 
for  the  two  phases  in  equilibrium.  Since  this  must  be  true  at  each  point 
of  the  VcMpor  pressure  line  in  the  P-T  plane,  we  can  find  the  slope  of  the 
vapor  pressure  curve  from  the  condition  that,  as  P  and  T  change*  in  the 
same  way  for  both  phases,  Ga  and  GI  must  undergo  equal  changes.  That 
is  to  say,  we  set 

d(G»  ~ 


n  \  —  n       i     v   °          l)  \  AT  _L 

0-Gi)  -  o  =  ^ — -^ — ^^y  + 

=  ~-(Sa  -  Si)dT  +  (V0  -  Vt)dP 
=  -L^~  +  (V0-  VfidP, 

dP  _  L 

W  =  (V~-  Vd~T'  (4'5) 

which  is  Clapc^yron's  equation.     In  deriving  Eq.  (4.5),  we  used  the  rela- 
tions (dG/dT),  =  -S,  (dG/dP)T  =  F,  from  Chap.  II. 

Hap^yron's  equation,  as  an  exact  result  of  thermodynamics,  is  useful 
in  several  ways.  In  the  first  place,  we  may  have  measurements  of  the 
equation  of  state  but  not  of  the  latent  heat.  Then  we  can  compute  the 
latent  heat.  This  is  particularly  useful  for  instance  at  high  pressures, 


176  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XI 

where  measurements  of  volume  and  temperature  are  fairly  easy,  but 
where  ealorimetric  measurements  such  as  would  be  required  to  find  the 
latent  heat  are  very  difficult.  Or,  in  the  second  place,  we  may  know  the 
latent  hoat  and  then  we  can  find  the  slope  of  the  equilibrium  curve,  and  by 
integration  we  may  find  the  whole  curve.  We  shall  discuss  the  applica- 
tion of  this  method  to  tho  vapor  pressure  curve  in  the  next  sections. 
Finally,  we  may  have  measurements  of  both  the  latent  heat  and  tho 
equilibrium  curve,  but  may  not  be  sure  of  their  accuracy.  We  can  test 
them,  and  perhaps  improve  them,  by  soring  whether  the  experimental 
values  satisfy  Chipeyron's  equation  exactly.  If  they  do  not,  it  is  certain 
evidence  that  they  are  in  error. 

5.  The  Integration  of  Clapeyron's  Equation  and  the  Vapor  Pressure 
Curve.  —  The  integration  of  Clapoyron's  equation  to  got  the  vapor  pres- 
sure curve  over  a  liquid  or  solid  from  a  measurement  of  the  latent  hoat  is 
one  of  its  principal  uses.  We  may  write  the  integral  of  Eq.  (4.3)  in  tho 
form 


-  f 

~  J 


LtiT 


This  can  be  evaluated  exactly  if  \ve  know  the  latent  hoat  L,  tho  volume 
of  tho  gas,  and  tho  volume  of  tho  solid,  as  functions  of  temperature.  In 
many  actual  oases  we  know  these1  only  approximately,  but  we  can  use 
thorn  to  got  an  approximate;  vapor  pressure  curve.  For  instance,  the 
simplost  approximation  is  to  assume  that  tho  latent  heat  is  a  constant, 
independent  of  temperature.  Furthermore,  in  the  ease  of  low  tempera- 
ture, where  tho  volume  of  tho  gas  will  bo  very  large  compared  with  the 
volume  of  the  solid  or  liquid,  we  may  neglect  tho  latter  and  furthermore 
assume  that  the  gas  obeys  tho  perfect  gas  law.  Then  Vg  —  Vs  =  nRT/P, 
approximately,  and  Eq.  (4.3)  becomes 


dT 

JP  T 

(5.2) 


PdT 

Equation  (5.2)  holds  whether  L  is  constant  or  not.     Assuming  it  to  be 
constant,  we  can  integrate  and  find 

In  P  =    --T>TT  +  const., 


or 

L__ 

P  =  const,  e   nRT.  (5.3) 

Equation  (5.3)  giving  P  in  terms  of  T  gives  a  first  approximation  to  a 


SBC.  5]  THE  EQUILIBRIUM  OF  SOLIDS  177 

vapor  pressure  curve.  Plainly  it  approaches  e~°°  =  0  at  low  tempera- 
ture, while  as  the  temperature  increases  the  pressure  rapidly  increases, 
agreeing  with  the  observed  form  of  the  curve.  By  making  more  elaborate 
assumptions,  taking  account  of  the  variation  of  L  with  pressure  and 
temperature  and  the  deviation  of  the  volume  of  the  gas  from  that  for  a 
perfect  gas,  the  equation  of  the  vapor  pressure  curve  can  be  obtained  as 
accurately  as  we  please.  Tho  formula  (5.3)  in  particular  becomes 
entirely  unreliable  near  the  critical  point.  Since  the  latent  heat 
approaches  zero  as  we  approach  the  critical  point,  and  the  volumes  of 
liquid  and  gas  approach  each  other  at  the  same  place,  f,he  ratio  dP/dT 
becomes  indeterminate  and  more  accurate  work  is  necessary  to  find  just 
what  the  slope  of  the  vapor  pressure  curve  is.  In  spite  of  this  difficulty, 
the  formula  (5.3)  is  a  very  useful  one  at  temperatures  well  below  the 
critical  point.  The  constant  factor,  of  course,  must  be  obtained  as  far  as 
thermodynamics  is  concerned  from  a  measurement  of  vapor  pressure  at 
one  particular  temperature. 

To  find  the  correct  vapor  pressure  equation,  we  shall  determine  the 
variation  of  latent  heat  with  temperature4.  In  introducing  the*  enthalpy 
H  =  U  +  PV  in  Chap.  11,  we  saw  that  the  change  in  enthalpy  in  any 
process  equalled  the  heat  absorbed  at  constant  pressure.  Now  the  latent 
heat  is  absorbed  at  constant  pressure;  therefore  it  equals  the  change  of 
the  enthalpy  between  solid  and  gas.  That  is, 

L  =  II,  -  //,  (5.4) 

Now  we  can  find  the  change*  in  L,  for  an  arbitrary  change  of  pressure 
and  temperature.  We  have 

+ 


I'.-    V.  -  Tf  +  7<  <//'  +  (Cfl  -  CMT,      (5.5) 

when1  we  have  used  thermodynamic  relations  from  the  table  in  Chap.  II. 
Now  we  assume?  as  in  the  last  paragraph  that  the  ve>lume  of  the  solid  can 
be  neglected  and  that  the  gas  obeys  the  perfect  gas  law.  The*  gas  law 
gives  at  once  Va  —  T(dVti/dT}p  =  0.  Thus  the  first  bracket  is  zero,  and 
we  have 

^  =  Cf.  -  CPf  (5.0) 

In  Chap.  VIII  \ve  have  expressed  the  specific  heat  of  a  gas  as 

Cf,  =  %nR  +  nC,.  (5.7) 


178  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XI 

Introducing  this  value  into  Eq.  (5.6)  and  working  with  one  mole,  we  can 
integrate  Eq.  (5.6)  to  get  the  latent  heat  L,  in  terms  of  L0,  the  latent  heat 
of  vaporization  at  the  absolute  zero: 


L  =  Lo  4-  %RT  +  (C*  -  Cf.)dT.  (5.8) 

Dividing  by  RT~  and  integrating  with  respect  to  7T,  wo  have  from  Eq. 
(5.2) 

j        *;  CT  jr  CT 

In  P  =  -~  °  +  ?  In  T  -         ~          (Cf,  -  CJdT  +  const.     (5.9) 


~  T 
This  expression,  or  the  corresponding  one 

"'*   _*£.          fT*LfT(t.     - 

P  =  const.  T  c   ~RTc     JoJmJo(Cl"       '  (5  10) 

is  the  complete  formula  for  a  vapor  pressure  curve,  as  obtained  from 
thermodynamics,  in  the  region  where  the  vapor  behaves  like  a  perfect 
gas.  We  shall  see  in  the  next  section  that  statistical  mechanics  can 
supply  the  one  missing  feature  of  Eqs.  (5.9)  and  (5.10):  it  can  give  the 
explicit  value  of  the  undetermined  constant. 

It  is  interesting  to  note  the  behavior  of  the  latent  heat  of  vaporization, 
as  given  by  Eq.  (5.8),  through  wide  temperature  ranges.  At  low  tempera- 
tures, since  CV.  and  Ct  both  are  very  small,  the  latent  heat  increases  with 
temperature.  This  tendency  is  reversed,  however,  as  the  specific  heat 
of  the  solid  becomes  greater  than  that  of  the  gas,  which  it  always  does. 
The  latent  heat  then  begins  to  fall  again.  At  the  triple  point,  as  we  have 
stated,  the  latent  heat  of  vaporization  of  the  solid  just  belo\\  the  triple 
point  equals  the  sum  of  the  latent  heat  of  fusion  and  the  latent  heat  of 
vaporization  of  the  liquid,  directly  above  the  triple  point.  Above  this 
temperature,  Eq.  (5.8)  is  to  be  replaced  by  one  in  which  C/>  of  the  liquid 
appears  rather  than  that  of  the  solid.  This  allows  us  to  use  the  same 
sort  of  method  for  finding  the  vapor  pressure  over  a  liquid.  Finally,  as 
the  temperature  approaches  the  critical  point,  it  is  no  longer  correct  to 
approximate  the  vapor  by  a  perfect  gas,  so  that  neither  Eq.  (5.2)  nor 
(5.8)  is  applicable,  though  we  already  know  that  the  latent  heat 
approaches  zero  at  the  critical  point,  and  of  course  Clapeyron's  equation 
can  be  applied  here  as  well  as  elsewhere. 

6.  Statistical  Mechanics  and  the  Vapor  Pressure  Curve.—  From 
statistical  mechanics,  we  know  how  to  write  down  the  Gibbs  free  energy 
of  the  solid  and  perfect  gas  explicitly.  All  we  need  do,  then,  to  find  the 
complete  equation  of  the  vapor  pressure  curve  is  to  equate  these  quanti- 
ties. Thus,  remembering  that  (dG/dT)P  =  -S,  we  can  write  the  Gibbs 


SEC.  6]  THE  EQUILIBRIUM  OF  SOLIDS  179 

free  energy  of  the  solid 

G.  =  f/o  -    f   8  dT 

Jo 

(*T          fTr, 

=  Uo  -    \    dT\    ±*dT.  (6.1) 

Jo        Jo     l 

Using  Kqs.  (1.19)  and  (1.20)  of  Chap.  VIII,  this  can  bo  rewritten 

CTdT  CT 

G.  =  £/„-  H    ~\    CP.dT.  (6.2) 

Jo   I   Jo 

Here  UQ  is  the  internal  energy,  or  free  energy,  at  absolute  zero,  a  function 
of  pressure  only.  Next  we  wish  the  Gibbs  free  energy  of  the  gas.  We 
use  Eq.  (1.25)  of  Chap.  VIII.  We  note,  however,  that  l/0  is  used  in  that 
equation  in  a  different  sense  from  what  it  has  been  here;  for  there  it 
means  the  internal  potential  energy  of  the  gas  at  absolute  zero,  while 
here  we  have  used  it  for  the  internal  energy  of  the  solid  at  absolute  zero. 
It  is  plain  that  the  energy  of  the  gas  at  absolute  zero  must  be  greater  than 
that  of  the  solid  by  just  the  latent  heat  of  vaporizal  ion  at  the  absolute  zero, 
orLo.  Using  this  fact,  we  have 

Gg  =  C/o  +  Lo  -  ^RT  In  T  +  RT  In  P 


Equating  Eqs.  (6.2)  and  (6.3),  we  have 

ln  p  -  -      +    ln  T  ~     '          (Cft  -  c')rfr  +  '•     (6'4) 


Equation  (6.4)  is  the  general  one  for  vapor  pressure,  and  it  shows  that  the 
undetermined  constant  in  In  P,  in  Eq.  (5.9),  is  just  the  chemical  constant 
that  we  have  already  determined  in  Eq.  (3.16)  of  Chap.  VIII.  The 
simplest  experimental  method  of  finding  the  chemical  constants  is  based 
on  Eq.  (6.4)  :  one  measures  the  vapor  pressure  as  a  function  of  the  tem- 
perature, finds  the  specific  heats  of  solid  and  gas,  so  that  one  can  calculate 
the  term  in  the  specific  heats,  and  computes  the  quantity 

T 


as  a  function  of  temperature.  Plotting  as  a  function  of  1/T,  Eq.  (6.4) 
says  that  the  result  should  be  a  straight  line,  whose  slope  is  —  L0  and 
whose  intercept  on  the  axis  1/T  =  0  should  be  the  chemical  constant  /. 
This  gives  a  very  nice  experimental  way  of  checking  our  whole  theory  of 
vapor  pressure  and  chemical  equilibrium:  the  same  chemical  constants 
obtained  from  vapor  pressure  measurements  should  correctly  predict  the 


180  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XI 

results  of  chemical  equilibrium  experiments.  It  is  found  that  in  fact  they 
do,  within  the  orror  of  experiment. 

7.  Polymorphic  Phases  of  Solids.  Kxpcrimontally,  polymorphism 
at  high  pressures  is  ordinarily  observed  by  discontinuous  changes  of 
volume.  As  the  pressure  is  changed  at  constant  temperature,  the  volume 
changes  smoothly  as  long  as  we  are  dealing  with  one  phase  only.  At  jhe 
equilibrium  pressures,  however,  the  volume  suddenly  changes  discontinu- 

OUslv   to   another  V^luer    whjf.fr    nf  nmirsn   jfi   always   fijTmllor  for   fho   ]} igJi 

pressure  modification.  Then  another  smooth  change  continues  from  the 
transition  pressure.  The  measurement  thus  gives  not  only  the  pressure 
and  temperature  of  a  point  on  the  equilibrium  line,  but  the  change  of 
volume  as  well.  Clapeyron's  equation  of  course  applies  to  equilibrium 
lines  between  solids,  and  that  moans  that  from  the  observed  slope  of  the 
transition  line  and  the  observed  change  of  volume,  we  can  find  the  latent 
heat  of  the  transition,  even  though  a  direct  thermal  measurement  of  this 
latent  heat  might  be  very  difficult.  Thus  we  can  find  energy  and  entropy 
differences  between  phases. 

It  is  very  hard  to  say  anything  of  value  theoretically  about  polymor- 
phic transitions.  The  changes  of  internal  energy  and  entropy  between 
phases  are  ordinarily  quite  small.  Any  calculation  that  we  should  tiy  to 
make  of  the  thermodynamics  properties  of  each  phase  separately  would 
have  errors  in  both  those  quantities,  at  least  of  the  order  of  magnitude 
of  the  difference  which  is  being  sought.  Thus  it  is  almost  impossible  to 
predict  theoretically  which  of  two  phases  should  bo  stable  under  given 
conditions,  or  whore  tho  equilibrium  line  between  thorn  should  lie. 
Nature  apparently  is  faced  with  essentially  the  same  problem,  for  in  many 
cases  polymorphism  scorns  to  be  a  haphazard  phenomenon.  It  has  been 
impossible  to  make  any  generalizations  or  predictions  as  to  what  sub- 
stances should  be  polymorphic  and  what  should  not,  and  in  many  cases 
substances  that  are  similar  chemically  show  quite  different  behavior  as  to 
polymorphism.  We  can,  however,  say  a  little  from  thermodynamics  as 
to  the  stability  of  phases  and  the  nature  of  equilibrium  lines. 

We  can  think  of  two  limiting  sorts  of  transitions:  one  in  which  the 
transition  always  occurs  at  tho  saino  tcmporaturo  independent  of  pros- 
sure,  the  other  whore;  it  is  always  at  the  same  pressure  independent  of 
temperature.  These  would  correspond  to  vortical  and  horizontal  lines 
respectively  in  Fig.  XI-3.  In  Clapoyron's  equation  dP/dT  =  L/TAV, 
these  correspond  to  the  case  dP/dT  =  oo  or  0  respectively.  Thus  in  the 
first  case  we  must  have  AF  =  0,  or  the  two  phases  have  the  same  volume, 
in  which  case  pressure  does  not  affect  the  transition.  And  in  the  second 
case  L  —  0,  or  AS  =  0,  there  is  no  latent  heat,  or  the  two  phases  have 
the  same  entropy,  in  which  case  temperature  does  not  affect  the  transition. 
Put  differently,  increase  of  pressure  tends  to  favor  the  phase  of  small 
volume,  increase  of  temperature  favors  the  phase  of  large  entropy. 


SEC.  7]  THE  EQUILIBRIUM  OF  SOLIDS  181 

Of  course,  in  actual  cases  we  do  not  ordinarily  find  two  phases  witli 
just  the  same  volume  or  just  the  same  entropy.  Qn_accaunt_oL_the 
parallelism  between  the  entropy  and _the_vplumc.  thore  is  a  tendency  for  n 
phase  of  larger  volume  also  to  haveji Jarger_on topy  Thus  t.ho  fonHpuny 
isjfor_the  latent  heat  and  theHiange  of  vojumMpJha/vo  the  same  sign.  so 
that  byjClapeyron's  equation  dP/dT  tends  to  be  positive,  or  the  equilib- 
rium lines  tend  to  slope  upward  tcTflie "  ngliTTTi  the  phase  diagram.  A 
statistical  study  of  the  phase  diagrams  of  many  substances  shows  that  in 
fact  this  is  the  case,  though  of  course  there  an1  many  exceptions.  In 
fact,  there  is^evenji  tendency  joward  a  fairly  definite  slope  dP/dT  charac- 
teristic  of  many  substances,  which  according;  To  Bridgman1  is  a  change  of 
somcfhing  less  ilian  12,000  atm.  for  a  temperature  range'oT  200°. 


In  each  region  of  the  P-T  diagram  then1  Ts  only  one  stable  phase, 
except  on  equilibrium  lines  or  at  triple  points  where  there  arc1  two  or  three 
respectively.  But  a  phenomenon  analogous  to  supercooling  is  very 
widespread  in  transitions  between  solids.  Particularly  at  temperatures 
well  below  the  melting  point,  transitions  occur  very  slowly.  A  stable 
phase  can  often  be  carried  into  a  region  where  it  is  unstable,  by  change  of 
pressure  or  temperature,  and  it  may  take  a  very  long  time  to  change  over 
to  the  phase  stable  at-  that  pressure  and  temperature.  This  makes  it- 
very  hard  in  many  cases  to  determine  equilibrium  lines  with  great  accu- 
racy, for  near  equilibrium  the  transitions  tend  to  be  slower  than  far  from 
equilibrium.  It  also  makes  it  hard  to  continue  investigations  of  poly- 
morphism to  low  temperatures.  In  Fig.  XI-3,  for  instance,  the  lines  are 
continued  about  as  far  toward  low  temperatures  as  it  is  practicable  to  go. 
Sometimes  these  slow  transitions  can  be  of  practical  value,  as  in  the  case 
of  alloys.  It  often  happens  that  a  modification  stable  at  high  tempera- 
ture, but  unstable  at  room  temperature,  has  properties  that  arc;  desirable 
for  ordinary  use.  In  such  a  case  the  material  can  often  be  quenched  and 
cooled  very  rapidly  from  the  high  temperature1  at  which  the  desired  modi- 
fication is  stable.  The  material  is  almost  instantly  cooled  so  far  down 
below  its  melting  point  that  the  transition  to  the  phase;  stable  at  room 
temperature  is  so  slow  as  to  be;  negligible  for  practical  purposes.  Thus 
the  desired  phase  is  made  practically  permanent  at  room  temperature, 
though  it  may  not  be  thermodynamically  stable.  The*  ordinary  process 
of  hardening  or  tempering  steel  by  quenching  is  an  example  of  this 
process.  In  some  cases  such  unstable  phases  change;  ove;r  to  the  stable 
form  in  a  period  of  years,  but  in  the  really  valuable;  cases  the  rate  is  so 
slow  that  it  can  be  disregarded  e;ven  for  many  years.  A  moderate  heat-- 
ing, however,  can  accelerate  the*  process  so  much  as  to  change  the  prop- 
erties entirely,  as  a  moderate  heating  can  destroy  the  temper  or  hardness 
of  steel. 

1  P.  W.  Bridgman,  Pmc.  Am.  Acnd.,  72,  45  (1937);  soe  p.  129 


CHAPTER  XII 
VAN  DER  WAALS'  EQUATION 

Real  gases  do  not  satisfy  the  perfect  gas  law  PV  =  nRT,  though  they 
approach  it  more  and  more  closely  as  they  become  less  and  less  dense. 
There  is  no  simple  substitute  equation  which  describes  them  accurately. 
There  is,  however,  an  approximate  equation  called  Van  der  Waals' 
equation,  which  holds  fairly  accurately  for  many  gases  and  which  is  so 
simple  and  reasonable  that  it  is  used  a  great  deal.  This  equation  is  not 
really  one  that  can  be  exactly  derived  theoretically  at  all.  Van  der 
Waals,  when  he  worked  it  out,  thought  he  was  giving  a  very  general  and 
correct  deduction,  but  it  has  since  been  seen  that  his  arguments  were  not 
conclusive.  Nevertheless  it  is  a  plausible  equation  physically,  and  it  is 
so  simple  and  convenient  that,  it  is  very  valuable  just  as  an  empirical 
formula.  We  shall  give,  first,  simply  a  qualitative  argument  for  justify- 
ing the  equation,  then  show  to  what  extent  it  really  follows  from  statistical 
mechanics.  Being  an  equation  of  state,  thermodynamics  by  itself  can 
give  no  information  about  it ;  we  remember  that  equations  of  state  have 
to  be  introduced  into  thermodynamics  as  separate  postulates.  Only 
statistical  mechanics  can  be  of  help  in  deriving  it. 

1.  Van  der  Waals'  Equation. — Van  der  Waals  argued  that  the  perfect 
gas  law  needed  revision  for  real  gases  on  two  accounts.  In  the  first  place, 
he  considered  that  the  molecules  of  real  gases  must  attract  each  other, 
exerting  forces  on  each  other  which  are  neglected  in  deriving  the  perfect 
gas  law.  The  fact  that  gases  condense  to  form  liquids  and  solids  shows 
this.  Surely  the  only  thing  that  could  hold  a  liquid  or  solid  together 
would  be  intermolecular  attractions.  These  attractions  he  considered 
as  pulling  the  gas  together,  just  as  an  external  pressure  would  push  it 
together.  There  is,  in  other  words,  an  internal  pressure  which  can  assist 
the  external  pressure.  In  a  liquid  or  solid,  the  internal  pressure  is  great 
enough  so  that  even  with  no  external  pressure  at  all  it  can  hold  the 
material  together  in  a  compact  form.  In  a  gas  the  effect  is  not  so  great 
as  this,  but  still  it  can  decrease  the  volume  compared  to  the  corresponding 
volume  of  a  perfect  gas.  To  find  the  way  in  which  this  internal  pressure 
depends  on  the  volume,  Van  der  Waals  argued  in  the  following  way. 
Consider  a  square  centimeter  of  surface  of  the  gas.  The  molecules  near 
the  surface  will  be  pulled  in  toward  the  gas  by  the  attractions  of  their 
neighbors.  For,  as  we  see  in  Fig.  XII-1,  these  surface  molecules  are 
subjected  to  unbalanced  attractions,  while  a  molecule  in  the  interior  will 

182 


Ssr.  1)  VAN  DER  WAALS'  EQUATION  183 

have  balanced  forces  from  the  molecules  on  all  sides.  Now  the  range  of 
action  of  these  intermolecular  forces  is  found  to  be  very  small.  Thus 
only  the  immediate  neighbors  will  be  pulling  a  given  molecule  to  any 
extent.  To  indicate  this,  we  have  drawn  a  thin  layer  of  gas  near  the 
surface,  including  all  the  molecules  exerting  appreciable  forces  on  the 
surface  molecules.  Now  the  total  force  on  one  surface  molecule  will  be 
proportional  to  the  number  of  molecules  that  pull  it.  That  is,  it  will  be 
proportional  to  the  number  of  molecules 
per  unit  volume,  times  the  volume  close 
enough  to  the  molecule  to  contribute 
appreciably  to  the  attraction.  The  total 
force  on  all  the  molecules  in  a  square 
centimeter  of  the  surface  layer  will  be 
proportional  to  the  number  of  molecules 
in  this  square  centimeter  times  the  force  Fi«.  xil-i.  intermolecular  attra«- 
011  each,  so  that  it  will  be  proportional  tlons> 

to  the  square  of  the  number  of  molecules  per  unit  volume,  or  to  (JV/F)2,  if 
there  are  N  molecules  in  the  volume  V,  or  to  (n/V)'2,  where  n  is  the  num- 
ber of  moles.  But  the  force  on  the  molecules  in  a  square  centimeter  of 
surface  area  is  just  the  internal  pressure,  so  that 


Internal  pressure  =  a(~t7y  '  (1-1) 

where  a  is  a  constant  characteristic  of  the  gas. 

The  second  correction  which  Van  dcr  Waals  made  was  on  account  of 
the  finite  volume  of  the  molecules.  Suppose  the  actual  molecules  of  a  gas 
were  rather  large  and  that  the  density  was  such  that  they  filled  up  a  good 
part  of  the  total  volume.  Then  a  single  one  of  the  molecules  which  we 
might  consider,  batting  around  among  the  other  molecules,  would  really 
not  have  so  large  a  space  to  move  in  as  if  the  other  molecules  were  not 
there.  Instead  of  having  the  whole  volume  V  at  its  disposal,  it  would 
move  much  more  as  if  it  were  in  a  smaller  volume.  If  there  are  n  moles 
of  molecules  present,  and  the  reduction  in  effective  volume  is  b  per  mole, 
then  it  acts  as  if  its  effective  volume  were 

Effective  volume  =  V  —  nb.  (1.2) 

If  the  volume  were  reduced  by  this  amount,  the  pressure  would  be  cor- 
respondingly increased,  since  the  molecule  would  collide  with  any  element 
of  surface  more  often. 

Making  both  these  corrections,  then,  Van  dor  Waals  assumed  that 
the  equation  of  state  of  an  imperfect  gas  was 


+  a(F  -  nb)  =  nRT.  (1.3) 


184 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  XII 


This  is  Van  der  Waals'  equation.  We  shall  later  come  to  the  question 
of  how  far  it  ran  be  justified  theoretically  by  statistical  mechanics.  First, 
however,  we  shall  study  its  properties  as  an  equation  of  state  and  see  how 
useful  it  is  in  describing  the  equilibrium  of  phases. 

2.  Isothermals  of  Van  der  Waals'  Equation. — In  Fig.  XII-2  we  give 
isothermals  as  computed  by  Van  der  Waals'  equation.  At  first  glance1, 
they  are  entirely  different  from  the  actual  isothermals  of  a  gas,  as  shown  in 
perspective  in  Fig.  XI-4,  because  for  low  temperatures  the  isothermals 
show  a  maximum  and  minimum,  the  minimum  corresponding  in  some 
cases  to  a  negative  pressure.  But  a  little  reflection  shows  that  this 
situation  is  not  alarming.  We  note  that  there  is  one  isothermal  at  which 
the  maximum  and  minimum  coincide,  so  that  there  is  a  point  of  inflection 


FIG.  XII-2.  — Isothermals  of  Van  dor  Waata*  equation. 

of  the  curve  here.  This  is  the  point  marked  C  on  Fig.  XII-2.  At  every 
lower  temperature,  there  arc  three  separate  volumes  corresponding  to 
pressures  lying  between  the  minimum  and  maximum  of  the  isothermal. 
This  is  indicated  by  a  horizontal  line,  corresponding  to  constant  pressure, 
which  is  drawn  in  the  figure  and  which  intersects  one  of  the  isothermals  at 
V\,  V%,  tind  F».  We  may  now  ask,  given  the  pressure  and  temperature 
as  determined  by  this  horizontal  line  and  this  isothermal  respectively, 
which  of  the  three  volumes  will  the  substance  really  have?  This  is  a 
question  to  which  there  is  a  perfectly  definite  answer.  Thermodynamics 
directs  us  to  compute  the  Gibbs  free  energy  of  the  material  in  each  of 
the  three  possible  states,  and  tells  us  that  the  one  with  the  lowest 
Gibbs  free  energy  will  be  the  stable  state.  The  material  in  one  of  the 
other  states,  if  it  existed,  would  change  irreversibly  to  this  stable  state. 
We  shall  actually  compute  the  free  energy  in  the  next  section,  and  shall 
find  which  state  has  the  lowest  value.  The  situation  proves  to  be  the 


SEC.  2] 


VAN  DEK  WAALS9  EQUATION 


185 


following.  Suppose  we  go  along  an  isothermal,  increasing  the  volume  at 
constant  temperature,  and  suppose  the  isothermal  lies  below  the  tempera- 
ture corresponding  to  C.  Then  at  first,  the  smallest  volume  V\  has  the 
smallest  Gibhs  free  energy.  A  pressure  is  reached,  however,  at  which 
volumes  V\  and  F3  correspond  to  the  same  free  energy.  At  still  lower 
pressures,  F3  corresponds  to  the  lowest  free  energy.  The  state  F2  has  a 
higher  free  energy  than  either  FI  or  Fa  under  all  conditions,  so  that  it  is 
never  stable.  We  see,  then,  that  above  a  certain  pressure  (below  a  certain 
volume)  the  state  of  smallest  volume  is  stable,  at  a  definite  pressure  this 
state  and  that  of  largest  volume  can  exist  together  in  equilibrium,  and 
below  this  pressure  only  the  phase  of  largest  volume  can  exist.  But  this 
is  exactly  the  behavior  to  be  expected  from  experience  with  actual  changes 
of  phase. 

The  isothermals  of  Van  dor  Waals'  equation,  then,  correspond  over 
part  of  their  length  to  states  that  are  not  thermodynamieally  stable,  in 


Fm.  XII-3. — Isothcrmals  of  Van  der  Waals'  equation,  showing  equilibrium  of  liquid  and 

gas. 

the  sense  that  their  free  energy  is  greater  than  that  of  other  states,  also 
described  by  the  same;  equation,  at  the  same  pressure  and  temperature. 
In  Fig.  XII-3  we  give  revised  isothormals,  taking  this  change  of  phase  into 
account.  In  this  figure,  corresponding  to  each  pressure  and  temperature, 
only  the  stable  phase  is  shown.  In  the  region  whore  two  phases  are  in 
equilibrium,  we  draw  horizontal  lines,  as  usual  in  such  diagrams,  indicat- 
ing that  the  pressure  and  temperature  are  constant  over  the  whole  range 
of  volumes  between  the  two  phases  in  which  the  stable  state  of  the  system 
is  a  mixture  of  phases.  The  isothermals  of  Fig.  XII-3  are  plainly  very 
similar  to  those  of  actual  gases  and  liquids. 

From  Fig.  XII-3,  it  is  plain  that  the  critical  point  is  the  point  C  of 
Fig.  XII-2,  at  which  the  maximum  and  the  minimum  of  the  isothermal 
coincide.  We  can  easily  find  the  pressure,  volume,  and  temperature;  of 
the  critical  point  in  terms  of  the  constants  a  and  6,  from  this  condition. 
The  most  convenient  way  to  state  the  condition  analytically  is  to  demand 
that  the  first  and  second  derivatives  of  P  with  respect  to  V  for  an  iso- 
thermal vanish  simultaneously  at  the  critical  point.  Thus,  denoting  the 


186  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XII 

critical  pressure,  volume,  and  temperature  by  Pc,  Ve,  Tc,  we  have 


-  n  =  _ 
~ 


_ 

e~  nb       Vs,' 

nRT, 


'        '(Fe  -  n6) 
_  _  2nRTc      _ 

3FVr  (V,  -  »6)*      "FJ  '         (F;  -  nbY  ~    V*'        (     } 

We  can  solve  Kqs.  (2.1),  (2.2),  (2.3)  simultaneously  for  Pc,  Tc,  and  Fc. 
Dividing  Eq.  (2.3)  by  Kq.  (2.2)  we  at  once  find  Ff.  Substituting  this  in 
Eq.  (2.2),  we  can  solve  for  T..  Substituting  both  in  Eq.  (2.1),  we  find 
Pc.  In  this  way  we  obtain 

P«  =  27F       v°  =  3nb>       RT''WH  (2'4) 

Equations  (2.4)  give  the  critical  point  in  terms  of  a  and  b.  Con- 
versely, from  any  two  of  the  Eqs.  (2.4)  we  can  solve  for  a  and  b  in  terms 
of  the  critical  quantities.  Thus,  from  the  first  and  third,  we  have 

'-i£- 

These  equations  allow  us  to  make  a  calculation  of  the  critical  volume: 


Fc  (Van  der  Waals)  =  3nb  =  c-  (2.6) 

o     rc 

If  Van  der  Waals'  equation  were  satisfied  exactly  by  the  gas,  the  critical 
volume  determined  in  this  way  from  the  critical  pressure  and  temperature 
should  agree  with  the  experimentally  determined  critical  volume.  That 
this  is  not  the  case  will  be  shown  in  a  later  chapter.  The  real  critical 


volume  and  ^     p      are  not  far  different,  but  the  latter  is  larger.     This  is 
o     rc 

one  of  the  simplest  ways  of  checking  the  equation  and  seeing  that  it  really 
does  not  hold  accurately,  though  it  is  qualitatively  reasonable. 

Using  the  values  of  PCJ  Vc,  and  Tc  from  Eq.  (2.4),  we  can  easily  write 
Van  der  Waals'  equation  with  a  little  manipulation  in  the  form 


Ip_ 3_¥Z  _  11  =  §  L 
p<    (-IV  LFc    3J    g7'- 
\    CJ  \ 


(2.7) 


This  form  of  the  equation  is  expressed  in  terms  of  the  ratios  P/PC,  V/VC, 
T/Tc,  showing  that  if  the  scales  of  pressure,  volume,  and  temperature  are 
adjusted  to  bring  the  critical  points  into  coincidence,  the  Van  der  Waals' 


SBC.  3]  VAN  DER  WAAL&  EQUATION  187 

equations  for  any  two  gases  will  agree.  This  is  called  the  law  of  cor- 
responding states.  Real  gases  do  not  actually  satisfy  this  condition  at  all 
accurately,  so  that  this  is  another  reason  to  doubt  the  accuracy  of  Van 
der  Waals'  equation. 

3.  Gibbs  Free  Energy  and  the  Equilibrium  of  Phases  for  a  Van  der 
Waals  Gas.  —  We  have  seen  that  the  equilibrium  between  the  liquid  and 
vapor  phase  is  determined  by  setting  the  Gibbs  free  energy  equal  for  the 
two  phases.  Let  us  carry  out  this  calculation  for  Van  der  Waals'  equa- 
tion. From  Eq.  (4.2),  Chap.  II,  we  have  dO  =  V  (IP  -  ASY  dT.  Thus 
we  can  calculate  the  Gibbs  free  energy  by  integrating  this  expression.  Wo 
arc  interested  only  in  comparing  free  energies  at  various  points  along  an 
isothermal,  however,  and  for  constant  temperatures  we  can  set  the  last 
term  equal  to  zero,  so  that  dG  =  V  dP  along  an  isothermal.  This  is  not 
a  convenient  form  for  calculation,  unfortunately,  for  Van  der  Waals' 
equation  cannot  be  solved  for  the  volume  in  terms  of  the  pressure  con- 
veniently. It  involves  the  solution  of  a  cubic  equation,  and  this  can 
usually  be  avoided  by  some  means  or  other.  To  avoid  this  difficulty,  we 
shall  instead  compute  the  Ilelmholtz  free  energy  A  =  (V  —  PV,  and  then 
find  the  Gibbs  free  energy  from  it.  We  have 

dA  =  -PdV  -  SdT  =  -PdV 
for  an  isothermal  process.     Thus 

A  —  —  JP  dV  +  function  of  temperature 


nRT         M2a\  ,„    ,   ,       ..        ,  . 
v  ^T~  /   ~~    \ri  r*   ~*~  funrtlon  °f  temperature 


-/( 

=  —nRT  In  (V  —  nb)  —  ~^-  +  function  of  temperature,     (3.1) 

and  for  the  Gibbs  free  energy  we  have 
G  =  A  +PV 

=  PV  —nRT In  (V  —  nb) y-  +  function  of  temperature 

nRTV        2n2a         nrri,    /T,         1X    ,   r       A.        f, 

=  ^ £ Tr nRTm  (V  —  nb)  +  function  of  temperature. 

V  —  no          V 

(3.2) 

Equation  (3.2)  expresses  G  as  a  function  of  volume  arid  temperature.  We 
wish  it  as  a  function  of  pressure  and  temperature,  and  it  cannot  con- 
veniently be  put  in  this  form  in  an  analytic  way,  on  account  of  the 
difficulty  of  solving  Van  der  Waals'  equation  for  the  volume.  It  is  an 
easy  matter  to  compute  a  table  of  values,  however.  We  plot  curves,  like 
Fig.  XII-3,  for  pressure  as  a  function  of  volume  in  Van  der  Waals' 
equation,  compute  values  of  G  from  Eq.  (3.2)  for  a  number  of  values  of  the 


188 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  XII 


volume,  and  read  off  the  corresponding  pressures  from  the  curves  of 
pressure  against  volume.  In  this  way  Hie  curve  of  Fig.  XII-4  was 
obtained.  In  this,  the  Gihbs  free  energy  is  plot  ted  as  a  function  of  pres- 
sure, for  a  particular  temperature  (T  =  0.95TC  in  this  particular  case). 
It  is  seen  that  for  a  range  of  pressures,  which  in  this  case  runs  from  about 
P  =  0.74PC  to  P  =  0.84PC,  there  are  three  values  of  G  for  each  pressure,  of 
which  the  lowest  one  represents  the  stable  state.  The  lowest  curves  cross 
at  about  P  =  0.82PC,  which  therefore  represents  the  vapor  pressure  or 
point  of  equilibrium  between  the  phases,  at  this  temperature.  Compari- 
son with  Fig.  XI-7  shows  that  Fig.  XII-4  really  represents  tho  correct 
form  of  this  function;  it  corresponds  to  a  section  of  the  solid  shown  in 
Fig.  XI-7  cut  at  constant  temperature  with  suitable  rotation  of  axes. 


0.7 


09 


10 


0.8 

P/Pc 

FIG.  XII-4.--  Gibbs  free  energy 
vs.  pressure,  at  constant  tempera- 
ture, for  Van  der  Waals'  equation,  at 
T  =  0.957V. 


07 


09 


10 


08 

T/TC 

Fits.  XI 1-5.  Vapor  pressure  by 
Van  der  Waals'  equation  compared 
with  values  for  HjO,  CO*. 


We  remember  that  dG  =  V  dP  at  constant  temperature.  That  is, 
(3G/dP)T  =  V,  or  the  slope  of  the  curve  in  Fig.  XII-4  measures  the 
volume.  Clearly  at  smaller  pressure  the  stable  state  is  that  with  greater 
slope  or  greater  volume,  while  the  state  of  smallest  volume  is  stable  at  the 
high  pressures.  The  figure  makes  it  clear  why  the  phase  of  intermediate 
volume  (Vz  on  Fig.  XII-2)  is  not  stable  at  any  pressure,  since  its  free 
energy  is  never  lower  than  that  of  the  other  two  phases.  The  discon- 
tinuity in  slope  of  the  Gibbs  free  energy  at  the  point  of  equilibrium 
between  phases  measures  the  change  of  volume  in  vaporization.  This 
discontinuity  becomes  less  and  loss  as  the  temperature  approaches  the 
critical  point,  and  the  small  pointed  loop  in  the  G  curve  diminishes,  until 
finally  at  the  critical  point  it  disappears  entirely,  and  the  curve  becomes 
smooth. 

In  the  way  we  have  just  described,  we  can  find  the  Gibbs  free  energy 
for  each  temperature  and  determine  the  vapor  pressure,  and  hence  the 


SEC.  31 


VAN  DER  WAALS'  EQUATION 


189 


correct  horizontal  line  to  draw  on  Fig.  XII-2.  This  gives  us  the  vapor 
pressure  curve,  and  we  show  this  directly  in  Fig.  XII-5.  For  comparison, 
we  have  plotted  on  a  reduced  scale  the  vapor  pressures  of  water  and 
carbon  dioxide.  We  see  that  while  the 
general  form  of  the  curve  predicted  by 
Van  der  Waals'  equation  agrees  with  the 
observed  curves,  the  actual  gases  show  a 
vapor  pressure  which  diminishes  a  good 
deal  more  rapidly  with  decreasing  pres- 
sure than  the  Van  der  Waals  gas. 

Our  methods  also  allow  us  to  calculate 
the  latent  heat  of  vaporization  from  Van 
der  Waals'  equation.  We  have 


-  Ul  +  P(V{,  - 


XIl-(i.~- Latent  heat  as  func- 


Furthermore, we  can  see  by  migrating  t^fU|^;u£^ dpr  Waala R1U, 
the  equation 


and  H2O  and  OO2 


—   I         =r-     T\  I         —     P 

~av)r~  1\^T)^ 

(hat  U  —   -an-/V  +  function  of  temperature  for  a  Van  der  Waals  gas. 
Thus 


(3.3) 


All  the  quantities  of  Eq.  (3.3)  can  be  found  when  \\e  have  carried  out  the 
calculation  above,  for  that  gives  us  the  volumes  of  gas  and  liquid.  Thus 
we  can  compute  the  latent  heat  as  function  of  temperature.  To  express 
it  in  terms  of  dimensionless  quantities,  we  can  write  Eq.  (3.3)  in  terms  of 
P/Pc,  etc.,  and  find 


•J_     „_      '/    _^        \ 


I 


(3.4) 


showing  that  the  latent  heats  of  two  gases  at  corresponding  temperatures 
should  be  in  the  proportion  of  a/6  to  each  other.  In  Fig.  XI 1-6  we  plot 
the  latent  heat  L/(o/6),  as  a  function  of  temperature,  as  derived  from  Van 
der  Waals'  equation.  For  comparison  we  give  the  latent  heats  of  water 
and  carbon  dioxide.  Both  Van  der  Waals'  equation  and  experiment 
agree  in  showing  that  the  latent  heat  decreases  to  zero  at  the  critical  point 
and  the  curves  are  of  similar  shape.  However,  the  scale  is  quite  different, 
Van  der  Waals'  equation  predicting  much  too  small  a  value  for  the  latent 
heat. 


190  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XII 

We  have  now  examined  Van  der  Waals'  equation  enough  to  see  that 
it  is  very  useful  as  an  empirical  equation,  even  if  it  has  no  theoretical 
justification  at  all.  As  a  matter  of  fact,  it  can  be  justified  from  statistical 
mechanics  as  a  first  approximation,  though  no  further.  In  the  next 
sections  we  shall  take  up  this  justification,  considering  the  problem  of 
the  equation  of  state  of  a  gas  whose  molecules  attract  each  other  at  large 
distances  but  have  a  finite  size,  so  that  they  repel  each  other  if  they  are 
pushed  too  closely  into  contact.  We  begin  by  taking  up  by  statistical 
mechanics  the  general  case  of  a  gas  with  arbitrary  intermolecular  forces, 
then  specializing  to  agree  with  Van  der  Waals'  assumptions  about  the 
nature  of  the  forces. 

4.  Statistical  Mechanics  and  the  Second  Virial  Coefficient. — The  way 
to  derive  the  thermodynamic  properties  of  an  imperfect  gas  theoretically 
is  clear:  we  find  the  energy  in  terms  of  the  coordinates  and  momenta, 
compute  the  partition  function,  and  derive  the  equation  of  state  and 
specific  heat  from  it.  The  only  trouble  is  that  the  calculation  is  almost 
impossibly  difficult,  beyond  a  first  approximation.  In  this  section  we 
shall  just  derive  that  first  approximation,  which  can  be  carried  through 
without  a  great  deal  of  trouble.  To  understand  the  nature  of  the  approxi- 
mation, we  write  the  equation  of  state  in  a  series  form  which  is  often  useful 
experimentally.  At  infinite  volume,  we  know  that  the  gas  will  approach 
a  perfect  gas,  with  an  equation  of  state  PV  =  nRT,  or  PV/nRT  =  1. 
At  smaller  volumes,  the  equation  will  begin  to  deviate  from  this.  That 
is,  we  can  expand  the  quantity  PV/nRT  in  series  in  1/F;  the  term  inde- 
pendent of  1/F  will  be  unity,  but  the  other  terms,  which  are  different 
from  zero  for  imperfect  gases,  will  bo  functions  of  the  temperature.  Thus 
we  can  write 

PV  I  n\  //A2 

•*     r  i        i        rwrrrxl'1'!       i       s'i/rn\l'l'\         i  /A    -j  \ 


nRT       A    ' 

Here  the  quantity  PV/nRT  is  often  called  the  virial;  and  the  quantities 
1,  B(T),  C(7T)>  etc.,  the  coefficients  of  its  expansion  in  inverse  powers  of 
the  volume  per  mole,  V/n,  are  called  the  virial  coefficients,  so  that  B(T) 
is  called  the  second  virial  coefficient,  C(T)  the  third,  etc.  The  experi- 
mental results  for  equations  of  state  of  imperfect  gases  are  usually  stated 
by  giving  #(7"),  C(T),  etc.,  as  tables  of  values  or  as  power  series  in  the 
temperature.  It  now  proves  possible  to  derive  the  second  virial  coeffi- 
cient B(T)  fairly  simply  from  statistical  mechanics. 

The  first  thing  we  must  know  is  the  energy  of  the  gas  as  a  function 
of  its  coordinates  and  momenta.  We  use  the  same  coordinates  and 
momenta  as  in  Chap.  VIII,  Sec.  3:  the  coordinates  of  the  center  of  gravity 
of  each  molecule  and  other  coordinates  determining  the  orientation  and 
vibration  of  the  molecule.  The  difference  between  our  present  problem 


SEC.  4]  VAN  DEtt  \YAALS'  EQUATION  191 

and  the  previous  one  of  the  perfect  gas  is  that  now  we  must  add  a  term  in 
the  potential  energy  depending  on  the  relative  positions  of  the  molecules, 
coming  from  intermolecular  attractions  and  repulsions.  Strictly  speak- 
ing, these  forces  depend  on  the  orientations  of  the  molecules  as  well  as  on 
their  distances  apart,  as  is  at  once  obvious  if  the  molecules  are  very 
unsymmetrical  in  shape,  but  we  shall  neglect  that  effect  in  our  approxi- 
mate treatment  here.  That  allows  us  to  write  the  energy,  as  before*,  as  a 
sum  of  terms,  tho  first  depending  only  on  the  coordinates  of  the  centers  of 
gravity  (the  kinetic  energy  of  the  molecules  as  a  whole,  and  the  potential 
energy  of  intermolecular  forces),  and  the  second  depending  only  on  orien- 
tations and  vibrations.  Then  the  partition  function  will  factor,  the  part 
connected  with  internal  motions  separating  off  as  before  and,  since  it  is 
independent  of  volume,  contributing  only  to  the  internal  specific  heat,  and 
not  affecting  the  equation  of  state.  For  our  present  purposes,  then,  we 
can  neglect  these  internal  motions,  treating  the  gas  as  if  it  were  mon- 
atomic  and  simply  adding  on  the  internal  specific  heat  at  the  end,  using 
the  value  computed  for  the  perfect  gas.  We  must  remember,  however, 
that  this  is  only  an  approximation,  neglecting  the  effect  of  orientation  on 
intermodular  forces. 

Neglecting  orientation  effects,  then,  we  deal  only  with  the  centers  of 
gravity  of  the  molecules.  We  must  now  ask,  how  does  the  potential 
energy  depend  on  those*  centers  of  gravity?  We  have  seen  the  general 
nature  of  Van  dcr  Waals's  answer  to  this  question.  For  the  moment,  let 
us  simply  write  the  total  potential  energy  of  interaction  between  two 
molecules  i  and  j,  at  a  distance  rty  apart,  as  <t>(rlt).  Then  we  may  reason- 
ably assume  that  the  whole  potential  energy  of  the  gas  is 


(4.2) 

pairs  t,y 

We  now  adopt  Eq.  (5.22),  Chap.  Ill,  for  the  partition  function,  but 
remember  that,  as  in  Sec.  3,  Chap.  VIII,  we  must  multiply  by  1/Nl,  or 
approximately  by  (c/N)N,  in  order  to  take  account  of  the  identity  of 
molecules.  We  then  have 

*'•*•''•        (4-3) 

The  energy  E  is  like  that  of  Eq.  (3.1)  of  Chap.  VIII,  except  that  for 
simplicity  we  are  leaving  the  internal  part  of  the  energy  out  of  account, 
and  we  have  our  potential  energy  Eq.  (4.2).  The  integral  (4.3)  still 
factors  into  a  part  depending  on  the  momenta  and  another  on  the  coor- 
dinates, however,  and  the  part  depending  on  the  momenta  is  exactly  as 
with  a  perfect  gas  and  leads  to  the  same  result  found  in  Chap.  VIII. 


192                       INTRODUCTION  TO  CHEMICAL  PHYSICS  [CIIAF.  XII 

Thus  we  have 

'*»                   -5-  ,,  ., 

•  (4-4) 


*»r  f      f  -5-^    . 

J  J  Je       </?i 


The  integral  over  coordinates  is  the  one  that  simply  reduced  to  VN  in 
the  case  of  the  perfect  gas.  The  variables  dq\  .  .  .  can  be  written  more 
explicitly  as  dxL  dyi  dz\  .  .  .  dx»  dy\  dzN. 

Tin1  integration  over  the  coordinates  can  be  carried  out  in  steps. 
First,   we  integrate  over  the  coordinates  of  the  Nth  molecule.     The 


quantity  e   kT  can  be  factored;  it  is  equal  to 

~"kf    ~/7,r~Ar" 
c        e 


(4.5) 


where  2'  represents  all  those  pairs  that  do  not  include  the  Nth  molecule. 
The  first  factor  then  does  not  depend  on  the  coordinates  of  the  Nth 
molecule  and  may  be  taken  outside  the  integration  over  its  coordinates, 
leaving 

y,  0(f,Af) 

-*TT 
fffe    '          dxNdyNdzN.  (4.6) 

We  rewrite  this  as 

rfrrv  dyN  dz,  -  J//(l  -  c~~^)dxN  dyN  dz»  =  V  -  W,     (4.7) 


/// 

where  the  first  term  is  simply  the  volume,  the  second  an  integral  to  be 
evaluated,  which  vanishes  for  a  perfect  gas.  To  investigate  W,  imagine 
all  the  molecules  except  the  Nth  to  be  in  definite  positions.  If  the  gas  is 
rare,  the  chances  are  that  they  will  be  well  separated  from  each  other. 
Now  if  the  point  xNy^N  is  far  from  any  of  these  molecules,  the  interatomic 
potentials  <£(rtV)  will  all  be  small,  and  the  integrand  will  be  practically 
1  —  e°  =  0.  Thus  we  have  contributions  to  this  integral  only  from  the 
immediate  neighborhood  of  each  molecule.  Each  of  these  will  be  equal  to 


/  *fr'*)\ 

=  J/J\1  -  e     kT  )dx  dy  dz. 


w  =  J/J\1  -  e     kT     dx  dy  dz.  (4.8) 

For  simplicity  we  put  the  ith  molecule  at  the  origin  of  coordinates  and 
integrate  to  infinity  instead  of  just  through  the  container;  the  integrand 
becomes  small  so  rapidly  that  this  makes  no  difference  in  the  answer. 
Then  we  have  » 

/.  /  »(r)\ 

w  -   f   4rrr2Vl  -  e    kT  )dr.  (4.9) 

jo 

In  terms  of  this,  we  then  have 

W  -  (N  -  l)u>.  (4.10) 


SEC.  4]  VAN  DER  WAALS'  EQUATION  193 

Now  when  we  integrate  over  the  coordinates  of  the  (N  —  l)st  mole- 
cule, we  have  the  same  situation  over  again,  except  that  there  are  only 
(N  —  2)  remaining  molecules,  and  so  on.  Thus  finally  we  have  for  the 
integral  over  coordinates  in  Eq.  (4.4) 

[V  -  (N  -  l)w][V  -  (N  -  2)w]  •  •  •  V.  (4.11) 

To  evaluate  the  quantity  (4.11),  we  can  most  easily  take  its  logarithm; 
that  is, 

N-l 

In  (l  -  ™y  (4.12) 

8  =  0 

Replacing  the  sum  over  s  by  an  integral,  this  becomes 


f   In  (l  - 


NlnV 


Nw 

*T  1         T7  ""^    l 

=  AT  In  V  -  -  In    1  - 


(4-13) 


Our  assumptions  are  only  accurate  if  Nw/V  is  small;  for  it  is  only  in  this 
case  that  we  can  assume  that  all  molecules  are  well  separated  from  each 
other.  In  this  limit,  wo  ran  expand  tho  logarithm  as 


,    / 
111 


Nw\          Nw 
-- 


Substituting  in  Eq.  (4.13)  and  retaining  only  the  leading  term,  we 

Nln  V  -  l^N2~  -  -  •  (4.15) 

The  quantity  (4.15)  for  the  logarithm  of  the  integral  over  coordinates 
in  Eq.  (4.4)  can  now  be  substituted  in  the  expression  for  Helmholtz  free 
energy,  giving  at  once 

A  =  -WlnZ 

=  -^NkT  In  T  -  NkT  In  V  +  ^^~ 
z  zv 

-  NkT\  lu  (2^*1*  +  l  -  In  (Nk)  I-     (4.16) 

L          A3  J 

Equation  (4.16)  agrees  exactly  with  Eq.  (3.6),  Chap.  VIII,  except  for  the 
internal  partition  function  Z»,  which  we  are  here  neglecting  for  simplicity, 
and  for  the  extra  term  N2kTw/2V.  This  represents  the  effect  of  inter- 


194  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XII 

atomic  forces  and  is  characteristic  of  the  imperfect  gas.     Differentiating 
A  with  respect  to  volume,  we  at  once  have  for  the  equation  of  state 


p  —  —I 

^  ~      Wr 

=.  J™T_+.™I}?. 

V    +    2V*   ' 
or,  substituting  JVfc  =  nft,  AT  =  nJVo, 


Equation  (4.18)  is  in  the  form  of  Eq.  (4.1  )  and  shows  that  the  second  virial 
coefficient  is  given  by 


B(T)  =          ;  (4.19) 

where  w  is  given  by  Eq.  (4.9).  This  deduction  of  the  second  virial  coeffi- 
cient is  exact,  in  spite  of  the  approximations  wo  have  made;  if  further 
terms  are  retained,  they  prove  to  affect  only  the  third  and  higher  virial 
coefficients.  But  the  calculation  of  these  higher  coefficients  is  much 
harder  than  the  treatment  we  have  given  here. 

5.  The  Assumptions  of  Van  der  Waals'  Equation.  —  The  formula 
(4.19)  for  the  second  virial  coefficient,  together  with  Eq.  (4.9),  furnishes 
a  method  for  deriving  this  quantity  directly  from  any  assumed  inter- 
molecular  potential  function,  though  generally  the  integration  is  so  diffi- 
cult that  it  must  be  carried  out  numerically.  With  the  assumptions  of 
Van  der  Waals'  equation,  however,  the  problem  is  simplified  enough  so 
that  we  can  treat  Eq.  (4.9)  analytically  at  high  temperatures.  We 
assume  that  the  molecules  attract  each  other  with  a  force  increasing 
rapidly  as  the  distance  decreases,  so  long,  as  they  are  not  too  close  together. 
We  assume,  however,  that  the  molecules  act  like  rigid  spheres  of  diameter 
r0,  so  that  if  the  intermodular  distance  is  greater  than  r0  the  attraction 
is  felt,  but  if  the  distance  r  is  equal  to  r0  a  repulsion  sets  in,  which  becomes 

__  ± 

infinitely  great  if  the  distance  becomes  less  than  r0.  Then  e  kT  is  zero,  if 
r  is  less  than  ro,  so  that  Eq.  (4.9)  becomes 


fr°4irr2dr  +  f  °°47rr2(l  -  e~^)dr. 

»/0  «/ro 


w  =        4irr2dr  +        47rr2l  -  e~^dr.  (5.1) 

»/0  «/ro 

The  first  term  is  simply  -jTrrj!,  the  volume  of  a  sphere  of  radius  ro,  or  eight 
times  the  volume  of  the  sphere  of  diameter  ro  which  represents  a  molecule. 
In  the  second  integral,  we  may  expand  in  power  series,  since  <£  is  relatively 


SBC.  5]  VAN  DEE  WAALS'  EQUATION  195 

small.    The  bracket  is    *  ~       ~        '  '  '        =       '    Thus  the  term  is 


is    *  ~  ^  ~  tt  '  '  '  )\  =  pr' 
±  \ 


dr  +  •  •  • 
Then  for  the  second  virial  coefficient  we  have 

~     '•        (5-2) 


We  may  write  this 

B(T)  ^b-~> 

where 

,       #o4 


(5.3) 

Here  b  is  four  times  the  volume  of  No  spheres  of  radius  r*o/2,  or  four  times 
the  volume  of  all  the  molecules  in  a  gram  mole.  Since  the  force  repre- 
sented by  the  potential  <t>  is  attractive,  <£  is  negative  and  the  quantity  a 
is  positive  and  measures  the  strength  of  the  intermolecular  attractions. 
It  is  found  experimentally  that  the  formula  (5,3)  for  the  second  virial 
coefficient  is  fairly  well  obeyed  for  real  gases,  showing  that  the  assump- 
tions of  Van  der  Waals  are  not  grc'atly  in  error.  This  formula  leads  to  the 
equation  of  state 


Equation  (5.4)  indicates  that  for  high  temperatures  (where  a/RT  is  less 
than  6)  the  pressure  should  be  greater  than  that  calculated  for  a  perfect 
gas,  while  at  low  temperatures  (a/RT  greater  than  6)  the  pressure  should 
be  less  than  for  a  perfect  gas.  The  temperature 

r.  =  &  (5-5) 

at  which  the  second  virial  coefficient  is  zero,  so  that  Boyle's  law  is  satisfied 
exactly  as  far  as  terms  in  l/V  arc  concerned,  is  called  the  Boyle  tempera- 
ture. 

We  can  now  take  Van  der  Waals'  equation  (1.3),  expand  it  in  the  form 
of  Eq-  (4.1),  and  see  if  the  second  virial  coefficient  agrees  with  the  value 


196  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XII 

given  in  Eq.  (5.4).     We  have 

nRT  /n\2      nRT, 


n6 


n&V1         /nV 

-T)   -a(v) 

1  /"V 

"  \~a(v)' 


nRT 

agreeing  with  Eq.  (5.4)  as  far  as  the  second  virial  coefficient.  In  our 
theoretical  deduction,  we  have  not  found  the  third  virial  coefficient,  but 
this  can  be  done  by  a  good  deal  more  elaborate  methods  than  we  have 
used.  When  this  is  done,  it  is  found  that  it  does  not  agree  with  the  cor- 
responding quantity  in  Eq.  (5.6).  In  other  words,  Van  der  Waals' 
equation  is  correct  as  far  as  the  second  virial  coefficient  is  concerned  but 
no  further,  as  a  theoretical  equation  of  state  for  a  gas  whose  molecules  act 
on  each  other  according  to  Van  der  Waals'  assumptions. 

6.  The  Joule -Thomson  Effect  and  Deviations  from  the  Perfect  Gas 
Law. — The  deviations  from  the  perfect  gas  law  are  rather  hard  to  measure 
experimentally,  since  they  represent  small  fractions  of  the  total  pressure 
at  a  given  temperature  and  volume.  For  this  reason,  another  method  of 
detecting  the  departure  from  the  perfect  gas  law,  called  the  Joule-Thom- 
son effect,  is  of  a  good  deal  of  experimental  importance.  This  effect  is  a 
slight  variation  on  the  Joule  experiment.  That  experiment,  it  will  be 
recalled,  is  one  in  which  a  gas,  originally  confined  in  a  given  volume,  is 
allowed  to  expand  irreversibly  into  a  larger  evacuated  volume.  If  the 
gas  is  perfect,  the  final  temperature  of  the  expanded  gas  will  equal  the 
initial  temperature,  while  if  it  is  imperfect  there  will  be  slight  heating  or 
cooling.  This  experiment  is  almost  impossible  to  carry  out  accurately,  for 
during  the  expansion  there  are  irreversible  cooling  effects,  which  com- 
plicate the  process.  The  Joule-Thomson  effect  is  a  variation  of  the 
experiment  which  gives  a  continuous  effect,  and  a  steady  state. 

Gas  at  a  relatively  high  pressure  is  allowed  to  stream  through  some 
sort  of  throttling  valve  into  a  region  of  lower  pressure  in  a  continuous 
stream.  The  expansion  through  the  throttling  valve  is  irreversible,  as  in 
the  Joule  experiment,  and  the  gas  after  emerging  from  the  valvr  is  in  a 
state  of  turbulent  flow.  It  soon  comes  to  an  equilibrium  state  at  the 
lower  pressure,  however,  and  then  it  is  found  to  have  changed  its  tempera- 
ture slightly.  To  make  the  approach  to  equilibrium  as  rapid  as  possible, 
the  valve  is  usually  replaced  by  some  sort  of  porous  plug,  as  a  plug  of 
glass  wool,  which  removes  all  irregular  currents  from  the  gas  before  it. 
emerges.  Then  all  one  has  to  do  is  to  get  a  steady  flow  and  measure  the 
difference  of  pressure  and  the  difference  of  temperature,  on  the  two  sides 
of  the  plug.  If  AP  is  the  change  of  pressure,  AT7  the  change  of  tempera- 


SEC.  0]  VAN  DER  WAALS'  EQUATION  197 

ture,  on  passing  through  the  plug,  the  Joule-Thomson  coefficient  is  defined 
to  be  AjT/AP.  It  is  zero  for  a  perfect  gas  and  can  be  either  positive  or 
negative  for  a  real  gas.  Wo  shall  now  evaluate  the  Joule-Thomson 
coefficient  in  terms  of  the  equation  of  state. 

It  is  easy  to  show  that  the  enthalpy  of  unit  mass  of  gas  is  unchanged 
as  it  flows  through  the  plug.  Let  a  volume  V\  of  gas  be  pushed  into  the 
pipe  at  pressure  PI]  then,  since  P\  is  constant  through  this  pipe,  work 
/Pi  (IV  i  —  P\V\  is  done  on  this  sample  of  gas.  After  passing  through 
the  plug,  the  same  mass  has  a  volume  F2,  and  does  work  P2F2  in  passing 
out  of  the  pipe.  Thus  the  external  work  done  by  the  gas  in  the  process  is 
PzVz  —  PiTi-  It  is  assumed  that  no  heat  is  absorbed,  so  that  if  f/i  is  the 
internal  energy  when  the  gas  enters,  C72  when  it  leaves,  the  first  law  gives 


or 

Ui  +  P,Fi  =  lJ\  +  P2K2,         7/t  =  /72.  (6.1) 

Thus  the  change  is  at  constant  #,  and  the  Joule-Thomson  coefficient  is 
(3T/dP)n.  But  this  can  be  evaluated  easily  from  our  Table  of  Thermo- 
dynamic  Relations  in  Chap.  II.  It  is 


(dT\    = 
WA 


(dg\ 

\dTjr 


71  |D    -  V 


(6.2) 

Vf 

From  Eq.  (6.2),  we  sec  that  for  a  perfect  gas,  for  which  V  is  proportional 
to  T  at  constant  P,  the  Joule-Thomson  coefficient  is  zero.  For  an  imper- 
fect gas,  we  assume  the  equation  of  state  (5.4).  We  have 


ev 


(dP\ 
WA 


T7         ,       2na 
V  ~  nb  +  ~ 


198  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XII 

where  we  have  regarded  (b—  nm)(  17)  &s  a  small  quantity  compared 
with  unity,  neglecting  its  square.     Substituting,  we  then  have 

dT\          n  2a 


From  Eq.  (6.4),  we  see  that  the  Joule-Thomson  coefficient  gives 
immediate  information  about  a  and  b.  If  we  measure  the  coefficient  and 
know  CP,  so  that  we  can  calculate  the  quantity  (Cp/n)(dT/dP)n,  we  can 
plot  the  resulting  function  as  a  function  of  l/T  and  should  get  a  straight 
line,  with  intercept  —  6,  and  slope  2a/R,  so  that  both  b  and  a  can  be 
found  from  measurements  of  the  Joule-Thomson  effect  as  a  function  of 
temperature.  We  notice  that  at  high  temperatures  the  coefficient  is 
negative,  at  low  temperatures  positive.  That  is,  since  AP  is  negative  in 
the  experiment,  corresponding  to  a  decrease  of  pressure,  the  change 
of  temperature  is  positive  at  high  temperatures,  loading  to  a  heating  of 
the  gas,  while  it  is  negative  at  low  temperatures,  cooling  the  gas.  The 
temperature  2a/Rb,  whore  the  effect  is  zero,  is  called  the  temperature  of 
inversion;  we  see  by  comparison  with  Eq.  (5.5)  that,  if  our  simple  assump- 
tions are  correct,  this  should  be  twice  the  Boyle  temperature.  The  Joule- 
Thomson  effect  is  used  practically  in  the  Lindc  process  for  the  liquefaction 
of  gases.  In  this  process,  the  gas  is  first  cooled  by  some  method  below 
the  temperature  of  inversion  and  then  is  allowed  to  expand  through  a 
throttling  valve.  The  Joule-Thomson  effect  cools  it  further,  and  by  a 
repetition  of  tho  process  it  can  be  cooled  enough  to  liquefy  it. 


CHAPTER  XIII 
THE  EQUATION  OF  STATE  OF  SOLIDS 

Next  to  perfect  gases,  regular  crystalline  solids  are  the  simplest  form 
of  matter  to  understand,  being  less  complicated  than  imperfect  gases  near 
the  critical  point,  or  liquids.  Unlike  perfect  gases,  there  is  no  simple 
analytic  equation  of  state  which  always  holds;  we  are  forced  either  to  use 
tables  of  values  or  graphs  to  represent  the  equation  of  state,  or  to  expand 
in  power  scries.  But  the  theory  is  far  enough  advanced  so  that  we  can 
understand  the  simpler  solids  fairly  completely.  As  with  gases,  we  shall 
start  our  discussion  from  a  thermodynamic  standpoint,  asking  how  one 
can  find  information  from  experiment,  and  then  later  shall  go  on  to  the 
theory,  seeing  how  far  one  can  go  by  statistical  mechanics  in  setting  up  a 
model  of  a  solid  and  predicting  its  properties.  Of  course,  it  is  obvious 
that  in  one  respect  the  subject  of  solids  is  a  much  wider  one  than  that  of 
gases:  there  is  tremendous  variety  among  solids,  whereas  all  gases  act 
very  much  alike.  This  comes  from  the  different  types  of  forces  holding 
the  atoms  together  and  the  different  crystal  structures.  We  shall  put  off 
most  of  the  discussion  of  the  different  types  of  solids  until  later  in  the 
book,  when  we  take  up  chemical  substances  and  their  properties.  When 
we  come  to  that,  we  shall  see  to  what  a  large  extent  the  fundamental 
atomic  arid  molecular  properties  of  a  solid  are  brought  out  in  the  behavior 
of  its  solid  state. 

1.  Equation  of  State  and  Specific  Heat  of  Solids. — To  know  the 
equation  of  state  of  a  solid,  we  should  have  its  pressure  as  a  function  of 
volume  and  temperature.  Really  we  should  know  more  than  this:  a 
solid  can  support  a  more  complicated  stress  than  a  pressure,  and  can  have 
a  more  complicated  strain  than  a  mere  change  of  volume.  Thus  for 
instance  it  can  be  sheared.  And  in  g£n_eraljthp  "equation  of  state "  kjL 
set  nfjvjfttfions  giving  the  stress  at  every  point  of  the  solid^as  a  jujicliaiL 
of  the^ strains  and  the  temperature.  TBut  we  shall"  not  concern  ourselves 
with  these  general  stresses  "a'mPsTraihs,  though  they  are  of  great  impor- 
tance both  practically  and  theoretically;  we  limit  ourselves  instead  to  the 
case  of  hydrostatic  pressure,  in  which  the  volume  and  temperature  are 
adequate  independent  variables.  Let  us  consider  what  we  find  from 
experiment  on  the  compression  of  solids  to  high  pressures.  At  zero  pres- 
sure, the  volume  of  a  solid  is  finite,  unlike  a  gas,  and  it  changes  with 
temperature,  generally  increasing  as  the  temperature  increases,  as  given 
by  the  thermal  expansion.  As  the  pressure  is  increased  at  a  given  tern- 

199 


200 


INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 


perature,  the  volume  decreases,  as  given  by  the  compressibility.  Com- 
bining these  pieces  of  information,  we  have  a  set  of  curves  of  constant 
temperature,  or  isothormals,  as  given  in  Fig.  XIII-1.  These  are  plainly 
very  different  from  the  isothormals  of  a  perfect  gas,  which  are  hyperbolas, 
the  pressure  being  inversely  proportional  to  tho  volume.  If  we  know 
nothing  experimentally  but  tho  thermal  expansion  and  the  compressi- 
bility, we  should  have  to  draw  the  lines  as  straight  lines,  with  equal  spac- 
ing for  equal  temperature  changes.  Fortunately  the  measurements  are 
more  extensive.  The  pressure  is  known  as  a  function  of  volume  over  a 
wide  pressure  rango,  enough  in  most  solids  to  change  tho  volumo  by  a 
per  cent,  and  with  very  compressible  solids  by  many  per  cont,  and  the 
volume  is  known  as  a  function  of  tomporaturo  for  wido  ranges  of  tompera- 
turo.  Tho  curves  must  stop  experimentally  at  zoro  pressure1,  but  wo  can 


4QOOO 


30,000 


20,000 


10,000 


i     i     i     I 


275°obs 


05  x<: 

V/V0 

FIG.  XIII-1. — Isothermals  for  a  solid  (sodium)  giving  pressure  as  a  function  of  volumo  at 

constant  temperature. 

imagine  that  they  could  bo  extrapolated  to  negative  pressures,  as  indi- 
cated by  the  dotted  lines  in  the  figure. 

To  carry  out  any  calculations  with  the  equation  of  state,  we  wish  to 
approximate  it  in  some  analytic  way.  First,  lot  us  consider  the  most 
convenient  variables  to  use.  The  results  of  experiment  are  usually 
expressed  by  giving  the  volume  as  a  function  oTpressurcltnd  temperature. 
Thus  thelhermal  expansion  Is" investigated  as  a  function  of  temperature 
at  atmospheric  pressure,  and  in  measurements  of  compressibility  the 
volume  is  found  as  a  function  of  pressure  at  certain  fixed  temperatures. 
On  the  other  hand,  for  deriving  results  from  statistical  mechanics,  it  is 
convenient  to  find  the  Helmholtz  free  energy,  and  hence  the  pressure,  as  a 
function  of  volume  and  temperature.  We  shall  express  the  equation  of 
state  in  both  forms,  and  shall  find  the  relation  between  the  two.  We  let 
F0  be  the  volume  of  our  solid  at  no  pressure  and  at  the  absolute  zero  of 
temperature.  Then  we  shall  assume 


V  =  Fn(l 


-  a,(7')P 


(1.1) 


SEC.  1]  THE  EQUATION  OF  STATE  OF  SOLIDS  201 

where  a0,  ai,  a2,  etc.,  are  functions  of  temperature,  the  signs  being  chosen 
so  that  they  are  positive  for  normal  materials.  Tho  meaning  of  the  q's 
is  easily  found.  Thus,  first  at  zero  pressure  (which  for  practical  purposes 
is  identical  with  atmospheric  pressure,  since  the  volume  of  a  solid  changes 
so  slowly  with  pressure)  the  volume  is  F0[l  +  ao(T)].  __  The  coefficient 
of  thermal  expansion  at  zero  pressure  is  then 


_  l(dV\ 
a  -  V\dTj 


I  1          dtto  rftto  •  ,1  /*    rk\ 

'A  "  TT^df  -  dT  approximately.  (1.2) 


If  the  material  has  a  constant  thermal  expansion,  so  that  the  change  in 
volume  is  proportional  to  temperature,  we  should  have  approximately 
dao/dT  =  a,  where  a  is  constant,  leading  to  a0(71)  =  aT.  This  is  a 
special  case,  however;  it  is  found  that  for  real  materials  the  coefficient 
of  thermal  expansion  becomes  smaller  at  low  temperatures,  approaching 
zero  at  the  absolute  zero;  for  this  reason  we  prefer  to  leave  a0(T)  as  an 
undetermined  function  of  the  temperature,  remembering  only  that  it, 
reduces  to  zero  at  the  absolute  zero  (by  the  definition  of  F0),  and  that  it 
is  very  small  compared  to  unity,  since  the  temperature  expansion  of  a 
solid  is  only  a  small  fraction  of  its  whole  volume. 

The  meaning  of  ai  is  simple:  rt  is  almost  exactly  equal  jto^jhc  com- 
pimsibilit/^^  The  compressibility  x  is  ordinarily  defined 

as  —  (\/V)(dV/dP)r,  to  be  computed  at  zero  pressure.  From  Kq.  (1.1), 
remembering  that  the  volume  at  zero  pressure  is  given  by  rn[l  +  a0(77)], 
we  have 


where  in  the  last  form  we  have  again  neglected  ao  compared  to  unity. 
The  compressibility  ordinarily  increases  with  increasing  temperature,  so 
that  a\(T)  must  increase  with  temperature,  enough  to  produce  a  net 
mcrease'ih  spite  of  the  increase  of  the  factor  1  +  ao  in  the  denominator  of 
Eq.  (1.3).  The  increase  is  not  very  great,  however;  most  compressi- 
bilities do  not  change  by  more  than  10  per  cent  or  so  between  absolute 
zera  and  hifih  temperatures.  I'h6  flUfthtity  a*  measures  essentially  the 
change  of  comgrejsgibility  with  pressure  .^JLtittkJs^  known  experimentally 
about  its  temperature  variation,  though  it  presumably  increases  with 
temperature  in  some  such  way  as  ai  does.  The  terms  of  the  series  in 
P  written  down  in  Eq.  (1.1)  represent  all  that  are  required  for  most 
materials  and  the  available  pressure  range.  Most  measurements  of 
solids  at  high  pressures  have  been  carried  out  by  Bridgman,1  who  has 
measured  changes  of  volume  up  to  pressures  of  12,000  atm.  with  many 

lSee  P.  W.  Bridgman,  "Physics  of  High  Pressures,"  Chap.  VI,  The  Macmillan 
Company,  1931,  and  later  papers. 


202  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 

solids  and  to  45,000  atm.  with  a  few  solids.  At  these  highest  pressures, 
the  most  compressible  solid,  Cs,  caesium,  has  its  volume  reduced  to  less 
than  half  the  volume  at  atmospheric  pressure,  and  the  other  alkali  metals, 
Li,  lithium,  Na,  sodium,  K,  potassium,  and  Rb,  rubidium,  have  reductions 
in  volume  of  from  20  to  50  per  cent.  To  represent  these  large  changes  of 
volume  accurately  requires  a  considerable  number  of  terms  of  such  a 
series  as  (1).  Those  are  extreme  cases,  however;  most  solids  are  much 
less  compressible,  and  changes  of  volume  of  only  a  few  per  cent  can  be 
produced  with  the  available  pressure,  so  that  we  can  approximate  quite 
accurately  by  a  quadratic  function  of  pressure,  as  in  Eq.  (1.1).  The 
experimental  results  are  usually  stated  by  giving  the  relative  change  of 
volume  as  a  power  scries  in  the  pressure.  That  is,  in  our  notation,  we 
have 


70(1  +  qQ)  —  V  =     aiP  ___  az    p2  n  4N 

^  F0(l  +  a0)  1  +  a0       1  +  a0     '  V      ; 

I 

The  constants  ai/(l  +  a0)  and  «2/(l  +  «o)  are  given  as  the  result  of 
experiments  on  compressibility.  If  a0  is  known  from  measurements  of 
thermal  expansion,  we  can  then  find  a\  and  a2  directly  from  experiment. 
The  equation  of  state  (1.1)  is  expressed  in  terms  of  pressure  and 
temperature  as  independent  variables.  We  shall  next  express  it  in  terms 
of  volume  and  temperature.  We  shall  do  this  in  the  form 


P  =  Pn(T)  +  P}(T)-       +  P2(T)-         •  •  •      (1.5) 

Here  Po(T),  P\(T),  and  Pz(T)  are  functions  of  temperature,  again  chosen 
to  be  positive.  The  meaning  of  PO  is  simple:  it  is  the  pressure  that  must 
be  applied  to  the  solid  to  reduce  its  volume  to  F0,  the  volume  which  it- 
would  have  at  the  absolute  zero  under  no  pressure.  Obviously  P0  goes 
to  zero  at  the  absolute  zero.  At  ordinary  temperatures,  while  it  repre- 
sents a  very  considerable  pressure,  still  it  is  small  compared  to  the  quan- 
tities Pi  and  Pz,  so  that  it  can  be  treated  as  a  small  quantity  in  our 
calculations  and  its  square  can  be  neglected.  We  shall  see  in  a  moment 
that  PI  is  approximately  the  reciprocal  of  the  compressibility,  or  equals 
the  pressure  required  to  reduce  the  volume  to  zero,  if  the  volume  decreased 
linearly  with  increasing  pressure  (which  of  course  it  does  not).  Obvi- 
ously this  is  much  greater  than  the  pressure  required  to  reduce  the  volume 
to  Fo. 

We  shall  now  find  the  relations  between  the  a's  of  Eq.  (1.1)  and  the 
quantities  P0,  PI,  Pz  of  Eq.  (1.5),  assuming  that  we  can  neglect  the  squares 
and  higher  powers  of  a0  and  P0.  To  do  this,  we  write  Eq.  (1.1)  in  the  form 

~  •  •  •  ,  (1.6) 


SBC.  1]  THE  EQUATION  Of  STATE  OF  SOLIDS  203 

substitute  in  Eq.  (1.5),  and  equate  the  coefficients  of  different  powers  of 
P.     We  have 


P  -  Po  +  Pi(-a0  + 

ofP2  •  •  •  ),     (1.7) 


where  we  have  neglected  a%.     Equating  coefficients,  we  have  the  equations 
0  =  Po  - 


-  2P2a0a, 
0  =  -P,a2  +  2P2aoa2  +  P^al  (1.8) 


Solving  for  the  a's,  AVO  have 

„        Po 
a«=p- 


-  2P2a»  ~  P 


P? 

Similarly  solving  for  the  P's  we  have 


Pi  -  - 


1      «?   / 
at 


(1.10) 

'1 


Since  we  know  how  to  find  the  a's  from  experiment,  Eqs.  (1.10)  tell  us 
how  to  find  the  P's.  We  observe  from  Eqs.  (1.10)  that,  as  mentioned 
before,  PI  is  equal,  apart  from  small  terms  proportional  to  a0,  to  the 
reciprocal  of  the  compressibility  given  in  Eq.  (1.3). 

In  addition  to  the  equation  of  state,  we  must  find  the  specific  heat 
from  experiment.  Ordinarily  ono  finds  the  specific  heat  at  constant 
pressure,  CP)  at  atmospheric  pressure,  or  practically  at  zero  pressure.  We 
shall  call  this  C£,  to  distinguish  it  from  the  general  value  of  CPj  which 
can  depencHm  pressure.  Le^usjSndJjic .  jgp^dence^pn  pressure.  From 
Eq.  (1.6),  Chap.  VIII,  we  have  ~  ~~  ~ 


>*P/T 

Substituting  for  V  from  Eq.  (1.1)  and  integrating  with  respect  to  pressure 


204  INTRODUCTION  TO  CHEMICAL  PHYSICS         [Cinr.  XIII 

from  P  —  0  to  P,  we  have 


-  v  r*p  -        ip*  +        *p*  •  •  -         n  in 

VoTP  ^P   +^P  (LH) 


In  case  an,  a\t  and  «2  can  be  approximated  by  linear  functions  of  tempera- 
ture, as  we  considered  earlier  for  a0,  the  second  derivatives  in  Eq.  (L.ll) 
will  be  zero  and  CP  will  be  independent  of  pressure.  Since  da^/dT  is 
essentially  the  coefficient  of  thermal  expansion,  we  see  that  the  term  in 
Eq.  (l.ll)  linear  in  the  pressure  depends  on  the  change  of  thermal  expan- 
sion with  the  temperature.  We  have  mentioned  that  the  thermal 
expansion  is  zero  at  the  absolute  zero,  increasing  with  temperature  to  an 
asymptotic  value.  Thus  we  may  expect  d2aQ/dT2  to  be  positive,  falling 
off  to  zero  at  high  temperatures,  so  that  from  Eq.  (1.11)  the  specific  heat 
will  decrease  with  increasing  pressure,  particularly  at  low  temperature. 
For  theoretical  purposes,  it  is  better  to  use  the  specific  heat  at  constant 
volume,  CV,  computed  for  the  volume  V0  which  the  solid  has  at  zero  pres- 
sure and  temperature.  We  shall  call  this  C°.  CV  will  depend  on  the 
volume  as  indicated  by  Eq.  (1.7)  of  Chap.  VIII: 


Using  Eq.  (1.5)  for  the  pressure,  we  obtain 


2 

1      •  •     (1.12) 


1   3  dT 

From  Eq.  (1.9),  P0  is  proportional  to  a0,  so  that  its  second  derivative  will 
likewise;  be  positive,  and  we  find  that  CV  will  decrease  with  decreasing 
volume  or  increasing  pressure,  just  as  we  found  for  Cp. 

Since  it  is  impracticable  to  find  CV,  or  Cy,  from  direct  experiment,  it  is 
important  to  be  able  to  find  these  quantities  from  CP.  From  Eq.  (5.2), 
Chap.  II,  we  know  how  to  find  CP  —  CV:  it  is  given  by  the  formula 
T(dV/dT)P(dP/dT)v.  This  gives  the  difference  of  specific  heats  at  a 
given  pressure  and  temperature.  We  are  more  interested,  however,  in 
the  difference  Cp  —  Cy,  in  which  CP  is  computed  at  zero  pressure,  Cy  at 
the  volume  Fo.  To  find  this  difference,  let  us  carry  out  a  calculation  of 
CV  at  zero  pressure,  from  Eq.  (1.12).  Here  we  have  Fo  —  F  =  —  F0a0, 

from  Eq.  (1.1).  Then  Eq.  (1.12)  gives  us  CV  =  C°v  +  VQTad~~  Using 
this  value  and  the;  equation  for  CP  —  CV,  which  we  calculate  for  zero 


SBC.  2]  THE  EQUATION  Ob'  STATK  OF  SOLIDS  205 

pressure,  we  have 

fK\       fin        v  Tn  ^2^°  -  v  Tda0dPQ 
Op  —  or  —  Kn/a<r2  —      ol        ~' 


,    ., 

(L13) 

In  tho  derivation  of  Eq.  (1.13),  we  have  neglected  the  variation  of  a^ 
with  temperature.  In  case  the  thermal  expansion  is  constant,  so  that 
ao  =  ctT,  and  the  specific  heat  is  independent  of  volume  or  pressure,  Eq. 
(1.13)  takes  the  simple  form 

° 


where  we  remember  that  a  is  Hie  coefficient  of  thermal  expansion,  ai  the 
compressibility,  to  a  good  approximation.  When  numerical  values  are 
substituted  in  Eq.  (1.14),  it  is  found  that  the  difference  of  specific  heats 
for  a  solid  is  much  less  than  for  a  gas,  so  that  no  great  error  is  committed 
if  we  use  one  in  place  of  the  other.  This  can  be  seen  from  the  fact  that 
the  difference  of  specific  heats  depends  on  aji,  as  we  see  in  Eq.  (1.13), 
whereas  elsewhere  we  have  considered  «n  as  being  so  small  that  its  square- 
could  be  neglected. 

We  have  now  discussed  all  features  of  the  specific  heats,  except  for 
the  dependence  of  C£  or  CfJ  themselves  on  temperature.  Experimentally 
it  is  found  that  the  specific  heat  is  zero  at  the  absolute  zero  and  rises  to  an 
asymptotic  valiuTat  high  temperatures,  much  like  the  specific  heat  of  an 
oscillator,  as  shown  in  Fig.  IX-a.  We  shall  see  later  that  the  t.frqrmal 
energy  oi  a  solid  comes  trom  iheTscillations  of  the  molecules,  so  that  there 
is  a  fundamental  reason  tor  this  behavior  of  the  specific  heat.  We  shall 
also  find  theoretical  formulas  later  which  express  the  specific  heat  with 
fairly  good  accuracy  as  a  function  of  the  temperature,  formulas  that  differ 
in  some  essential  respects  from  Eq.  (5.8),  Chap.  IX,  from  which  Fig.  IX-3 
was  drawn.  For  the  present,  however,  where  we  are  discussing  thermo- 
dynamics, we  must  simply  assume  that  the  specific  heat  is  given  by 
experiment  and  shall  treat  C?  and  Cv  as  unknown  functions  of  the  tem- 
perature, which  however  always  reduce  to  zero  at  the  absolute  zero  of 
temperature. 

2.  Thermodynamic  Functions  for  Solids.  —  In  the  preceding  section 
we  have  seen  how  to  express  the  equation  of  state  and  specific  heat  of  a 
solid  as  functions  of  pressure,  or  volume,  and  temperature.  Now  we  shall 
investigate  the  other  thermodynamic  functions,  the  internal  energy, 
entropy,  Helmholtz  free  energy,  and  Gibbs  free  energy.  For  the  internal 


206 


INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 


energy  as  a  function  of  volume  and  temperature,  we  have  the  relations 
(dU/dT)v  =  Cv,  (dU/dV)T  =  T(dP/dT)v  -  P.  Let  the  energy  of  the 
solid  at  volume  Fo  and  zero  temperature  be  C/oo.  Then  we  find  the 
energy  at  any  temperature  and  volume  by  starting  at  VQ  at  the  absolute 
zero,  raising  the  temperature  at  volume  VQ  to  the  desired  temperature, 
and  then  changing  the  volume  at  this  temperature.  Using  Eq.  (1.5),  we 
find  at  once 


U  -  t/oo  + 


(2.1) 


The  internal  energy  of  metallic  sodium  is  shown  as  a  function  of  volume  in 
Fig.  XIII-2,  as  an  illustration.  On  account  of  the  large  compression 
that  can  be  attained  with  sodium,  more  terms  of  the  power  series  must 
be  retained  than  are  given  in  Eq.  (2.1),  but  it  is  easier  to  show  the  prop- 
erties of  this  metal  than  of  a  less  compressible  one.  Lot  us  consider  the 
behavior  of  the  internal  energy  as  a  function  of  volume  at  fixod  tempera- 
ture. If  the  thermal  expansion  is  independent  of  temperature,  so  that  P0 

is  proportional  to  the  temperature  and 
dPQ/dT  is  a  constant,  the  coefficient  of 
the  term  in  (F0  -  V)  in  Eq.  (2.1)  is 
zero  and  the  term  in  (Vo  —  V)'2  is  the 
principal  term  in  U.  In  this  term,  PI, 
which  is  the  reciprocal  of  the  compressi- 
bility, is  large  compared  to  T  dPi/dT,  so 
that  the  coefficient  of  (F0  —  F)2  is  posi- 
tive and  the  internal  energy  has  a  mini- 
mum at  Fo,  just  as  it  does  at  the  absolute 
zero.  If  the  thermal  expansion  depends 


,.5 


1.0 


0.5 


08 


0.9 


10 
V/V0 


for  various  temperatures.    The  dotted  ordinarily  to   smaller   volumes. 

line  connects  points  at  zero  pressure.      .  . 

is    an    interesting    consequence 


12  on  temperature,  the  term  in  (F0  —  V) 
will  have  a  small   coefficient  different 

FIG.  XIII-2.  —  Internal  energy  of  a  .    .  1.1,1 

solid  (sodium)  as  function  of  volume  from  zero,  shifting  the  minimum  slightly, 

There 
of   the 

fact  that  the  minimum  of  U  is  approximately  at  F0.  At  ordinary 
temperatures,  the  volume  of  the  solid  at  zero  pressure,  which  as  we  have 
seen  is  Fo(l  +  «o),  will  be  greater  than  Vo.  Then  on  compressing  the 
solid,  its  internal  energy  will  decrease  until  we  have  reduced  its  volume 
approximately  to  Fo,  when  it  will  begin  to  increase  again.  Of  course, 
work  is  constantly  being  done  on  the  solid  during  the  compression,  but 
so  much  heat  flows  out  to  maintain  the  temperature  constant  that  the 
total  energy  decreases,  with  moderate  compressions.  The  internal 


SBC.  2]  THE  EQUATION  OF  STATE  OF  SOLIDS  207 

energy  of  course  increases  as  the  temperature  is  raised  at  constant  volume, 
as  we  see  from  the  obvious  relation  (dU/dT)v  =  CV,  so  that  the  curves 
corresponding  to  high  temperatures  lie  above  those  for  low  temperature. 
Furthermore,  since  the  specific  heat  is  higher  at  large  volume,  as  we  saw 
from  Eq.  (1.12),  the  spacing  of  the  curves  is  greater  at  large  volume, 
resulting  in  the  slight  shift  of  the  minimum  to  smaller  volume  with 
increasing  temperature. 

The  entropy  is  most  easily  determined  as  a  function  of  volume  and 
temperature  from  the  equation  (dS/dT)v  =  CV/T.  At  the  absolute  zero 
of  temperature,  the  entropy  of  a  solid  is  zero  independent  of  its  volume  or 
pressure.  The  reason  goes  back  to  our  fundamental  definition  of  entropy 

in  Chap.  Ill,  8  —  ~^2/t  In  /t>  where/,-  represents  the  fraction  of  all 


systems  of  the  assembly  in  the  ith  state.  At  the  absolute  zero,  according 
to  the  canonical  assembly,  all  the  systems  will  be  in  the  state  of  lowest 
energy,  which  will  then  have/  =  1,  all  other  states  having/  =  0.  Thus 
automatically  S  =  0.  We  can  then  find  the  entropy  at  any  temperature 
and  volume  as  follows.  First,  at  absolute  zero,  wo  change  the  volume* 
to  the  required  value,  with  no  change  of  entropy.  Then,  at  this  constant 
volume,  we  raise  the  temperature,  computing  the  chango  of  entropy  from 
JCv/TdT.  We  can  use  Eq.  (1.12)  for  the  specific  heat  at  arbitrary 
volume.  Carrying  out  the  integration  from  that  equation,  we  have  at 
once 


_  fVr-  v\^(Y^L 

Jo     *T       y°[dT\     V, 


YVo  -2  ,  «jL- 

+  2'dT\     V,          ^'  "  (      } 

The  entropy  of  sodium,  as  computed  from  Eq.  (2.2),  is  plotted  in  Fig. 
XIII-3  as  a  function  of  temperature,  for  several  volumes.  Starting  from 
zero  at  the  absolute  zero,  the  entropy  first  rises  slowly,  since  its  slope, 
Cv/Tj  goes  strongly  to  zero  at  the  absolute  zero.  As  tho  temperature 
rises,  the  curve  goes  over  into  something  more  like  the  logarithmic  form 
which  it  must  have  at  high  temperature,  where  CV  becomes  constant,  and 
S  =  JCV  dT/T  =  CV  In  T  +  const.  From  the  curves,  it  is  plain  that 
the  entropy  increases  with  increasing  volume,  at  constant  temperature. 
This  can  be  seen  from  Eq.  (2.2),  in  which  the  leading  term  in  the  variation 

7jp 

with  volume  can  be  written  -j7ft(V  —  F0),  where  from  Eq.  (1.10)  we  see 

»p 

that  -j~  is  approximately  the  thermal  expansion  divided  by  the  com- 

pressibility.    It  can  also  be  seen  from  the  thermodynamic  relation 


208 


INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIII 


dV 


which  is  seen,  if  we  multiply  and  divide  by  1/F,  to  be  exactly  the  thermal 
expansion  divided  by  the  compressibility.  The  reason  for  the  increase  of 
entropy  with  increasing  volume  is  simple:  if  the  volume  increased,  or  the 
pressure  decreased,  adiabatically,  the  material  would  cool;  to  keep  the 
temperature  constant  heat  must  flow  in,  increasing  the  entropy. 

The  Helmholtz  free  energy  A  =  U  — •  TS  can  be  found  from  Kqs. 
(2.1)  and  (2.2)  for  U  and  S,  or  can  be  found  by  integration  of  the  equations 


0  100  200 

Deg.  Abs. 

Flu.  XIII-3.  —  Entropy  of  a  solid   (sodium)   as  function  of  the  temperature  tit  constant 

volume. 

(dA/dT)v  =  —S,  (dA/dV)T  =  -P.  The  latter  method  is  perhaps  a 
little  more  convenient.  At  the  absolute  zero  and  volume  To,  the  Helm- 
holtz free  energy  equals  the  internal  energy  and  is  given  by  f/oo,  as  in  Eq. 
(2.1).  From  that  point  we  increase  the  temperature  at  volume  VQ  to 
the  desired  temperature,  and  then  change  the  volume  at  this  temperature. 
We  find  at  once 


A  = 


In  Fig.  XIII-4  we  show  A  as  a  function  of  volume  for  a  number  of  tem- 
peratures. At  the  absolute  zero,  as  we  have  mentioned  above,  the 
Helmholtz  free  energy  equals  the  internal  energy,  as  given  in  Fig.  XIII-2. 


SEC.  2] 


THE  EQUATION  OF  STATE  OP  SOLIDS 


209 


From  the  equation  (dA/dV)T  =  — P,  we  see  that  the  negative  slope  of  the 
Helmholtz  free  energy  curve  is  the  pressure,  and  the  change  of  Helmholtz 
free  energy  between  two  volumes  at  constant  temperature  gives  the 
external  work  done  in  changing  the  volume.  It  is  for  this  reason,  of 
course,  that  it  is  called  the  free  energy.  Thus  the  minimum  of  each  curve 
corresponds  to  the  volume  where  the  pressure  is  zero.  It  is  obvious  from 
the  graph  that  this  minimum  moves  outward  to  larger  volumes  with 
increase  of  temperature;  this  represents 
the  thermal  expansion.  In  particular, 
it  is  plain  that  this  shift  of  the  mini- 
mum is  very  small  for  low  temperatures, 
corresponding  to  the  small  thermal 
expansion  at  low  temperatures.  Since 
the  slope  of  the  free  energy  curve  gives 
t  he  negative  pressure,  it  is  only  the  part 
of  the  curve  to  the  left  of  the  minimum 
that  corresponds  to  positive  pressure  _£• 
and  has  physical  significance. 

Finally,  we  consider  the  Gibbs  free 
energy, 

G  =  U  +PV  -  TS  =  A  +  Plr, 

as  a  function  of  pressure  and  temper- 
ature. This  is  most  conveniently  found 
from  the  relations  (dU/dP)? •=  T, 
(dG/dT)?=-S.  Starting  at  the  abao-  „  VTTT  ^  U1  .  14  , 

'         '  **  FIG.  XIII-4. — Helmholtz  free  energy 

lute  zero  and  zero  pressure,  where  the  of  a  solid  (sodium)  as  function  of  tho 
value  of  G  is  Z70o,  we  first  increase  the  volume  at  c»n«*ant  temperature, 
temperature  at  zero  pressure,  then  increase  the  pressure  at  constant 
temperature,  finding 

•r      j 


-  f( 


PV.(l 


a,)  - 


(2.4) 


111  Fig.  XIII-5,  we  plot  G  as  a  function  of  pressure,  for  a  number  of  tem- 
peratures. The  term  PVo(l  +  a0)  is  by  far  the  largest  one  in  (7,  resulting 
approximately  in  straight  lines  proportional  to  P.  The  spacing  of  the 
curves  is  determined  by  the  entropy:  (dG/dT)P  =  —  S,  showing  that  G 
decreases  with  increasing  temperature  at  constant  pressure  and  that  the 
decrease  is  greater  at  low  pressure  (large  volume)  than  at  high  pressure. 
These  details  of  the  change  of  the  Gibbs  free  energy  with  temperature 
are  not  well  shown  in  Fig.  XIII-5,  however,  on  account  of  scale,  and 


210 


I  \TRODUCTWN  TO  CHEMICAL  I'HYMC'ti          [CHAP.  Xii( 


this  sort  of  plot  does  not  give  a  great  deal-of  useful  information.  Before 
leaving  it,  it  is  worth  while  pointing  out  the  resemblance  to  Fig.  XII-4, 
where  we  plotted  G  as  a  function  of  pressure  for  a  liquid  and  gas  in  equilib- 
rium, as  given  by  Van  der  Waals'  equation,  and  found  again  almost 
straight  lines. 

The  more  useful  way  to  give  G  graphically  is  to  plot  it  as  a  function  of 
temperature  for  constant  pressure,  as  we  do  in  Fig.  XIII-6.  The  slope  of 
these  curves,  being  —5,  is  zero  at  the  absolute  zero,  negative  at  all  higher 
temperatures,  so  that  the  curves  slope  down.  The  Gibbs  free  energy 
decreases  more  slowly  with  temperature  at  high  pressure,  where  the 
entropy  is  lower,  than  at  zero  pressure.  At  zero  pressure,  the  term  PV 


0 


10,000  2QOOO 

P  Atmospheres 


FIG.  XIII-5. — Gibbs  free  energy  of  a  solid  (sodium)  as  function  of  the  pressure  at  constant 
J  temperature. 

is  zero,  so  that  the  Gibbs  free  energy  G  equals  the  Helmholtz  free  energy 
A.  The  difference  between  the  two  functions  is  small  at  low  pressures,  so 
that  at  pressures  of  a  few  atmospheres  the  two  functions  can  be  used 
interchangeably  for  solids.  This  of  course  does  not  hold  for  gases,  for 
which  the  volume  V  is  much  greater,  and  the  term  PV  is  very  large  even 
at  small  pressures.  As  we  can  see  from  Chap.  XI,  Sees.  3  and  4,  this 
diagram,  of  G  as  a  function  of  77,  is  the  important  one  in  discussing  the 
equilibrium  of  phases,  since  the  condition  of  equilibrium  is  that  the  two 
phases  should  have  the  same  Gibbs  free  energy  at  the  same  pressure  and 
temperature.  Thus  if  we  draw  G  for  each  phase,  as  a  function  of  tem- 
perature, for  the  pressure  at  which  the  experiment  is  carried  out,  the  point 
of  intersection  will  give  the  equilibrium  temperature  of  the  two  phases  at 
the  pressure  in  question. 


SEC.  3) 


THE  EQUATION  OF  STATE  OP'  SOL1DX 


211 


We  have  already  shown,  in  Figs.  XI-4,  XI-5,  XI-6,  and  XI-7,  the 
equation  of  state,  entropy,  and  Gibbs  free  energy  of  a  substance  in  all  of 
its  three  phases.  Examination  of  the  parts  of  those  figures  dealing  with 
solids  will  show  the  similarity  of  those  curves  to  tho  ones  found  in  tho 
present  section  in  a  more  explicit  and  detailed  way. 

3.  The  Statistical  Mechanics  of  Solids. — The  first  step  in  discussing 
a  solid  according  to  statistical  mechanics  is  to  set  up  a  model,  describing 
its  coordinates  and  momenta,  finding  its  energy  levels  according  to  the1 
quantum  theory,  and  computing  the  partition  function.  This  represents 
an  extensive  program,  of  which  only  the  outline  can  bo  given  in  the 
present  chapter.  The  typical  solid  is  a  crystal,  a  regular  repeating 


l_ 

s. 
3 


P-  20. 000  atm 


P- 10,000  atm 


0  100  200  300 

Deg  Abs. 

FIG.  XIII-6. — Gibbs  free  energy  of  a  solid  (sodium)  as  function  of  temperature  at  constant 

pressure. 

structure  composed  of  molecules,  atoms,  or  ions.  The  repeating  unit  is 
called  the  unit  cell.  The  crystal  is  hold  together  by  forces  between  the 
molecules,  atoms,  or  ions — forces  that  resemble  those  between  atoms  in 
diatomic  molecules,  as  discussed  in  Chap.  IX,  in  that  they  lead  to  attrac- 
tion at  large  distances,  repulsion  at  small  distances,  with  equilibrium 
between.  The  interplay  of  the  attractive  and  repulsive  forces  of  all 
atoms  of  the  crystal  leads  to  a  state  of  equilibrium  in  which  each  atom 
has  a  definite  position,  in  which  no  forces  act  on  it.  At  the  absolute  zero 
of  temperature,  the  atoms  will  be  found  just  in  these  positions  of  equilib- 
rium. At  higher  temperatures,  however,  they  will  vibrate  about  the 
positions  of  equilibrium,  to  which  they  are  held  by  forces  proportional  to 
the  displacement,  if  the  displacements  are  small.  We  shall  divide  our 
discussion  of  the  model  into  two  parts:  first,  the  crystal  at  the  absolute 


212  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIII 

zero  with  its  atoms  at  rest  in  their  equilibrium  positions;  secondly,  the 
thermal  vibrations  of  the  atoms  about  these  positions. 

Let  us  first  consider  the  crystal  at  the  absolute  zero.  The,  energy 
will  depend  on  the  state  of  strain  of  the  crystal;  as  was  mentioned  in 
Sec.  1,  we  omit  discussion  of  shearing  strains  and  types  of  deformation 
other  than  change  of  volume.  Thus  we  are  interested  simply  in  the 
dependence  of  energy  on  volume.  As  the  volume  is  changed,  of  course 
each  unit  cell  changes  in  the  same  proportion  and  the  atoms  change  their 
positions  in  the  crystal.  The  interatomic  energies  also  change,  and  the 
change  in  energy  of  the  whole  crystal  is  simply  the  sum  of  the  changes 
of  the  energies  of  interaction  of  the  various  atoms.  We  cannot  say 
anything  further  about  the  energy  as  a  function  of  volume,  without 
investigating  specific  examples,  as  we  shall  do  in  later  chapters.  But  at 
least  we  may  assume  that  the  energy  of  interaction  of  two  atoms  is  most 
conveniently  expressed  as  a  function  of  the  distance  of  separation,  and  if 
the  whole  energy  is  a  sum  of  these  energies  of  interaction,  it  also  may  be 
expected  to  be  particularly  simple  when  regarded  as  a  function  of  a  linear 
dimension  of  the  crystal,  rather  than  as  a  function  of  the  volume.  We 
shall,  therefore,  express  the  energy  of  the  crystal  at  the  absolute  zero  as  a 
function  of  a  quantity  r,  which  may  be  a  distance  between  atoms,  a  side  of 
a  unit  cell,  or  some  other  linear  dimension  of  the  crystal,  and  shall  find 
results  that  will  later  be  useful  to  us,  when  we  knowr  more  about  the  nature 
of  the  interatomic  forces. 

Consider  a  crystal  of  volume  V,  containing  N  atoms  or  molecules. 
(We  purposely  leave  the  description  slightly  vague,  so  as  to  allow  more 
generality  in  the  result.)  Then  V/N  is  the  volume  per  atom  or  molecule, 
a  quantity  which  of  course  can  be  changed  by  application  of  external 
pressure.  We  shall  limit  the  present  discussion  to  cubic  crystals,  in 
which  only  the  volume,  and  not  the  shape,  changes  under  pressure;  many 
crystals  do  not  have  this  property,  but  the  ones  that  we  shall  discuss 
quantitatively  happen  to  be  cubic.  Then  V/N  will  be  a  numerical 
constant  times  r3,  the  volume  of  a  cube  of  side  r,  since  in  a  uniform  com- 
pression the  whole  volume  and  the  volume  r3  will  change  in  proportion. 
Thus  let 

jj  =  <**,  (3-1) 

where  c  will  be  a  definite  number  for  each  structure,  which  we  can  easily 
evaluate.  We  define  a  quantity  r0  in  terms  of  F0,  the  values  respectively 
of  r  and  V  when  the  crystal  is  under  no  pressure  at  the  absolute  zero. 
Thus  we  have 

Fo  -V       r*  -r3  _ 
V~o    ~  rjj 


SBC.  3]  THE  EQUATION  OF  STATE  OF  SOLIDS  213 

We  now  take  the  expression  (2.1)  for  the  internal  energy  as  a  function 
of  volume,  set  the  temperature  equal  to  zero,  and  use  Eq.  (3.2),  finding 
the  internal  energy  at  absolute  zero  as  a  function  of  the  linear  dimensions. 
Calling  this  quantity  t/o,  we  have 


where  PJ,  P£  are  the  values  of  the  quantities  PI,  P2  of  Eq.  (2.1)  at  the 
absolute  zero  of  temperature.  (We  note  that  P0  ==  0  at  the  absolute 
zero.)  Substituting  from  Eq.  (3.2)  and  retaining  terms  only  up  to  the 
third,  we  have 


[/00  +  NcrP°~-~      -  9(P?  -  pojTir.     (3.4) 


Equation  (3.4)  will  later  prove  to  bo  convenient,  in  cases  where  we  have 
a  theoretical  way  of  calculating  t/o  from  assumed  interatomic  forces. 
In  these  cases,  P?  and  PI  can  be  found  directly  from  the  theory,  using 
Eq.  (3.4). 

Our  next  task  is  to  consider  the  solid  at  a  higher  temperature  than 
the  absolute  zero.  The  molecules  and  atorps  will  have  kinetic  energy  and 
will  vibrate.  We  can  get  a  simple,  but  incorrect,  picture  of  the  vibrations 
by  thinking  of  all  the  atoms  but  one  as  being  fixed  and  asking  how  that 
one  would  move.  It  is  in  a  position  of  stable  equilibrium  at  the  absolute 
zero,  being  held  by  its  interactions  with  its  neighbors  in  such  a  way  that 
it  is  pushed  back  to  its  position  of  equilibrium  with  a  force  proportional 
to  the  displacement.  Thus  it  will  execute  simple  harmonic  motion,  with 
a  certain  frequency  v.  To  discuss  the  heat  capacity  of  this  oscillation, 
we  may  proceed  exactly  as  in  Chap.  IX,  Sec.  5,  where  we  were  talking 
about  the  heat  capacity  of  molecular  vibrations.  Each  atom  can  vibrate* 
in  any  direction,  so  that  its  #,  ?/,  and  z  coordinates  separately  can  execute 
simple  harmonic  motion.  It  is  then  found  easily  that  the  classical 
partition  function  for  vibration  for  a  single  atom  is  (kT/hv)*,  similar  to 
Eq.  (5.4),  Chap.  IX,  but  cubed  on  account  of  the  three  dimensions.  This 
corresponds  to  a  heat  capacity  of  3k  per  atom,  or  3R  per  mole,  if  the 
material  happens  to  be  monatomic,  with  corresponding  values  for  poly- 
atomic substances.  This  law,  that  the  heat  capacity  of  a  monatomic 
substance  should  be  3R  or  5.96  cal.  per  mole,  at  constant  volume,  is  called 
the  law  of  Dulong  and  Petit.  It  is  a  law  that  holds  fairly  accurately  at 
room  temperature  for  a  great  many  solids  and  has  been  known  for  over  a 
hundred  years.  It  was  first  found  as  an  empirical  law  by  Dulong  and 
Petit.  At  lower  temperatures,  however,  the  specific  heats  of  actual  solids 
are  less  than  the  classical  value,  and  decrease  gradually  to  zero  at  the 
absolute  zero. 


214  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 

It  was  to  explain  these  deviations  from  the  law  of  Dulong  and 
Petit  that  Einstein  developed  his  theory  of  specific  heats.  He  treated  the 
vibrations  of  the  separate  atoms  by  quantum  theory,  just  as  we  did  in 
Sec.  5,  Chap.  IX,  and  derived  the  formula 


/     hv  \2 

(«w-i) 


k2 

-  1; 
e 


6 

(3.5) 


where 


__ 
y 
-  l) 


0  =  ~  (3.6) 


Equation  (3.5)  is  analogous  to  Eq.  (5.7),  Chap.  IX,  but  is  multiplied  by 
3  on  account  of  the  three  degrees  of  freedom.  As  we  have  seen  in  Fig. 
IX-3,  this  gives  a  specific  heat  rising  from  zero  at  the  absolute  zero  to  the 
classical  value  3/2  at  high  temperatures.  It  is  found  that  values  of  the 
frequency  i>,  or  of  the  corresponding  characteristic  temperature  9,  of 
Eq.  (3.6),  can  be  found,  such  that  the  Einstein  specific  heat  formula  (3.5) 
gives  a  fairly  good  approximation  to  the  observed  specific  heats,  except  at 
very  low  temperatures.  Close  to  the  absolute  zero,  the  Einstein  formula 
predicts  a  specific  heat  falling  very  sharply  to  zero.  The  actual  specific 
heats  do  not  fall  off  so  rapidly,  but  instead  are  approximately  propor- 
tional to  T3  at  low  temperatures.  Thus,  while  Einstein's  formula  is 
certainly  a  step  in  the  right  direction,  we  cannot  consider  it  to  be  correct. 
The  feature  that  we  must  correct  is  the  one  in  which  we  have  already 
rioted  that  this  treatment  is  inadequate:  the  atoms  are  not  really  held  to 
positions  of  equilibrium  but  merely  to  each  other.  In  other  words,  we 
must  treat  the  solid  as  a  system  of  many  atoms  coupled  to  each  other,  arid 
we  must  find  the  vibrations  of  these  atoms.  This  is  a  complicated  prob- 
lem in  vibration  theory,  something  like  the  problems  met  in  the  vibrations 
of  polyatomic  molecules.  We  shall  not  take  it  up  until  the  next  chapter. 
In  the  meantime,  however,  there  are  certain  general  results  that  we  can 
find  regarding  such  vibrations,  which  are  enough  to  allow  us  to  make 
considerable  progress  toward  understanding  the  equation  of  state  of 
solids. 

A  system  of  N  particles,  held  together  by  elastic  forces,  has  in  general 
3AMJ  vibrational  degrees  of  freedom,  as  we  saw  in  Eq.  (6.1),  Chap.  IX, 
where  we  were  talking  about  polyatomic  molecules.  Really  a  whole 
crystal,  or  solid,  can  be  regarded  as  an  enormous  molecule,  and  for  large 
values  of  N  we  can  neglect  the  6,  saying  merely  that  there  are  SN  vibra- 


SEC.  4]  THE  EQUATION  OF  STATE  OF  SOLIDS  215 

tional  degrees  of  freedom.  In  general,  there  will  then  be  3N  different 
normal  modes  of  vibration,  as  they  are  called.  Each  normal  mode 
consists  of  a  vibration  of  all  the  atoms  of  the  crystal,  each  with  its  own 
amplitude,  direction,  and  phase,  but  all  with  the  same  frequency.  Each 
atom  then  finds  itself  surrounded,  not  by  a  stationary  group  of  neighbors, 
but  by  neighbors  which  are  oscillating  with  the  same  frequency  as  its  own 
motion.  At  each  point  of  its  path,  it  will  always  find  the  neighbors  in 
definite  locations,  so  that  the  forces  exerted  on  it  by  its  neighbors  will 
depend  only  on  its  position;  but  the  forces  will  not  be  the  same  as  if  the 
neighbors  remained  at  rest,  for  the  positions  will  be  different.  Thus  the 
frequency  will  not  be  the  same  as  assumed  in  Einstein's  theory.  Our 
problem  in  the  next  chapter  will  be  to  consider  these  3N  modes  of  vibra- 
tion and  to  find  their  frequencies,  which  in  general  will  all  be  different. 
For  the  present,  however,  we  may  simply  assume  the  frequencies  to  be 
known,  and  equal  to  vi  .  .  .  PS.V.  The  most  general  motion  of  the 
atoms,  of  course,  is  not  one  of  these  normal  modes  but  a  superposition  of 
all  of  them,  with  appropriate  amplitudes.  This  has  a  simple  and  in  fact 
a  very  fundamental  analogy  in  the  theory  of  sound.  The  normal  modes 
of  a  vibrating  string  or  other  musical  instrument  are  simply  the  different 
harmonic  overtones  in  which  it  can  vibrate.  Each  one  consists  of  a  purely 
sinusoidal  vibration,  in  which  the  string  is  divided  up  by  nodes  into  cer- 
tain vibrating  segments.  The  simplest  type  of  vibration  of  the  string  is 
an  excitation  of  only  one  of  these  overtones,  so  that  it  vibrates  with  a 
pure  musical  tone.  But  the  more  general  and  common  type  of  vibration 
is  a  superposition  of  many  overtones,  each  with  an  appropriate  amplitude 
and  phase;  it  is  such  a  superposition  which  gives  a  sound  of  interesting 
musical  quality.  As  a  matter  of  fact,  if  we  ask  about  the  3N  vibrations 
of  a  piece  of  matter,  for  studying  its  specific  heat,  we  find  that  the  vibra- 
tions of  low  frequency  are  exactly  those  acoustical  vibrations  which  are 
considered  in  the  theory  of  sound.  As  we  go  to  higher  and  higher  fre- 
quencies and  shorter  and  shorter  wave  lengths,  however,  the  vibrations 
begin  to  depart  from  the  simple  ones  predicted  by  the  ordinary  theory  of 
sound,  and  finally  when  the  wave  length  begins  to  be  comparable  with  the 
interatomic  distance,  the  departure  is  very  great.  We  shall  investigate 
the  nature  of  these  vibrations,  as  well  as  their  frequencies,  in  the  next 
chapter. 

4.  Statistical  Mechanics  of  a  System  of  Oscillators. — I^jiamically, 
we  have  seen  that  a.jprystal  can  be  approximated  by  a  set  of  3N  vibra- 
tions, if  there  are  jV^atonaa  in  the,  crystal.  These  vibrations  have  fre- 
quencies which  we  may  label  vi  .  .  .  VZN,  varying  through  a  wide  range  of 
frequencies.  To  the  approximation  to  which  the  restoring  forces  can  be 
treated  as  linear,  these  oscillations  are  independent  of  each  other,  each  one 
corresponding  to  a  simple  harmonic  oscillation  whose  frequency  is  inde- 


216  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 

pendent  of  its  own  amplitude  or  of  the  amplitudes  of  other  harmonics. 
This  is  only  an  approximation,  but  it  is  sufficient  for  most  purposes. 
Then  the  energy  is  the  sum  of  the  energies  of  the  various  oscillators,  and 
each  of  these  is  quantized.  That  is,  the  energy  of  the  jth  oscillator 
ran  take  on  the  values  (n/  +  ?)hvj,  where  n/,  an  integer,  is  the  quantum 
number  associated  with  this  oscillator.  We  see  that  3N  quantum  num- 
bers aro  necessary  to  describe  the  total  energy  and  to  define  a  stationary 
state.  All  these  quantum  numbers  should  then  appear  as  subscripts  of 
the  energy,  and  we  have  the  relation 


where  U'Q  is  the  energy  which  the  lattice  would  have  if  the  amplitudes 
of  all  oscillations  were  zero.  Actually,  even  at  the  absolute  zero  of 
temperature,  however  each  oscillation  has  a  half  quantum  of  energy. 
Thus  we  may  write 

3N 


where 

3tf 

f/o  =  U'0  +  5iAiv.  (4.3) 


The  quantity  t/o  is  the  same  as  that  givon  in  Eq.  (3.3),  representing  the 
energy  of  the  lattice,  as  a  function  of  volume,  at  the  absolute  zero  of  tem- 
perature. The  subscripts  n\  .  .  .  HS.V  take  the  place  of  the  single  index  i 
which  we  ordinarily  use  in  defining  the  partition  function.  Thus  we  find 
for  the  partition  function 


_  __        - 

z  =  ekT  =  2  •  •  •  2c  jkT  ekT  (4*4) 

HI  U3N 

We  can  write  the  exponential  as  a  product  of  terms  each  coming  from  a 
single  value  of  j,  and  can  carry  out  the  summations  separately,  obtaining 

(4.5) 

Each  of  the  summations  in  Eq.  (4.5)  is  of  the  form  already  evaluated  in 
Sec.  5,  Chap.  IX.     Thus  we  have 

1 


SBC.  4]  THE  EQUATION  OF  STATE  OF  SOLIDS  217 

a  product  of  3N  terms.    Taking  the  logarithm,  we  have  at  once 

A  «  C7o  +  2*!T  In  (l  -  c~$).  (4.7) 

3 

Differentiating  with  respect  to  7\  we  have 

(hv,     v 
/  _^\  IT      \ 

-1»  U  -''   "7  +-J-  —  )•  (4-8) 

e*r  -  I/ 
Finally,  from  Eq.  (5.20),  Chap.  Ill,  we  hav« 


By  differentiating  Eq.  (4.9),  we  find  the  specific  heat  in  agreement  with 
the  value  previously  found,  Eq.  (3.5),  for  the  special  case  whoro  all  3N 
frequencies  are  equal. 

Having  found  the  Helmholtz  free  energy  (4.7),  we  can  find  the  pressure 
by  differentiating  with  respect  to  volume.     We  have 


p  =  —  I  — 

\dV 


where 

The  first  two  terms  in  Eq.  (4.10)  are  the*  pressure  at  the  absolute  zero  of 
temperature,  which  wo  have  already  discussed.  The  summation  repre- 
sents the  thermal  pressure.  It  is  different  from  zero  only  because  the 
7/s  are  different  from  zero;  that  is,  because  the  vibrational  frequencies 
depend  on  volume.  Wenaturally  expect  this  dependence;  as  the  crystal 
is  compressed  it  becomes  harder,  the  restoring  forces  become  greater,  and 
vibrational  frequencies  increase,  so  that  the  P/S  increase  with  decreasing 
volume  and  the  7/5  are  positive.  If  we  consider  the  7/3  to  be  inde- 
pen9ehEDf~ temperature,  each  term  of  the  summation  in  Eq.  (4.10)  is 
proportional  to  the  energy  of  the  corresponding  oscillator  as  a  function  of 
temperature,  given  by  Eq.  (4.9),  divided  by  the  volume  V. 

At  high  temperatures,  we  know  that  the  quantum  expression  for  the 

i  hv- 

energy  of  an  oscillator,  ^hvj  +  -A- — - ->  approaches  the  classical  value 

~ — —  -~—  -  ••IIIIIMH_     imi  n  miiii  a  i«  ••«II"W«B<II«  nil  <  >ui«iM»M«»»mMk « „«,,,    ,   _ 

\  ekT  —  1 


218  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 

kT.     Thus,  at  high  temperature  the  thermal  pressure  approaches^ 


The  first  term  is  an  expression  similar  to  the  pressure  of  a  perfect  gas, 
NkT/V,  except  that  the  constant  of  proportionality  is  now  Vy/  instead  of 

3 

N,  the  number  of  molecules.  We  shall  find  that  the  7/s  are  generally 
between  1  and  2,  so  that,  since  there  are  3N  terms  in  the  summation 
ovorj,  whore  N  is  the  number  of  atoms,  the  thermal  pressure  asjndicated 
by  Eq.  (4.12)  has  a  term  which  is  from  three  to  six  times  as  greatjts  the 
corresponding  pressure  of  a  perfect  gas  of  the  same  number  of  atoms,  at 
the  same  temperature  and  volume  as  the  solid.  The  second  term  of 
Eq.  (4.12),  coming  from  the  zero  point  energy,  gives  a  decrease  of  thermal 
pressure  compared  to  this  gas  pressure  which  is  independent  of  tempera- 
ture, until  wo  go  down  to  low  temperatures.  There  the  thermal  pressure 
does  not  decmise  so  rapidly  with  decreasing  temperature  as  we  should 
estimate  from  the  gas  law,  but  instead  falls  to  the  value  zoro  at  the 
absolute  zero.  The  change  of  thermal  pressure  with  temperature,  involv- 
ing the  derivative  of  Eq.  (4.10),  approaches  zoro  very  strongly  at  the 
absolute  zoro,  just  as  the  specific  boat  does,  and  sinoo  this  derivative 
enters  the  formula  for  thermal  expansion,  this  quantity  goes  to  zero  at 
the  absolute  zoro.  In  fact,  as  we  shall  soo  in  the  noxt  paragraph,  there  is  a 
close  connection  between  the  thermal  expansion  and  the  specific  heat. 

To  got  a  complete  equation  of  state  from  statistics,  we  need  only 
take  the  expression  (4.10)  for  P  and  expand  tho  summation,  the  thermal 
pressure,  as  a  power  series  in  (Fo  —  F)/Fo.  This  can  be  donqjf  we  can 
nniiJj£  de£)endence  of  .the  i//s  on  volume,  from  the  theory.  Then  we  can 
identify  tho  resulting  equation  with  Eq.  (1.5),  equating  coefficients,  and 
findPo,  -Pi,  and  P^  which  determine  the  pressure,  in  terms  of  the  structure 
of  the  crystal.  Knowing  the  moaning  of  P0,  PI,  and  PZ  from  our  earlier 
discussion,  this  allows  us  to  find  the  thermal  expansion,  compressibility, 
and  change  of  compressibility  with  pressure,  as  functions  of  temperature. 
Since  very  fow  experiments  are  available  dealing  with  the  changes  of  these 
quantities  with  temperature,  we  shall  confine  our  attention  to  the  thermal 
expansion  at  zero  pressure;.  From  Kq.  (1.2),  this  is  approximately 
ddo/dT,  where  from  Eq.  (1.9)  we  have  a0  =  Po/Pi.  Comparing  with 
Eq.  (4.10)  above,  we  soe  that  P0  is  the  value  of  the  summation  when 
V  =  Fo.  The  quantity  PI,  which  we  have  seen  to  be  the  reciprocal  of 
the  compressibility,  equals  PJ  plus  a  small  term  coming  from  the  summa- 
tion, which  we  can  neglect  for  a  very  rough  discussion,  though  of  course  it 
would  have  to  be  considered  for  accurate  work.  Since  P?  is  independent 


SEC.  4]  THE  EQUATION  OF  STATE  OF  XOLIDS  219 

of  temperature,  this  gives  us 


k  "%ri      \kT/ 
Thermal  expansion  =  X-  >j  y,  \u.  —  ry  (4.13) 


u.   —  r 
^-  1/ 


where  x  is  the  compressibility.  Comparison  of  Eq.  (4.13)  with  the 
formula  for  heat  capacity  of  linear  oscillators,  for  example  Eq.  (3.5), 
shows  at  once  the  close  relation  between  the  heat-  capacity  and  the  thermal 
expansion.  Each  term  in  Eq.  (4.13)  is  proportional  to  the  term  in  the 
heatjjapacity  arising  from  the  same  oscillator,  so  that  the  thermal  expan- 
sion shows  qualitatively  the  same  sort  of  behavior,  becoming  constant  at 
high  temperatures  but  reducing  to  zero  as  the  temperature  approaches 
the  absolute  zero.  While  experimental  data  for  thermal  expansions  are 
not  ne^lj^ojBxUinsive  jis  those  for  specific  heats,  still  they  are  sufficient 
tojjjhow  that  this  is  actually  the  observed  behavior. 

To  allow  the  construction  of  a  simple  theory  of  thermal  expansions, 
Grlineisen  assumed  that  the  quantities  7,  were  all  equal  to  each  other  and 
to  a  constant  7,  which  he  regarded  as  an  empirical  constant.  To  see  the 
meaning  of  7,  we  assume  that  the  frequencies  v}  are  given  in  terms  of  the 
volume  by  the  relation 

",  =  |~>  (4.14) 

where  c3  is  a  constant,  so  that  the  frequencies  are  inversely  proportional 
to  the  7  power  of  the  volume.  Since  surely  the  p/s  increase  with  decreas- 
ing volume,  this  is  a  reasonable  form  of  dependence  to  assume.  Then  we 
find  at  once  that 

-l1^      -      y  (Al&) 

d  In  V  "  7'  C4-15' 

so  that  the  7  defined  in  Eq.  (4.14)  is  the  same  as  the  7,  of  Eq.  (4.J1). 
We  see  that  7  =  1  or  7  =  2  respectively  corresponds  to  the  frequencies 
being  inversely  proportional  to  the  volume  or  to  the  square  of  the  volume. 
If  we  assume  with  Griineiseri  that  7,-  =  7,  we  then  have  from  Eq.  (4.13) 

Cv     ] 
Thermal  expansion  =    -^--      '  (4.16) 

y  o 

Equation  (4.16)  is  a  relation  between  the  thermal  expansion,  compressi- 
bility, specific  heat,  volume,  and  the  parameter  7.  If  we  have  an  inde- 
pendent theoretical  way  of  finding  7,  we  can  use  it  to  compute  the  thermal 
expansion.  Otherwise,  we  can  use  measured  values  of  thermal  expansion, 
compressibility,  specific  heat,  and  volume,  to  find  empirical  values  of  y. 
Both  types  of  discussion  will  be  given  in  later  chapters,  where  we  discuss 


220  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIII 

specific  types  of  solids.  We  shall  find  that  the  agreement  between  the 
various  methods  of  finding  7  is  rather  good,  and  that  values  for  most 
ordinary  materials  are  between  1  and  3,  generally  in  the  neighborhood 
of  2. 

6.  Polymorphic  Transitions.  —  It  has  been  mentioned  in  Chap.  XI, 
Sec.  7,  that  the  transition  lines  in  the  P-T  diagram  between  polymorphic 
phases  of  the  same  substance  tend  to  have  positive  slope,  a  change  of 
pressure  of  something  less  than  12,000  atm.  corresponding  to  a  change  of 
temperature  of  about  200°.  With  our  present  knowledge  of  the  equation 
of  state  of  solids,  we  can  attempt  a  theoretical  explanation  of  this  relation. 
To  work  out  the  slope,  we  note  that  Clapeyron's  equation  can  be  written 
dP/dT  =  AS/A7,  where  AS  is  the  difference  of  entropy  between  one 
phase  and  the  other,  AF  the  difference  of  volume.  We  have  seen  in  the 
present  chapter  that  the  entropy  of  a  single  phase  increases  when  its  vol- 
ume increases,  and  we  have  found  quantitative  methods  of  calculating  the 
amount  of  change.  Let  us,  then,  tentatively  assume  that  the  relation 
of  change  of  entropy  to  change  of  volume  in  going  from  one  phase  to 
another  is  about  the  same  as  when  we  change  the  volume  of  a  single 
phase.  This  is  certainly  a  very  crude  assumption,  but  we  shall  find  that 
it  gives  results  of  the  right  order  of  magnitude.  The  assumption  we  have 
just  made  amounts  to  replacing  AS/AF  by  dS/dV,  computed  for  a  single 
phase.  Now  from  Eq.  (4.8)  we  can  write  the  entropy  of  a  substance  per 
gram  atom  as 

f  /  _±A        hv/kT  I 

S  =  3Nok\  -  In  VI  -  e   kT)  +  -jr  -  ,  (5.1) 

L  **  -  ij 

where  No  is  Avogadro's  number,  and  we  assume  for  simplicity  that  all 
frequencies  vt  are  the  same.  We  then  have 

rfS  _  dS  dv^  _  __    v_  dS  f 

dV  ~  dvdV  ~      7F  dv  (b'Z) 

where  7  has  the  same  significance  as  in  the  last  section.  Differentiating 
Eq.  (5.1),  this  leads  to 


dS  _  3AT0fc7AA2       F*  _  _  yCv  ,.  ^ 

dV          V    \kf)  ^  Ji)*~~V  ' 

Now  Cv  is  about  3R  per  gram  atom,  and  from  the  preceding  section  we  see 
that  7  is  about  2  for  most  substances.  Furthermore,  examination  of 
experimental  values  shows  that  V  is  of  the  order  of  magnitude  of  10  cc. 
per  gram  atom,  for  most  substances.  Putting  in  these  values  and  putting 
proper  units  in  Eq.  (5.3),  we  find  that  dS/dV  is  approximately  50  atm.  per 
degree,  or  10,000  atm.  for  200  degrees,  just  about  the  value  that  Bridgman 


SBC.  5]  THE  EQUATION  OF  STATE  OF  SOLIDS  221 

finds  to  be  most  common  experimentally.  Thejaot  1ftftt'  this  fifl.)milfl.t.inn 
agrees  so  welLwithjbhc  average  behavior  of  many  materials  is  some  justi- 
fication_for  thinking  that  the  major  part  of  the  entropy  change  from  Qwe 
polymorphic  phase  .to  another  is  simply  that  associated  with  the  change  of 
volume.  The  individual  variations  are  so  great,  however,  that  no  groat 
claim  for  accuracy  can  be  made  for  such  a  calculation  as  we  have  just 
made. 

In  the  present  chapter,  we  have  laid  the  foundations  for  a  statistical 
study  of  the  equation  of  state  of  solids,  though  we  have  not  made  any  use 
of  a  model,  and  hence  have  not  been  able  to  compute  the  thcrmodynamic 
quantities  we  have  been  talking  about.  We  proceed  in  the  next  chapter 
to  a  discussion  of  atomic  vibrations  in  solids,  with  a  view  to  finding  more 
accurate  information  about  specific  heats  and  thermal  expansion.  Later, 
when  we  study  different  types  of  solids  more  in  detail,  we  shall  make 
comparisons  with  experiment  for  many  special  cases. 


CHAPTER  XIV 
DEBYE'S  THEORY  OF  SPECIFIC  HEATS 

We  have  seen  in  the  last  chapter  that  the.  essential  step  in  investigating 
the  specific  heat  and  thermal  expansion  of  solids  is  to  find  the  frequencies 
of  the  normal  modes  of  elastic  vibration.  We  shall  take  this  problem 
up,  in  its  simplest  form,  in  the  present  chapter.  The  vibrations  of  a  solid 
are  generally  of  two  sorts:  vibrations  of  the  molecules  as  a  whole  and 
internal  vibrations  within  a  molecule.  This  distinction  of  course  can  be 
found  only  in  molecular  crystals  and  is  lacking  in  a  crystal,  like  that  of 
a  metal,  whore  all  the  atoms  are  of  the  same  sort.  For  this  reason  solids 
of  the  elements  have  simpler  specific  heats  than  compounds,  and  we  take 
them  up  first,  postponing  discussion  of  compounds  to  the  next  chapter. 
At  first  sight,  on  account  of  the  large  number  of  atoms  in  a  crystal,  it 
might  seem  to  be  impossibly  hard  to  solve  the  problem  of  their  elastic 
vibrations,  but  as  a  matter  of  fact  it  is  just  the  largo  number  of  atoms 
that  makes  it  possible  to  handle  the  problem.  For  the  vibrations  of  a 
finite  continuous  piece  of  matter  can  bo  handled  by  the  theory  of  elas- 
ticity, and  for  waves  long  compared  to  atomic  dimensions  this  theory  is 
correct.  We  start  therefore  by  considering  the  elastic  vibrations  of  a 
continuous  solid,  and  later  ask  how  the  vibrations  arc  affected  by  the 
fact  that  the  solid  is  really  made1  of  atoms.  We  have  already  mentioned 
briefly  in  Chap.  XIII,  Sec.  3,  the  close  relation  between  the  normal  modes 
of  vibration  of  a  solid  composed  of  atoms  and  the  harmonic  or  overtone 
vibrations  of  acoustics. 

1.  Elastic  Vibrations  of  a  Continuous  Solid.  -  It  is  well  known  that 
elastic  waves  can  be  propagated  through  a  solid.  The  waves  are  of  two 
sorts,  longitudinal  and  transverse,  having  different  velocities  of  propaga- 
tion. The  longitudinal  waves  are  analogous  to  tin;  sound  waves  in  a  fluid, 
while  the  transverse  waves,  which  cannot  exist  in  a  fluid,  also  have  many 
properties  similar  to  sound  waves  and  are  ordinarily  treated  as  a  branch 
of  acoustics.  The  velocities  of  both  sorts  of  waves  are  determined  by  the 
elastic  constants  of  the  material  and  are  independent  of  the  frequency,  or 
wave  length,  of  the  waves,  within  wide  limits.  The  waves  with  which  we 
are  familiar  have  frequencies  in  the  audible  range,  less  than  10,000  or 
15,000  cycles  per  second.  The  velocities  of  clastic  waves  in  solids  are  of 
the  order  of  magnitude  of  several  thousand  meters  per  second  (something 
like  ten  times  the  velocity  in  air).  Since  we  have 

222 


SBC.  1]  DE BYE'S  THEORY  OF  SPECIFIC  HEATS  223 

\v  =  y,  (1.1) 

where  X  is  the  wave  length,  v  the  frequency  or  number  of  vibrations  per 
second,  and  v  the  velocity,  the  {shortest  sound  waves  with  whieh  we  are 
familiar  have  a  wave  length  of  the  order  of  magnitude  of 

(5  X  105)       _ 

—  Q- =50  cm., 

taking  the  velocity  to  be  5000  m.  per  second,  the  frequency  10,000  cycles. 
By  methods  of  supersonics,  frequencies  up  to  100,000  cycles  or  more  can 
be  investigated,  corresponding  to  waves  of  something  like  5  cm.  length. 
There  is  every  reason  to  suppose,  however,  that  this  is  not  the  limit  for 
elastic,  waves.  In  fact,  we  have  every  reason  to  believo  that  waves  of 
shorter  and  shorter  wave  length,  and  higher  and  higher  frequency,  arc 
possible,  up  to  the  limit  in  which  the  wave  length  is  comparable  with  the 
distance  between  atoms.  It  is  obvious  that  the  wave  length  cannot  be 
appreciably  shorter  than  interatomic  distances.  In  fact,  if  the  wave 
length  were  just  the  interatomic  distance*,  successive  atoms  would  be  in 
the  same  phase  of  the  vibration,  and  there  would  not  really  be.  a  vibration 
of  one  atom  with  respect  to  another  at  all.  The  shortest  wave  which 
we  can  really  have  comes  when  successive  atoms  vibrate  opposite  to  each 
other,  so  that  the  \\ave  length  is  twice  the  distance  between  atoms.  It  is 
interesting  to  find  the  corresponding  order  of  magnitude  of  the  frequency 
of  vibration.  If  we  set  X  =  5  X  10~~s,  of  the  order  of  magnitude  of  twice1 
an  interatomic  distance  in  a  metal,  we  have1 


v 


(5  X  105)         inll        .  i 

=  (5~X  10~:~s)  =  cycles  per  second. 


This  is  a  frequency  of  an  order  of  magnitude  of  those  found  in  the  infrared 
vibrations  of  light  waves.  There  is  good  experimental  evidence  that 
such  frequencies  really  represent  the  maximum  possible  frequencies  of 
acoustical  vibrations. 

The  situation,  then,  is  that  there  is  a  natural  upper  limit  set  to  fre- 
quencies, and  lower  limit  to  wave  lengths,  of  elastic  waves,  by  the  atomic 
nature*  of  matter.  It  can  be  shown  theoretically  that  as  this  limit  is 
approached,  the  velocity  of  the  waves  no  longer  is  independent  of  wave* 
length.  However,  the  change  is  not  great;  it  changes  by  something  not 
more  than  a  factor  of  two.  This  change  is  the  only  important  difference 
between  a  vibration  theory  based  on  the  theory  of  elasticity  and  a  theory 
based  directly  on  interatomic  forces,  provided  only  that  we  recognize  the 
lower  limit  to  wave  lengths.  Our  first  approach  to  a  theory,  the  one  made 
by  Debye,  takes  account  of  the  lower  limit  of  wave  lengths  but  neglects 
the  change  of  velocity  with  frequency. 


224  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIV 

In  a  finite  piece  of  solid,  such  for  instance  as  a  rod,  standing  waves 
are  set  up.  These  arise  of  course  from  constructive  interference  between 
direct  and  reflected  waves,  and  they  exist  only  when  the  wave  length  and 
the  dimensions  of  the  solid  have  certain  relations  to  each  other.  For  the 
transverse  vibrations  of  a  string,  these  relations  are  very  familiar:  tho 
length  of  the  string  must  be  a  whole  number  of  half  wave  lengths.  For 
other  shapes  of  solid,  the  relations  are  similar  though  not  so  simple. 
Arranging  the  standing  waves  in  order  of  decreasing  wave  length  or 
increasing  frequency,  we  have  a  series  of  frequencies  of  vibration,  often 
called  characteristic  frequencies,  or  harmonics.  For  the  string,  these 
are  simply  a  fundamental  frequency  and  any  integral  multiple  of  this 
fundamental.  The  resulting  harmonics  or  overtones  form  the  basis  of 
musical  scales  and  chords.  For  other  shapes  of  solids,  the  relations  are 
not  simple,  and  the  overtones  do  not  form  pleasing  musical  relations 
with  the  fundamental.  Now  for  all  one  knows  in  ordinary  acoustical 
theory,  the  number  of  possible  overtones  is  infinite,  though  of  course  few 
of  them  can  be  heard  on  account  of  the  limitations  of  the  ear.  Thus  if  we 
have  a  string,  with  frequencies  which  are  integral  multiples  of  a  funda- 
mental, there  seems  no  reason  why  the  integer  cannot  be  as  large  as  wo 
please.  This  no  longer  holds,  as  we  can  immediately  see,  when  we  con- 
sider the  atomic  nature  of  the  solid.  For  we  have  just  mentioned  that 
there  is  an  upper  limit  to  possible  frequencies,  or  a  lower  limit  to  possible 
wave  lengths,  set  by  interatomic  distances.  The  highest  possible  over- 
tone will  have  a  frequency  of  the  order  of  this  limiting  frequency.  That 
means  that  the  solid  has  a  finite  number  of  possible  overtone  vibrations. 
And  now  we  see  the  relation  between  our  acoustical  treatment  and  the 
vibration  problem  we  started  with:  these  overtone  vibrations  are  just  the 
normal  modes  of  vibration  of  the  atoms  in  the  crystal,  which  we  wanted  to 
investigate.  If  there  are  N  atoms,  with  3N  degrees  of  freedom,  we  have 
mentioned  in  Chap.  XIII  that  we  should  expect  3N  modes  of  oscillation; 
when  we  work  out  the  number  of  overtones,  we  find  in  fact  that  there  are 
just  3AT  allowed  vibrations. 

The  most  general  vibrational  motion  of  our  solid  is  one  in  which 
each  overtone  vibrates  simultaneously,  with  an  arbitrary  amplitude  and 
phase.  But  in  thermal  equilibrium  at  temperature  T,  the  various  vibra- 
tions will  be  excited  to  quite  definite  extents.  It  proves  to  be  mathe- 
matically the  case  that  each  of  the  overtones  behaves  just  like  an 
independent  oscillator,  whose  frequency  is  the  acoustical  frequency  of  the 
overtone.  Thus  we  can  make  immediate  connections  with  the  theory  of 
the  specific  heats  of  oscillators,  as  we  have  done  in  Chap.  XIII,  Sec.  4.  If 
the  atoms  vibrated  according  to  the  classical  theory,  then  we  should  have 
equipartition,  and  at  temperature  T  each  oscillation  would  have  the  mean 
energy  kT.  This  means  that  each  of  the  N  overtones  would  have  equal 


SEC.  2]  DE  BYE'S  THEORY  OF  SPECIFIC  HEATS  225 

energy,  on  the  average,  so  that  the  energy  of  all  of  them  put  together 
would  be  3NkT,  just  as  we  found  in  Chap.  XIII,  Sec.  3,  by  considering 
uncoupled  oscillators.  The  fundamental  and  first  few  harmonics,  which 
are  in  the  audible  range,  would  have  the  average  energy  kT,  just  like  the 
harmonics  of  higher  frequency.  This  does  not  mean  that  we  should  be 
able  to  hear  the  rod  in  thermal  equilibrium,  because  kT  is  such  a  small 
energy  that  the  amplitude  of  each  overtone  would  be  quite  inappreciable. 
Of  the  3N  harmonics,  by  far  the  largest  number  come  at  extremely  high 
frequencies,  and  it  is  here  that  the  thermal  energy  is  concentrated.  The 
superposition  of  these  high  frequency  overtone  vibrations,  each  with 
energy  proportional  to  the  temperature,  is  just  what  we  mean  by  tem- 
perature vibration,  and  the  energy  is  the  ordinary  internal  energy  of  the 
crystal.  Actually  the  oscillations  take  place  according  to  the  quantum 
theory  rather  than  the  classical  theory,  and  we  have  seen  in  Chap.  XIII, 
Sec.  4,  how  to  handle  them.  Each  frequency  v}  can  have  a  characteristic 
temperature  6,  associated  with  it,  according  to  the  equation 


(1.2) 
Then  the  heat  capacity  is 


so  that  the  heat  capacity  associated  with  each  oscillator  will  be  zero  at 
temperatures  much  below  9,,  rising  to  the  classical  value  at  temperatures 
considerably  above  O/.  For  the  lower  harmonics,  the  characteristic 
temperatures  are  extremely  low,  so  that  these  vibrations  are  excited  in  a 
classical  manner  at  any  reasonable  temperature.  The  highest  harmonics, 
however,  have  values  of  0,  in  the  neighborhood  of  room  temperature,  and 
since  many  of  the  harmonics  come  in  this  range,  the  specific  heat  does  not 
attain  its  classical  value  until  temperatures  somewhat  above  room 
temperatures  are  reached. 

2.  Vibrational  Frequency  Spectrum  of  a  Continuous  Solid. — To  find 
the  specific  heat,  on  the  quantum  theory,  we  must  superpose  Einstein 
specific  heat  curves  for  each  natural  frequency  vn  as  in  Eq.  (1.3).  Before 
we  can  do  this,  we  must  find  just  what  frequencies  of  vibration  are 
allowed.  Let  us  assume  that  our  solid  is  of  rectangular  shape,  bounded  by 
the  surfaces  x  =  0,  x  =  X,  y  =  0,  y  =  Y,  z  =  0,  z  =  Z.  The  fre- 
quencies will  depend  on  the  shape  and  size  of  the  solid,  but  this  does  not 
really  affect  the  specific  heat,  for  it  is  only  the  low  frequencies  that  art* 
very  sensitive  to  the  geometry  of  the  solid.  As  a  first  step  in  investigating 
the  vibrations,  let  us  consider  those  particular  waves  that  are  propagated 
along  the  x  axis. 


226  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIV 

It  is  a  familiar  fact  that  there  arc  two  sorts  of  waves,  traveling  and 
standing  waves,  and  that  a  standing  wave  can  be  built  up  by  superposing 
traveling  waves  in  different  directions.  We  start  with  a  traveling  wave 
propagated  along  the  x  axis.  Let  us  suppose  that  the  point  which  was 
located  at  x,  yy  z  in  the  unstrained  medium  is  displaced  by  the  wave  to 
the  point  .r  +  £,  //  +  r/,  z  +  f ,  so  that  £,  17,  f  are  the  components  of  dis- 
placement. To  describe  the  wave,  we  must  know  £,  r/,  f  as  functions  of 
x,  y,  z,  t.  For  a  wave  propagated  along  the  x  axis,  £  represents  a  longi- 
tudinal displacement,  r?  and  f  transverse  displacements.  For  the  sake  of 
definiteness  lot  us  consider  a  longitudinal  wave.  Then  the  general  expres- 
sion for  a  longitudinal  wave  traveling  along  the  x  axis,  with  velocity  v, 
frequency  vy  amplitude1  A,  and  phase  a,  is 


•HH- 


£  =  A  sin  27r   pit  -  ^ )  -  a  •  (2.1) 

Rather  than  using  the  phase  constant  a,  it  is  often  convenient  to  use  both 
sine  and  cosine  terms,  with  independent  amplitudes  .4  and  B,  obtaining 

£  =  A  cos  2irp(  t  -  -  )  +  B  sin  2irv(t  -  -Y  (2.2) 

\        v/  \        v) 

an  expression  equivalent  to  Eq.  (2. 1)  if  the  constants  A  and  B  of  Eq.  (2.2) 
have  the  proper  relation  to  the  A  and  a  of  Eq.  (2. 1).  Still  another 
way  to  write  such  a  wave,  this  time  using  complex  notation,  is 

£  -  Ae\~\  (2.3) 

where  the  A  of  Eq.  (2.3)  is  still  another  constant,  which  may  be  complex 
and  so  take  care  of  the  phase.  In  Eq.  (2.3)  it  is  to  be  understood  that 
the  real  part  of  the  complex  expression  is  the  value  to  be  used. 

/      A 

By  writing  expressions  in  l^+~)  analogous  to  Eqs.   (2.1),   (2.2), 

and  (2.3),  we  get  waves  propagated  along  the  negative  x  axis.  Adding 
such  a  wave  to  the  one  along  the  positive  x  axis,  we  have  a  standing 
wave.  As  a  simple  example,  we  take  the  case  of  Eq.  (2.2)  and  let  B  =  0. 
Then  we  have 


=  A  cos  2wp(t  -  ~ J  +  A  cos  2Trv\t  +  * 

A  (        rt     ,.         2irvx    .     .     n     J    .     2irvx 
=  A  I  cos  2wpt  cos h  sin  2wvt  sin 


.          rt     .  .     n     .    .     2irvx\ 

+  cos  2wvt  cos sm  2vvt  sin I 

v  / 

=  2A  cos  27Tpt  cos         -  (2.4) 


SEC.  2]  DEBYE'S  THEORY  OF  SPECIFIC  HEATS  227 

By  using  different  combinations  of  functions,  we  can  get  standing  waves 


P    ,  ,>    „    •     2irvx     .     rt    A         2irvx        i    •     rt    ,    - 

of  the  form  cos  2irvt  sin  -  >  sin  2irvt  cos  -  ;  and  sin  2^vt  sm  -  as 

'    v  v  v 

well.  The  particular  characteristic  of  a  standing  wave  is  that  the  dis- 
placement is  the  product  of  a  function  of  the  time  and  a  function  of  the 
position  x.  As  a  result  of  this,  the  shape,  given  by  the  function  of  x,  is 
the  same  at  any  instant  of  time,  only  the  magnitude  of  the  displacement 
varying  from  instant  to  instant. 

Certain  boundary  conditions  must  be  satisfied  at  the  surfaces  of  the 
solid.  For  instance,  the  surface  may  be  hold  rigidly  so  that  it  cannot 
vibrate,  or  it  may  be  in  contact  with  the  air  so  that  it  cannot  develop  a 
pressure  at  the  surface.  The  allowed  overtones  will  depend  on  the 
particular  conditions  we  assume,  but  again  this  is  important  only  for  tho 
low  overtones  and  is  immaterial  for  the  high  frequencies.  To  bo  specific, 
then,  let  us  assume  that  the  surface  is  hold  rigidly,  so  that  the  displace- 
ment £  is  zoro  on  the  surface*,  or  when  x  =  0,  x  =  X.  The  first  con- 
dition can  be  satisfied  by  using  a  standing  wave  containing  the  factor 

sin  -  j  rather  than  cos  -  >  since  sin  0  =  0.     Then  for  the  second  condi- 
v  v 

tion  we  must  have 

.     2irvX       _  /rt  _ 

sin  -  =  0.  (2.5) 

Condition  (2.5)  can  be  satisfied  in  many  ways,  for  we  know  that  the 
sine  of  any  integer  times  IT  is  zero.  Thus  we  satisfy  our  boundary  condi- 
tion if  we  make 


";     =  0,  1,  2,  •••=*,  (2.6) 

where  s  is  an  integer.     Using  the  relation  v/v  =  I/A,  this  can  be  written 

1  C  tt\ 

I  «  *A  v-  /rfc  _x 

A  =  2X'         °r        -2   -  X<  (2'7) 

showing  that  a  whole  number  of  half  wave  lengths  must  be  contained 
in  the  length  of  the  solid.  Equation  (2.6)  or  (2.7)  solves  entirely  the 
problem  of  the  allowed  vibrations  of  a  continuous  solid,  so  long  as  we 
limit  ourselves  to  longitudinal  waves  propagated  along  the  x  direction. 
If  we  introduce  the  additional  condition,  demanded  by  the  atomic  nature 
of  the  medium,  that  the  minimum  wave  length  is  twice  the  distance 
between  atoms,  we  can  immediately  find  the  number  of  such  possible 
overtones.  Let  there  be  No  atoms  in  a  row  in  the  length  X.  Then  the 
distance  between  atoms,  along  the  x  axis,  is  X/No.  Our  condition  for 
the  maximum  possible  overtone  is  then  X/2  =  X/N*,  or  AT0X/2  =  X, 
showing  that  there  are  just  NQ  overtones  corresponding  to  propagation 


228  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIV 

along  the  x  axis.  If  we  investigate  transverse  vibrations,  but  propagation 
along  the  x  axis,  we  obtain  exactly  analogous  results,  but  with  t\  or  f 
substituted  for  £.  The  allowed  wave  lengths  for  transverse  vibrations 
are  the  same  as  for  longitudinal  ones,  but  on  account  of  the  fact  that 
the  velocity  of  transverse  waves  is  different  from  that  of  longitudinal 
waves,  the  frequencies  are  different.  There  are  No  possible  vibrations 
for  each  of  the  two  directions  of  transverse  vibration,  giving  3No  vibra- 
tions in  all  corresponding  to  propagation  along  the  x  axis. 

We  can  now  use  the  results  that  we  have  obtained  as  a  guide  to  the 
general  problem  of  waves  propagated  in  an  arbitrary  direction.  To 
describe  the  direction  of  propagation,  imagine  a  unit  vector  along  the 
wave  normal.  Let  the  x,  y,  and  z  components  of  this  unit  vector  be 
I,  m,  n.  These  quantities  are  often  called  direction  cosines,  for  it  is 
obvious  that  they  are  equal  respectively  to  the  cosines  of  the  angles 
between  the  direction  of  the  wave  normal  and  the  #,  y,  z  axes.  Then  in 

place  of  the  quantity  sin  2wvlt ),'  or  similar  expressions,  appearing  in 

Eqs.  (2.1),  (2.2),  and  (2.3),  we  must  use  the  expression 


sin  2irAt  -  --J-^-l-  -I-  (2.8) 

Let  us  verify  the  fact  that  Eq.  (2.8)  represents  the  desired  plane  wave. 
At  time  t,  the  expression  (2.8)  is  zero  when 

rt     /.       Ijc  +  My  +  nz\  vv        .   , 

2irv(t  —      — -      —I  =  TT  X  an  integer, 

or 

Ix  +  my  +  nz  =  —  (-  X  — 2")  ~*~  **"  ^'^ 

Now 

/j  +  ra?/  +  nz  =  a  (2.10) 

is  the  equation  of  a  plane  whose  normal  is  a  vector  with  components 
proportional  to  Z,  rrc,  n,  and  whose  perpendicular  distance  from  the  origin, 
measured  along  the  normal  drawn  through  the  origin,  is  a.  Thus  the 
surfaces  given,  by  putting  different  integers  in  Eq.  (2.9),  are  a  series  of 
equidistant  parallel  planes  with  normal  I,  m,  n,  the  distance  apart  being 
$v/v,  and  the  distance  from  the  origin  increasing  linearly  with  the  time, 
with  velocity  v.  This  is  what  we  should  expect  for  the  zeros  of  a  traveling 
wave  of  wave  length  X  =  v/v,  so  that  the  zeros  come  half  a  wave  length 
apart. 

By  superposing  traveling  waves  of  the  nature  of  Eq.  (2.8),  we  can  set 
up  the  standing  waves  that  we  wish.     We  must  superpose  eight  waves, 


SBC.  2]  DE  BYE'S  THEORY  OF  SPECIFIC  HEATS  229 

having  all  eight  possible  combinations  of  ±  signs  for  the  three  terms 
Ix,  my,  nz.  One  of  the  many  typos  of  standing  waves  which  wo  can  sol, 
up  in  this  way  has  the  form 


A    .     0         .     2-irvlx    .     Zirvmy    .     Zirvnz  f        . 

A  sm  2wt  sin  -  sin  -    —  -  sin  -  >  (2.  11) 

V  V  V 

and  this  proves  to  be  the  one  that  we  nood.  Wo  impose  the  boundary 
condition  that  the  displacement  bo  zero  whon  x  =  0,  x  =  X,  y  =  0, 
y  =  }r,  z  —  0,  z  =  Z.  The  conditions  at  a*  =  0,  y  =  0,  z  —  0,  arc*  auto- 
matically satisfied  by  the  function  we  have  chosen  in  Eq.  (2.11).  To 
satisfy  those  at  x  =  X,  y  —  Y,  z  =  Z,  wo  must  make 

2vlX  _  2vmY  _  2mZ  _  (f>      . 

'   7T     ~~  'SiM  7,       ~~  '9'"  ?>       ~  **'  v^'l^y 

where  s^,  sy,  sa  are  integers.     From  Eq.  (2.12),  wo  have 

I  =  «-TO-V'         m  =  5^ov;         w  =  's*o^>         whrro         X  =     •       (2.13) 

L\  £tl  &/J  V 

Since  I,  m,  n  arc*  tho  components  of  unit  voctor,  we  must  have 

P  +  m-  +  n-  =  1, 


or 

K\2       /  •  \2       /  •  V21/\V2 
j)  +  (T)  +\t)  M  =  '•  (2-u) 

Equation  (2.14)  can  be  used  to  find  the  allowed  wave  lengths,  in  terms 
of  the  integers  sx,  sy,  sz: 

1  (2.15) 


or 

i         //TV      /  o  \2      7~TV 

(2.16) 

We  can  now  introduce  the  condition  demanded  by  the  atomic  nature; 
of  the  medium.  We  shall  do  this  only  for  the  simplest  case  of  a  simple 
cubic  lattice,  but  similar  results  hold  in  general.  Let  the  atoms  bo  spaced 
with  lattice  spacing  d,  such  that  X  =  JV*d,  Y  =  Nyd,  Z  =  Nzd,  ami 

NJfJf.  =  N  (2.17) 

is  the  total  number  of  atoms  in  the  crystal.     We  assume  as  the  condition 
for  the  maximum  overtone  that  the  minimum  distance  between  nodes, 


230  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIV 

along  any  one  of  the  three  axes,  is  d.  That  is,  considering  for  instance  the 
x  direction  and  referring  to  Eq.  (2.11),  we  assume  that  increasing  x  by  the 
amount  d  increases  the  argument  of  the  sine,  or  2irvlx/v^  by  TT.  Expressed 
otherwise,  this  states  that  for  the  maximum  overtone  we  must  have 
sx  =  X/d,  and  nimilarly  sy  =  F/d,  sg  =  Z/d.  This  means  that  all  values 
of  sx,  sv,  sz  are  possible  up  to  the  values  sx  =  Nx,  sv  =  Nv,  sz  =  Ng.  Wo 
can  visualize  the  situation  most  easily  by  considering  a  three-dimensional 
space  in  which  sx,  suy  sz  are  plotted  as  three  rectangular  coordinates. 
Then  the  integral  values  of  sx,  sy,  sz,  which  represent  overtones,  form  a 
lattice  of  points,  one  per  unit  volume  of  the  space.  The  overtones 
allowed  for  an  atomic  crystal  are  represented  by  the  points  lying  between 
sx  =  0,  sx  —  Nx,  sv  =  0,  sy  =  Ny,  sz  =  0,  sx  =  Nz.  The  volume  of  space 
filled  with  allowed  points  is  thus  NXNVNZ  =  N,  and  since  there  is  ono 
point  per  unit  volume  there  are  N  allowed  overtones.  In  place  of  using 
this  three-dimensional  space,  it  is  often  more  convenient  to  use  what  is 
called  a  reciprocal  space.  This  is  one  in  which  sx/X,  sy/Y,  sz/Z  are 
plotted.  The  allowed  points  in  the  reciprocal  space  then  form  a  lattice 
with  spacings  I/ A",  1/Y,  l/Z,  so  that  there  is  one  point  in  volume 
l/XYZ  =  1/F,  if  F  =  XYZ  is  the  volume  of  the  crystal.  For  the  maxi- 
mum overtone  we  have  sx/X  =  sy/Y  =  sz/Z  =  1/d,  so  that  the  allowed 
overtones  fill  a  cube  of  volume  1/d3,  or  the  reciprocal  of  the  volume  of  unit 
cell  in  the  crystal.  The  number  of  allowed  overtones,  given  by  the 
volume  (l/dtj)  divided  by  the  volume  (1/F)  per  allowed  overtone,  is 
F/d3  =  N,  as  before.  It  is  plain  why  this  space  is  called  a  reciprocal 
space,  since  distances,  volumes,  etc.,  in  it  are  reciprocals  of  the  cor- 
responding quantities  in  ordinary  space. 

We  have  so  far  omitted  discussion  of  the  fact  that  we  have  both  longi- 
tudinal and  transverse  vibrations.  For  a  single  traveling  wave  like 
Eq.  (2.8),  there  are  of  course  three  possible  modes  of  vibration,  one 
longitudinal  along  the  direction  Z,  m,  n,  and  two  transverse  in  two  direc- 
tions at  right  angles  to  this  direction.  The  longitudinal  wave  will  travel 
with  velocity  vi,  the  transverse  ones  with  velocity  vt.  We  can  superpose 
eight  longitudinal  progressive  waves  to  form  a  longitudinal  standing 
wave,  and  by  superposing  transverse  progressive  waves  we  can  form  two 
transverse  standing  waves.  Three  standing  waves  can  be  set  up  in  this 
way  for  each  set  of  integers  sx,  sy,  sz.  These  three  waves  will  all  corre- 
spond to  the  same  wave  length,  according  to  Eq.  (2.16),  but  to  different 
frequencies,  according  to  Eq.  (1.1).  Considering  the  three  modes  of 
vibration,  there  will  be  in  all  3N  allowed  overtones,  just  the  same  as  in 
the  theories  of  Dulong-Petit  and  Einstein,  discussed  in  Chap.  XIII, 
Sec.  3. 

From  Eqs.  (1.1)  and  (2.16),  we  can  now  set  up  the  frequency  distribu- 
tion, or  spectrum,  as  it  is  often  called  from  the  optical  analogy,  of  our 


SBC.  2]  DE BYE'S  THEORY  OF  SPECIFIC  HEATS  231 

oscillations.     We  have  at  once 


In  our  reciprocal  space,  where  sx/X,  sy/Y,  sz/Z  are  the  three  coordinates, 


the  quantity  V(W^)M:T^7F)"r+  (sz/Z)2  is  simply  the  radius  vector, 
which  we  may  call  r.     Thus  we  have 


the  frequency  being  proportional  to  the  distance  from  the  center.  Now 
we  can  easily  find  the  number  of  overtones  whose  frequencies  lie  in  the 
range  dv  of  frequencies.  For  the  points  in  the  reciprocal  space  represent- 
ing thorn  must  lie  in  the  shell  between  r  and  r  +  dr,  where  r  is  given  by 
Eq.  (2.19).  This  shell  has  the  volume  4wr2  dr,  or  32-jrv*  dv/v\  Only 
|X)sitivo  values  of  the  integers  sx,  sy,  $z  are  to  be  used,  however,  so  that  we 
have  overtones  only  in  the  octant  corresponding  to  all  coordinates  being 
positive.  This  means  that  the  part  of  the  shell  containing  allowed 
overtones  is  one  eighth  of  the  value  above  or  4irv2  dv/tf.  We  have  seen 
that  the  number  of  overtones  per  unit  volume  in  the  reciprocal  space  is 
V.  Thus  the  number  of  allowed  overtones  in  dv,  for  one  direction  of 
vibration,  is 


dN  =  .  v.  (2.20) 

Considering  the  three  directions  of  vibration,  the  number  of  allowed 
overtones  in  dv  is 


dN  =  ±Trv*dvV(  ~  +  -I-  (2.21) 

\vi       vt/ 

Formulas  (2.20)  and  (2.21)  hold  only  when  sx/X,  sv/Y,  sz/Z  are  less  than 
1  /d;  that  is,  when  the  spherical  shell  lies  entirely  inside  the  cube  extending 
out  to  sx/X  —  ±l/d,  etc.  For  larger  values  of  the  frequency,  the  shell 
lies  partly  outside  the  cube,  so  that  only  part  of  it  corresponds  to  allowed 
vibrations.  It  is  a  problem  in  solid  geometry,  which  we  shall  not  go  into, 
to  determine  the  fraction  of  the  shell  lying  inside  the  cube.  When  this 
fraction  is  determined,  we  must  multiply  the  formula  (2.20)  by  the 
fraction  to  get  the  actual  number  of  allowed  states  per  unit  frequency 


range.     In  Fig.  XIV-1  we  plot  the  quantity     y-j->  tbe  number  of 

vibrations  per  unit  volume  per  unit  frequency  range,  computed  in  this 
way,  for  one  direction  of  vibration.     The  curve  starts  up  as  a  parabola, 


232 


INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIV 


(1  \dN      /43T\ 
,r  J-J-  «=  ( -y  ]v2j  but  starts  down  when  v  =  v/2d,  the  point  where  the 

sphere  is  inscribed  in  the  cube,  and  reaches  zero  at  v  =  \/3v/2d,  where 
the  sphere  is  circumscribed  about  the  cube.  The  area  under  the  curve  of 
Fig.  XIV-1  of  course  is  N/V,  where  N  is  the  total  number  of  atoms. 
When  we  consider  both  types  of  vibration,  the  longitudinal  and  the 
transverse,  we  must  superpose  two  curves  like  Fig.  XIV-1,  with  different 
scales  on  account  of  the  difference  between  vi  and  vt.  This  results  in  a 
curve  similar  to  Fig.  XIV-2,  having  two  peaks,  the  one  at  lower  fre- 
quencies corresponding  to  the  transverse  vibration,  which  has  a  lower 
velocity  than  the  longitudinal  vibration. 

In  Fig.  XIV-2  we  have  a  representation  of  the  frequencies  of  vibration 
of  an  clastic  solid,  under  tho  assumption  that  the  waves  are  propagated 
as  in  an  isotropic  solid,'  the  velocity  of  propagation  boing  independent 


§13 


05 


10 


15 


FIG.  XIV-1.  —  Number  of  vibrations  of  one  direction  of  polarization,  per  unit  frequency 
range,  in  a  simple  cubic  lattice  with  lattice  spacing  d,  and  constant  velocity  of  propagation  v. 

of  direction  and  wave  length,  but  the  number  of  overtones  being  limited 
by  the  conditions  that  the  maximum  values  of  sx/X,  sy/Y,  sz/Z  are  1/d, 
where  d  is  the  interatomic  spacing.  This  is  tho  condition  appropriate  to 
a  simple  cubic  arrangement  of  atoms,  an  arrangement  which  does  not 
actually  exist  in  the  real  crystals  of  elements.  It  is  not  hard  to  make 
the  changes  in  the  conditions  that  are  necessary  for  other  types  of  struc- 
ture, such  as  body-centered  cubic,  face-centered  cubic,  and  hexagonal 
close-packed  structures,  which  will  be  discussed  in  a  later  chapter.  The 
general  situation  is  not  greatly  changed.  We  can  still  describe  a  wave  by 
the  three  integers  sx,  syj  sz  and  the  frequency  is  still  given  by  Eq.  (2.19), 
in  terms  of  the  radius  vector  in  tho  reciprocal  space.  Thus  the  number  of 
overtones  in  dv  is  still  given  by  Kq.  (2.20)  or  (2.21),  provided  the  fro- 
quency  is  small  enough  so  that  the  spherical  shell  lies  entirely  within  the 
allowed  region  in  the  reciprocal  space.  The  only  difference  comes  in  the 
shape  of  this  allowed  region.  Instead  of  being  a  cube,  it  can  be  shown 
that  the  region  takes  the  form  of  various  regular  polyhedra,  depending 


SBC.  2] 


DE BYE'S  THEORY  OF  SPECIFIC  HEATS 


233 


on  the  crystal  structure.  These  polyhedra,  which  are  often  called  zones 
or  Brillouin  zones,  are  important  in  the  theory  of  electronic  conduction 
in  metals  as  well  as  in  elastic  vibrations.  The  volumes  of  those  zones  in 
each  case  are  such  that  they  allow  just  N  vibrations  of  each  polarization. 
The  zones  for  the  three  crystal  structures  mentioned  resemble  each  other 
in  that  they  are  more  nearly  like  a  sphere  than  the  cubical  zone  of  the 
simple  cubic  structure.  That  is,  the  radii  of  the  inscribed  and  circum- 
scribed spheres  are  more  nearly  equal  than  for  a  cube.  That  means  that 

the  region  in  which  the  curve  of  ( -T?  )~7~  ^s  falling  from  its  maximum  to 


FIG.  XIV-2. — Number  of  vibrations  per  unit  fit-quonry  range,  in  u  simple  cubic  lattice 
with  constant  velocity  of  propagation.  It  is  assumed  that  the  \elocity  of  the  longitudinal 
wave  is  twice  that  of  the  transverse  waves.  Dotted  curve  indicates  Debye's  assumption. 

zero  is  more  concentrated  than  in  Fig.  XI  V-l,  and  corresponds  to  a  higher 
maximum  and  more  precipitate  fall.  If  the  zone  were  a  sphere  instead  of 
a  polyhedron,  the  fall  would  be  perfectly  sharp,  as  shown  by  the  dotted 
line  iii  Fig.  XIV-2,  the  distribution  being  given  by  a  parabola  below  a 
certain  limiting  frequency  j>niax,  and  falling  to  zero  above  this  frequency. 
The  calculation  which  we  have  carried  out  in  this  section  has  been 
limited  in  accuracy  by  our  assumption  that  the  velocity  of  propagation  of 
the  elastic  waves  was  independent  of  direction  and  of  wave  length. 
Actually  neither  of  these  assumptions  is  correct  for  a  crystal.  Even  for 
a  cubic  crystal,  the  elastic  properties  are  more  complicated  than  for  an 
isotropic  solid  and  the  velocity  of  propagation  depends  on  direction. 


234  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIV 

And  we  have  stated  that  on  account  of  the  atomic  nature  of  the.  material 
the  velocity  depends  on  wave  length,  when  the  wave  length  becomes  of 
atomic  dimensions.  These  limitations  mean  that  the  frequency  spectrum 
which  we  have  so  far  found  is  not  vory  close  to  the  truth.  Nevertheless, 
our  model  is  good  enough  so  that  valuable  results  can  be  obtained  from 
it,  and  we  go  on  now  to  describe  the  approximations  made  by  Debye, 
loading  to  the  specific  heat  curve  known  by  his  name. 

3.  Debye's  Theory  of  Specific  Heats. — Debye's  approximation  con- 
sists in  replacing  the  actual  spectral  distribution  of  vibrations  by  the 
dotted  line  in  Fig.  XIV-2.  That  is,  he  assumed  that  dN  is  given  by  Eq. 
(2.21)  as  long  as  v  is  less  than  a  certain  v^^  and  is  zero  for  greater  v's.  It 
is  obvious  that  this  is  not  a  very  good  approximation.  Nevertheless,  it 
reproduces  the  form  of  the  correct  distribution  curve  at  low  frequencies 
and  has  the  correct  behavior  of  predicting  no  vibrations  above  a  certain 
limit.  To  find  the  proper  j>max  to  use,  Debye  simply  applied  the  condi- 
tion that  the  area  under  his  dotted  curve,  giving  the  total  number  of 
allowed  overtones,  must  be  3N  to  agree  with  the  correct  curve.  That  is, 
he  assumed 


*  4.    (l     J.    2\    f"""    2 

,-    —   4x1  — ,  +  -;  I   I  V 

\»i     Wo 


2N 

V 


from  which 

_/9A* 1       .  .    n 

"-"'  ~     4xF  T~2      '  (AA> 

\         vl  +  v*J 

In  terms  of  the  assumed  frequency  distribution  and  the  formula  (1.3), 
we  can  now  at  once  write  down  a  formula  for  the  specific  heat.     This  is 


J^ 


The  integration  in  Eq.   (3.2)  cannot  be  performed  analytically.     It  is 
worth  while,  however,  to  rewrite  the  expression  in  terms  of  a  variable 

x  "  EF'       with       Xn  **  TT'  (3'3) 

We  then  have 

1     CXQ        ~i»x 

.dx.  (3.4) 


SEC.  3]  DEBYE'S  THEORY  OF  SPECIFIC  HEATS  235 

It  is  customary  to  define  a  so-called  Dobye  temperature  6/>  by  the 
equation 

9*  =  *!=*•  (3.5) 

Then  we  have 

1         T 

--  =  —  (3.6) 

XQ          00 

so  that  Eq.  (3.4)  gives  the  specific  heat  in  terms  of  T/&D,  the  ratio  of  tho 
actual  temperature  to  the  Dobye  temperature.  That  means  that  tho 
specific  heat  curve  should  bo  the  same  for  all  substances,  at  temperatures 
which  are  tho  same  fraction  of  the  corresponding  Dobye  temperatures. 

Whon  integrated  numerically,  the  function  (3.4)  proves  to  bo  not 
unlike  an  Einstein  specific  heat  curve,  except  at  low  temperatures.  To 
facilitate  calculations  with  tho  Dobye  function,  wo  give  in  Table  XIV-1 

TABLE  XIV-1. — SPECIFIC'  HEAT  AS  A  FUNCTION  OF  .r,,  =  O/>/7T,  ACCORDING  TO 

DEBYE'S  THEORY 

XQ  C\ 

0  5.955 

1  5  670 

2  4  918 

3  3  948 

4  2  996 

5  2  197 

6  1  582 

7  1  137 

8  0  823 

9  0.604 

10  0  452 

11  0.345 

12  0  267 

13  0  211 

14  0  169 

15  0  137 

16  0  113 

17  0.0945 

18  0  0796 

19  0  0677 

20  0.0581 

A  more  extended  table  will  be  found  in  NeniHt,  "Die  GiundltiKcn  des  neucn  Waiinenataes  " 

the  specific  boat  per  mole,  calculated  by  Debye's  theory,  as  a  function  of 
XQ.  We  also  give  in  Fig.  XIV-3  a  graph  of  the  Debye  specific  heat  curve 
as  a  function  of  temperature.  We  can  easily  investigate  the  limit  of  low 
temperatures  analytically.  If  T  «  6z>,  we  have  XQ  >  >  1.  Then, 
approximately,  we  can  carry  the  integration  in  Eq.  (3.4)  from  0  to  oo, 
since  the  integrand  becomes  very  small  for  large  values  of  x  anyway.  It 


236  INTRODUCTION  TO  CHEMICAL  PHYSICS          (CHAP.  XIV 

can  be  shown1  that 

f  *        ~4*x  A 

(3.7) 


f  *  4/#  J. 

Jo  ^lAr  ^4> 


Then  we  have,  for  low  temperatures 

r   -  12, 

Cr  =  -=-n 


IF 


(3.8) 


From  Eq.  (3.8),  we  see  that  the  specific  heat  should  be  proportional  to 

6 1 


fX 

~o 


02       04       06      08       10        12       1-1        K> 

T/0 

FIQ.  XIV-3. — Specific  heat  of  a  solid  as  a  function  of  the  temperature,  according  to  Debyc's 

theory. 

the  third  power  of  the  absolute  temperature,  for  low  temperatures.     This 
feature  of  Debye's  theory  proves  to  be  true  for  a  variety  of  substances. 

TABLE  XIV-2. — OBSERVED  SPECIFIC  HEAT  OF  ALUMINUM,  COMPARED  WITH  DEBYE'S 

THEORY 


1 

T,  °abs. 

CP  observed 

Cv  observed 

Cv,  Debye 

54.8 

1.129 

1.127 

1.11 

70.0 

1.856 

1.851 

1  88 

84.0 

2.457 

2.446 

2  51 

112.4 

3.533 

3.502 

3  54 

141.0 

4.239 

4.183 

4.23 

186.2 

4.932 

4.833 

4  87 

257.5 

5.558            ;              5.382 

5.35 

278  9 

5.698             !               5.499 

5.42 

296  3 

5.741 

5  526 

5  48 

Data  for  this  table  are  taken  from  the  article  by  Eucken,  in  "Handbuch  der  Experimentalphysik,' 
Vol.  8,  a  useful  reference  for  the  theory  and  experimental  discussion  of  specific  heats. 

1  See  Debye,  Ann.  Physik,  39,  789  (1912). 


SEC.  3] 


DEBYE'S  THEORY  OF  SPECIFIC  HEATS 


237 


Considering  the  crudeness  of  the  assumptions,  Debye's  theory  works 
surprisingly  well  for  a  considerable  number  of  elements.  Thus  we  give- 
in  Table  XIV-2  the  observed  specific  heat  of  aluminum,  and  the  values 
calculated  from  Debye's  theory,  using  GD  =  385°  abs.  In  Table  XIV-2, 
specific  heats  are  given  in  calories  per  mole.  The  value  of  CV  observed  is 
computed  from  Cp  observed  by  the  use  of  Eq.  (1.14)  of  Chap.  XIII.  It 
is  interesting  to  note  how  much  less  the  difference  CP  —  CV  is  for  such  a 
solid  than  the  value  R  =  1.98  cal.  per  mole  for  a  gas.  Calculations  for 
many  other  substances  show  agreement  with  experiment  of  about  the 
accuracy  of  Table  XIV-2.  We  have  already  pointed  out  the  shortcom- 
ings of  Debye's  treatment,  and  the  remarkable  thing  is  that  it  agrees  as 
well  with  experiment  as  it  does. 

TABLE  XIV-3. — DEB  YE  TEMPERATURES  FOR  ELEMENTS 


Substance 

Oz>,  high  tem- 
perature, °ahs. 

6/>  (T») 

9y>  calculated 

C  (diamond) 

1840 

2230 

Na 

159 

Al 

398 

385 

399 

K 

99 

Fe 

420 

428 

467 

Cu 

315 

321 

329 

Zn 

235 

205 

Mo 

379 

379 

Ag 

215 

212 

Cd. 

160 

129 

168 

Sn 

160 

127 

Pt  

225 

226 

Au 

180 

162 

Pb 

• 

88 

72 

Data  for  this  table,  as  for  Tablp  XIV-2,  are  from  Eucken's  attiele  in  Vol.  8  of  the  "Handbuch  der 
Experimentalphysik. ' ' 

In  Table  XIV-3,  we  give  Dcbye  temperatures  for  a  number  of  elements 
for  which  the  specific  heat  has  been  accurately  determined.  We  give 
three  columns,  and  the  agreement  of  these  three  is  a  fair  indication  of 
the  accuracy  of  Debye's  theory.  The  first  column,  0/,  (high  tempera- 
ture), gives  temperatures  determined  empirically  from  the  specific  heat,  so 
as  to  make  the  agreement  between  theory  and  experiment  as  good  as 
possible  through  the  temperature  range  in  the  neighborhood  of  6/>/2, 
where  the  specific  heat  is  fairly  large.  The  second  column,  6i>(r3),  is  a 
Debye  temperature  determined  to  make  the  T3  part  of  the  curve,  at  very 
low  temperatures,  fit  as  accurately  as  possible.  If  the  Debye  curve 
agreed  perfectly  with  experiment,  these  two  temperatures  would  of  course 
be  equal.  Finally,  in  the  third  column  we  give  9z>  (calc.),  calculated 


238  INTRODUCTION*  W  ffHEM/(&L  PHYSICS         [CHAP.  XIV 

from  the  elastic  constants.  To  find  these,  Eq.  (3.1)  is  used  to  find  vmax 
in  terms  of  the  velocity  of  propagation  of  longitudinal  and  transverse 
waves,  and  Eq.  (3.5)  to  find  QD  in  terms  of  i>inftx.  We  shall  not  go  into 
the  theory  of  elasticity  to  find  the  velocity  of  propagation  in  terms  of  the 
elastic  constants,  but  shall  merely  state  the  results,  in  terms  of  x  the  vol- 
ume compressibility,  <r  Poisson's  ratio,  and  p  the  density.  In  terms  of 
these  quantities,  it  can  be  shown1  that 


Using  Eq.  (3.9),  the  velocity  can  be  found  in  terms  of  tabulated  quantities. 
The  agreement  between  the  columns  in  Table  XIV-3  is  good  enough  so 
that  it  is  plain  that  Debye's  theory  is  a  good  approximation,  but  far  from 
perfect.  It  is  interesting  to  note  from  the  table  the  inyer,s£_rclatioii 
between  compressibility  and  Debye  temperature,  which  can  be  seen 
analytically  "from  Kqs.  (3.1),  (3.5),  and  (3.9)'.  Thus  cITamohd7'a  suHstnnce 
with  extremely  low  compressibility,  has  a  very  high  Debye  temperature, 
while  lead,  with  very  high  compressibility,  has  a  very  low  Dobye  tempera- 
ture. This  moans  that  at  room  temperature  the  specific  heat  of  diamond 
is  far  below  the  Dulong-Petit  value,  while  that  of  lead  has  almost  exactly 
the  classical  value. 

4.  Debye's  Theory  and  the  Parameter  y  —  Debye's  theory  furnishes 
us  with  an  approximation  to  the  frequency  spectrum  of  a  solid,  and  we 
can  use  this  approximation  to  find  how  the  frequencies  change  with 

volume,  arid  hence  to  find  the  parameter  7  =  —  j-.  —  ^  which  is  important 

in  the  theory  of  thermal  expansion,  as  we  saw  in  Chap.  XIII,  Sec.  4. 
According  to  Debye's  theory,  the  frequency  spectrum  is  entirely  deter- 
mined by  the  limiting  frequency  j>max,  and  if  this  frequency  changes,  all 
other  oscillations  change  their  frequencies  in  the  same  ratio.  Thus 
Grimeisen's  assumption  that  y  is  the  same  for  all  frequencies  is  justified, 
and  we  can  set 

d   In    J>max  /  A     .  v 


From  Eq.  (3.1)  we  see  that  the  Debye  frequency  vmax  varies  proportion- 
ally to  the  velocity  of  elastic  waves,  divided  by  the  cube  root  of  the 
volume,  and  from  Eq.  (3.9)  we  see  that  the  velocity  of  either  longitudinal 
or  transverse  waves  varies  inversely  as  the  square  root  of  the  compressi- 
bility times  the  density,  if  we  assume  that  Poisson's  ratio  is  independent 
of  the  volume.  As  we  shall  see  later,  this  assumption  can  hardly  be 

1  For  a  derivation,  see  for  instance,  Slater  and  Frank,  "Introduction  to  Theoretical 
Physics,"  McGraw-Hill  Book  Company,  Inc.,  1933.  Combine  results  of  paragraphs 
109,  110  with  result  of  Prob.  3,  p.  183. 


SEC.  4]  DE 'BYE'S  THEORY  OF  SPECIFIC  HEATS  239 

justified;  Poisson's  ratio  presumably  increases  as  the  volume  increases. 
But  for  the  moment  we  shall  assume  it  to  be  constant.     Then  we  have 

(4.2) 

On  the  other  hand,  the  density  is  inversely  proportional  to  the  volume,  so 
that  we  have 


In  *>max  =  &  In  V  —  £  In  x  +  const., 

7  "  ~6  +  2  d  In  V  (4'3) 

The  compressibility  concerned  in  Eq.   (3.9)  is  that  computed  for  the* 

actual  volume  of  solid  considered;  that  is,  it  is  —  v(  Yp  )  ,  where  V  is 

\  \&r/7 
the  actual  volume,   rather  than  the  volume  at   zero  pressure.     Thus 

In  x  =  -  In  V  - 
and 

/         '' 


din  V 

(4.5) 


To  evaluate  the  derivative  in  Kq.  (4.5),  we  express  P  as  a  function  of  V 
according  to  Eq.  (1.5),  Chap.  XIII: 


=    ?P2; 

from  which,  computing  for  V  = 


This  simple  formula  will  be  compared  with  experiment  in  later  chapters, 
computing  PI  and  P2  both  by  theory  from  atomic  models,  and  by  experi- 
ment from  measurements  of  compressibility.  We  may  anticipate  by 
saying  that  in  general  the  agreement  is  fairly  good,  certainly  as  good  as  we 


240  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XIV 

could  expect  from  the  crudity  of  the  approximations  made  in  deriving 
the  equation. 

In  the  calculations  we  have  just  made,  we  have  assumed  that  Poisson's 
ratio  was  independent  of  volume.  We  have  mentioned,  however,  that 
actually  Poisson's  ratio  increases  with  the  volume.  We  recall  the  mean- 
ing of  Poisson's  ratio:  it  is  the  ratio  of  the  relative  contraction  in  diameter 
of  a  wire,  to  the  relative  increase  in  length  when  it  is  stretched.  For  most 
solids,  it  is  of  the  order  of  magnitude  of  £.  For  a  liquid,  however,  it- 
equals  \.  One  cannot  see  this  directly,  since  a  wire  cannot  be  made  out  of 
a  liquid,  but  the  value  $  indicates  that  a  wire  has  no  change  of  volume 
when  it  is  stretched,  and  this  is  the  situation  approached  by  a  solid  as  it 
becomes  more  and  more  nearly  like  a  liquid.  Now  as  the  volume  of  a 
solid  increases,  either  on  account  of  heating  or  any  other  agency,  it 
becomes  more  and  more  like  a  liquid,  the  atoms  moving  farther  apart  so 
that  they  can  flow  past  each  other  more  readily.  Thus  we  may  infer  that 
Poisson's  ratio  increases,  and  experiments  on  the  variation  of  Poisson's 
ratio  with  temperature  indicate  that  this  is  in  fact  the  case.  If  we  look 
at  Eq.  (3.9),  we  see  that  an  increase  of  Poisson's  ratio  decreases  the 
velocities  of  both  longitudinal  and  transverse  waves.  In  fact,  the  change 
of  vt  is  so  great  that  for  a  liquid,  for  which  a  =  £,  the  velocity  of  transverse 
waves  becomes  zero,  in  agreement  with  our  usual  assumption  that  trans- 
verse waves  cannot  be  propagated  through  a  liquid.  Thus,  when  we 
consider  Poisson's  ratio,  we  find  that  it  provides  an  additional  reason  why 
the  velocity  of  elastic  waves  and  the  Debyc  frequency  should  decrease 
with  increasing  volume.  In  other  words,  it  tends  to  increase  7  over  the 
value  found  in  Eq.  (4.3).  The  exact  amount  of  increase  is  impossible  to 
calculate,  since  the  available  theories  do  not  predict  how  Poisson's  ratio 
should  vary  with  volume,  and  there  are  not  enough  experimental  data 
available  to  compute  the  variation  from  experiment. 

In  considering  thermal  expansion,  we  must  remember  that  Debye's 
theory  is  but  a  rough  approximation,  and  that  really  the  elastic  spectrum 
has  a  complicated  form,  as  indicated  in  Fig.  XIV-2.  If  it  were  not  for 
the  variation  of  Poisson's  ratio  with  volume,  our  discussion  would  still 
indicate  that  the  whole  spectrum  should  shift  together  to  higher  fre- 
quencies with  decrease  of  volume,  since  the  velocities  of  both  transverse 
and  longitudinal  waves  would  then  vary  in  the  same  way  with  volume, 
according  to  Eq.  (3.9).  When  we  consider  the  Poisson  ratio,  however,  we 
see  that  the  velocity  of  the  transverse  waves  should  increase  more  rapidly 
than  that  of  the  longitudinal  waves  with  decreasing  volume,  so  that  the 
shape  of  the  spectrum  would  change.  Thus  Griineisen's  assumption,  that 
the  Vs  should  be  the  same  for  all  frequencies,  is  not  really  justified,  and 
we  cannot  expect  a  very  satisfactory  check  between  his  theory  and 
experiment. 


CHAPTER  XV 
THE  SPECIFIC  HEAT  OF  COMPOUNDS 

In  the  preceding  chapter,  where  we  have  been  considering  the  specific 
heat  of  elements,  there  was  no  need  to  speak  of  internal  vibrations  within 
a  molecule.  In  considering  compounds,  however,  this  is  essential.  A 
real  treatment  of  the  mathematical  problem  of  the  vibrations  is  far  beyond 
the  scope  of  this  book.  Nevertheless,  we  can  take  up  a  simple  one- 
dimensional  model  of  a  molecular  crystal,  which  can  furnish  a  guide  to 
the  real  situation.  Suppose  we  have  a  one-dimensional  chain  of  diatomic 
molecules.  That  is,  we  have  an  alternation  of  two  sorts  of  atoms,  with 
alternating  spacings  and  restoring  forces.  The  vibrations,  transverse 
or  longitudinal,  of  such  a  chain  have  analogies  to  the  vibrations  in  a 
molecular  crystal,  and  yot  they  form  a  simple  enough  problem  so  that  we 
can  carry  it  through  completely.  As  a  preliminary,  we  take  up  the 
simpler  case  of  a  chain  of  like  atoms,  equally  spaced,  analogous  to  the  case 
of  an  elementary  crystal.  This  preliminary  problem  in  addition  will  give 
justification  for  the  discussion  of  the  preceding  chapter,  in  which  we  have 
arbitrarily  broken  off  the  vibrations  of  a  continuum  at  a  given  wave 
length  and  have  said  that  that  resulted  from  the  atomic  nature  of  the 
medium.  Also  it  will  allow  us  to  investigate  the  change  of  velocity  of 
propagation  with  wave  length,  which  we  have  mentioned  before  but  have 
not  been  able  to  discuss  mathematically. 

1.  Wave  Propagation  in  a  One -dimensional  Crystal  Lattice. — Let 
us  consider  N  atoms,  each  of  mass  m,  equally  spaced  along  a  lino,  with 
distance  d  between  neighbors.  Let  the  x  axis  be  along  the  line  of  atoms. 
We  may  conveniently  take  the  positions  of  the  atoms  to  be  at  x  =  rf, 
2d,  3dj  .  .  .  Nd,  with  y  =  0,  z  =  0  for  all  atoms.  These  are  the  equilib- 
rium positions  of  the  atoms.  To  study  vibrations,  we  must  assume  that 
each  atom  is  displaced  from  its  position  of  equilibrium.  Consider  tho 
jth  atom,  which  normally  has  coordinates  x  =  jd,  y  =  z  =  0,  and  assume 
that  it  is  displaced  to  the  position  x  =  jd  +  £/,  y  =  ?;/,  z  =  f „  so  that 
£;>  'nn  f ;  are  the  three  components  of  the  displacement  of  the  atom.  If 
the  neighboring  atoms,  the  (j  —  l)st  and  the  (j  +  l)st,  are  undisplaced, 
we  assume  that  the  force  acting  on  the  jth  atom  has  the  components 

Fx  -  -a*/,        Fv  -  -&,,,        F,  -  -ify,  (1.1) 

each  component  being  proportional  to  the  displacement  in  that  direct  ion 

241 


242  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XV 

and  opposite  to  the  displacement.  We  assume  that  the  force  constant  a 
for  displacement  longitudinally,  or  along  the  line  of  atoms,  is  different 
from  tho  constant  6  for  displacement  transversely,  or  at  right  angles  to 
tho  lino  of  atoms.  This  can  well  happen,  for  in  a  longitudinal  displace- 
ment tho  jth  atom  changes  its  distance  from  its  neighbors  considerably, 
while  in  a  transverse  displacement  it  moves  at  right  angles  to  the  linos 
joining  it  to  its  neighbors  and  stays  at  an  almost  constant  distance  from 
the  neighbors. 

Instead  of  assuming  that  the  force  (1.1)  depends  only  on  the  position 
of  the  jih  particle,  wo  now  assume  that  it  is  really  tho  sum  of  two  forces, 
exerted  on  the  jih  atom  by  its  neighbors  the  (j  —  l)st  and  (j  +  l)st 
atoms.  The  force  exerted  by  the  (j  —  l)st  on  thojth,  whon  the  (j  —  l)st 
is  at  its  position  of  equilibrium,  has  components  (  —  a/2)(f/,  (  —  6/2)77;, 
(  — 6/2)f/.  But  we  must  suppose  that  this  force  doponds  only  on  the 
relative  positions  of  the  two  atoms  in  question,  not  on  their  absolute 
positions  in  space.  Thus  it  must  dcporid  on  tho  differences  of  coordinates 
of  the  j'th  and  (j  —  l)st  atoms,  so  that  the  general  expression  for  the  force 
has  components  (— a/2)(£,  —  £/_i),  (  —  b/2)(rjj  —  r7;_i),  (  — 6/2)(f/  —  f/-i). 
Similarly,  the  force  exerted  on  the  jth  atom  by  the  (j  +  l)st  must  have 
components  (— a/2)(£/  —  £/+i),  (  —  6/2)(tj/  —  T?,+I),  (  — 6/2)(f7-  —  £/+i). 
Combining,  the  total  force  acting  on  the  jth  atom  is 

p    =  _&     +  &(        + 

F,  -  -6f,  +g(f,-i  +  f,+i).  (1-2) 

Using  the  expressions  (1.2)  for  the  force,  we  can  set  up  the  equations 
of  motion  for  the  particles,  using  Newton's  law  that  the  force  equals  the 
mass  times  the  acceleration.  Thus  we  have 

a 
m%j  =  — Q>%J  +  ^ 

,       ,  & 


(1-3) 

where  £/  indicates  the  second  time  derivative  of  {/.  We  now  inquire 
whether  we  can  solve  the  equations  (1.3)  by  assuming  that  the  displace- 
ments form  a  standing  wave  of  the  sort  discussed  in  the  preceding  chapter. 
Let  us  consider  a  longitudinal  vibration,  for  which  £  is  different  from 


SEC.  1]  THE  SPECIFIC  HEAT  OF  COMPOUNDS  243 

zero,  vf  and  f  equal  to  zero,  so  that  only  the  first  equation  of  (1.3)  is 
significant,  and  let  us  assume 

£  =  A  sin  2wvt  sin-^->  (1.4) 

A 

by  analogy  to  the  standing  waves  in  a  continuous  medium.     In  particular, 
for  thojth  particle,  whose  undisplaced  position  x  is  equal  tojW,  wo  assume 


£/  =  A  sin  2irvt  sin  — ~~-  (1.5) 

Then  we  have 

£j*i  =  A  sin  2irvt  sin  2ir(j  ±  1)- 

A 


A     •     o          •  i  • 

=  A  sin  27ml  sin  cos  -^-  ±  cos  —^--  sin  ~-    ; 

A  A  A 


i    >  >«     •     o     ,/«    •     27r;d         2?rd\ 

_i  +  fy+i  =  A  sm  2?r^(  2  sm  -y-  cos  —  1 


„  _N 
(1.6) 


Substituting  Eqs.  (1.5)  and  (1.6)  in  Eq.  (1.3),  we  find  that  the  factor 

f 

A  sin  2wvt  sin  - 
factor,  we  have 


A  sin  27r^  sin  -~-  is  common  to  each  term.     Canceling  this  common 

A 


A     9    9  i  n 

47r2i>2rw  =  —a  +  o  2  cos 

&  A 


aA  27rd\       2a    . 

—  I    1    —    COS    -r-  )     =    -   SI 

m\  X  /        w 


sin*  — ' 

A 

1     /2a    .     ird  ,,  -N 

p  =  7T--V/—  sm  T"  (J-7) 

2?r  \  m          X 

If  the  frequency  ?  and  ihe  wave  length  X  are  related  by  Ea.  (1.7).  the 
values  of  f,  in  Eq.  (1.5)  will  satisfy  the  equations  (1.3).  If  the  velocity 
wcro  constant,  we  should  have  v  =  v/\.  the  frequency  being  inversely 
proportional  to  the  wave  length.  Fronf  Eq.  (1.7)  we  can  see  that  this  is 
the  case  for  long  waves,  or  low  frequencies,  where  we  can  approximate  the 
sine  by  the  angle.  In  that  limit  we  have 

*2awd 
X' 


.  -  X,  -  (1.8) 

a  value  that  can  easily  be  shown  to  agree  with  what  we  should  find  b}' 
Elasticity  theory.     For  higher  frequencies,  however,  since  sin  wd/\  is 


244  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XV 

less  than  ird/\,  we  see  that  the  velocity  of  propagation  must  be  less  than 
the  value  (1.8).  This  is  the  dependence  of  velocity  on  wave  length  which 
was  mentioned  in  the  preceding  chapter. 

NcxtL  we  must  impose.  boundary  conditions  on  our  chain  pf  atoms  as 
WG.  did  wjf-h  t.hp  ftoflfarmona  solid.  We  are  assuming  N  atoms,  with 
undisplaced  positions  at  x  =  d,  2d,  .  .  .  Nd.  We  shall  assume  that  the 
chain  is  held  at  the  ends,  and  to  be  precise  we  assume  hypothetical  atoms 
at  x  =  0,  x  =  (N  +  l)d,  which  are  held  fast.  As  with  the  continuous 
solid,  the  precise  nature  of  the  boundary  conditions  is  without  effect  on 
the  higher  harmonics.  In  Eq.  (1.3),  then,  we  assume  that  the  equations 
can  be  extended  to  include  terms  £0  and  &\r+i,  but  with  the  subsidiary 
conditions 

fo  =  fr+i  =  0.  (1.9) 

The  first  of  these  is  automatically  satisfied  by  the  assumption  (1.5), 
setting,;  =  0.  For  the  second,  we  have  the  additional  condition 

sin  2ir(N  +  1)^  =  0.  (1.10) 

A 

Since  sin  TTS  =  0,  where  s  is  any  integer,  Eq.  (1.10)  can  be  satisfied  by 
2(N  +  1)-  =  s,  where  s  is  an  integer.  0-H) 

A 

The  significance  of  s  is  the  same  as  in  Eq.  (2.7),  Chap.  XIV:  s  =  1  corre- 
sponds to  the  fundamental  vibration,  s  =  2  to  the  first  harmonic,  etc. 
Introducing  the  condition  (1.11),  we  may  rewrite  Eqs.  (1.5)  and  (1.7)  for 
displacement  and  frequency.  For  the  vibration  described  by  the  integer 
s,  we  introduce  a  subscript  s,  obtaining 


A9  sin  2irv,t  sin 


1     I2a   .          ITS  /«  «rt\ 

*  =         8in  (L12) 


On  examining  Eq.  (1.12),  we  see  that  both  £/,  and  vt  are  periodic  in 
s.  Aside  from  a  question  of  sign,  which  is  trivial,  £/,  repeats  itself  when  ,s 
increases  by  (N  +  1)  or  any  integral  multiple  of  that  quantity,  and  vs 
similarly  repeats  itself.  That  is,  all  the  essentially  different  solutions  are 
found  in  the  range  between  s  =  0  and  s  =  N  +  1.  These  two  limiting 
values,  by  Eq.  (1.12),  correspond  to  all  (•'  s  equal  to  zero,  so  that  they  are 
not  really  modes  of  vibration  at  all.  The  essentially  different  values  then 
correspond  to  s  =  1,  2,  ...  N,  just  N  possible  overtones.  This  verifies 


SEC.  1] 


THE  SPECIFIC  HEAT  OF  COMPOUNDS 


245 


the  statement  of  the  previous  chapter  that  the  number  of  allowed  over- 
tones, for  one  direction  of  vibration,  equals  the  number  of  atoms  in  the 
crystal.  It  is  interesting  to  consider  the  periodicity  in  connection  with 
the  reciprocal  space  of  Sec.  2,  Chap.  XIV.  In  that  section,  we  imagined 
s/X  to  be  plotted  as  an  independent  variable.  Hero,  since  X,  the  length 

of  the  chain  of  atoms,  equals  (N  +  l)d,  we  should  plot  ,AT--r  +\~,  —  T 

\J\    ~T~    1  )Cl          A 

as  variable.  Rather  than  plotting  v9  as  a  function  of  this  quantity,  ue 
prefer  to  plot  its  square,  *>*.  This  is  done  in  Fig.  XV-1,  where  the  perio- 
dicity is  clearly  shown.  We  see,  furthermore,  that  the  region  from 
2/X  =  0  to  2/X  =  \/d  includes  all  possible  values  of  1%.  This  funda- 
mental region  corresponds  to  those  described  in  See.  2,  Chap.  XIV.  In 
addition  to  the  sinusoidal  curve  giving  v2,  we  plot  a  parabola  v2  =  02/X2, 
where  v  is  the  velocity  of  propagation  for  long  waves.  Tins  parabola  is 


FIG.  XV-l. — v-  vs.  2/X,  for  one-dimensional  crystal. 


l/d  2/X=S/X 

Dotted  parabola,  case  of  no  dispersion. 


the  curve  which  we  should  find  if  there  were  no  dependence  of  velocity 
on  wave  length  or  no  dispersion  of  the  waves.  We  see  from  Fig.  XV-1 
that  the  effect  of  dispersion  is  to  reduce  tho  frequencies  of  the  highest 
overtones,  compared  to  the  values  which  we  should  find  from  the  theory 
of  vibrations  of  a  continuum.  While  we  cannot  at  once  apply  this  result 
to  the  three-dimensional  case,  it  is  natural  to  suppose  that,  for  instance  in 
Fig.  XIV-1  of  the  previous  chapter,  the  effect  will  be  to  shift  the  peak  of 

1  dtf 


—  -r—  );  the  number  of  overtones  per  unit  frequency  range,  to  lower 

frequencies,  and  at  the  same  time  to  make  the  peak  higher,  so  as  to  keep 
the  number  of  overtones  the  same.  This  is  a  type  of  change  that  makes 
the  curve  resemble  Debye's  assumption  (the  dotted  curve  of  Fig.  XIV-2) 
more  closely  than  before.  Very  few  actual  calculations  of  specific  heat 
have  yet  been  made  using  the  more  exact  frequency  spectrum  that  we 
have  found. 


246 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  XV 


In  Fig.  XV-1  we  have  seen  graphically  the  way  in  which  the  square 
of  the  frequency,  v2,  is  periodic  in  the  reciprocal  space.  It  is  informing 
to  see  as  well  how  the  actual  displacements  ?/,  repeat  themselves.  From 
Eq.  (1.12)  we  see  that  £/,  contains  the  factor  sin  irjs/(N  +  1).  This  of 
course  equals  zero  for  s  =  N  +  1,  and  two  values  of  s  which  are  greater 
than  N  +  1  and  less  than  N  +  1  by  the  same  amount  respectively 
will  havo  equal  and  opposite  values  of  the  sine.  In  other  words,  the 
values  of  the  sine  which  wo  have  when  s  goes  from  0  to  N  +  1  repeat  each 


(a)  s«  N  +  l 


(b)s  = 


N-H 


M/ 


Cc)  s  almost  N-M 


(d)  s  slightly  greater  thanN+1.  Displacements  as  in 
(c)  with  opposite  sign 


Ce)  s  =  "2  (N+1).  Displacements  as  in  (b),with  opposite  sign 


(f)  s  almost  2  (N+1).  Displacements  as  i'n(a),with  opposi4esign 
FIG.  XV-2.  —  Displacements  of  atoms,  for  different  overtones. 


other  iii  opposite  order  as  s  goes  from  N  +  1  to  2(N  +  1).  As  far  as  we 
can  tell  from  the  displacements  of  the  particles,  a  value  of  s/X  greater 
than  !/(/,  in  other  words,  doos  not  correspond  to  a  shorter  value  of  wavo 
length  than  2d  but  to  a  longer  wave  length,  and  when  s/X  equals  2/d  the 
actual  wave  length  is  not  (/  but  infinity,  corresponding  to  110  wave  at  all. 
These  paradoxical  results  are  illustrated  in  Fig.  XV-2,  where  we  show 
TTJS 


curves  of  sin 


indicated  as  functions  of  a  continuous  variable 


SBC.  2]  THE  SPECIFIC  HEAT  OF  COMPOUNDS  247 

j,  with  the  integral  values  of  j  shown  by  dots,  for  several  values  of  s.  It  is 
clear  from  the  figure  that  increase  of  5  does  not  always  mean  decrease  of 
the  actual  wave  length,  but  that  there  is  a  definite  minimum  wave  length 
for  the  actual  disturbance,  oqual  as  we  have  previously  stated  to  twice  the 
interatomic  spacing.  From  the  fact  which  we  have  pointed  out  that  the 
range  of  s/X  from  \/d  to  2/d  repeats  the  range  from  0  to  1/d  in  opposite 
order,  we  understand  why  the  curve  of  i>2  vs.  s/X,  in  Fig.  XV-1,  has  a 
maximum  for  s/X  =  1/d,  falling  to  zero  again  for  2/d. 

2.  Waves  in  a  One  -dimensional  Diatomic  Crystal.  —  In  the  preceding 
section  we  havo  seen  that  the  atomic  nature  of  a  one-dimensional  mon- 
atomic  crystal  lattice  leads  to  a  dependence  of  the  velocity  of  elastic 
waves  on  wave  length  and  to  a  limitation  of  the  number  of  allowed  over- 
tones to  the  number  of  atoms  in  the  crystal.  Now  wo  attack  our  real 
problem,  the  vibrations  of  a  diatomic  one-dimensional  crystal,  using 
analogous  methods.  Wo  assume  each  molecule  to  have  two  atoms,  one 
of  mass  m,  the  other  of  mass  m'.  Lot  there  bo  N  molecules,  and  in 
equilibrium  let  the  atoms  of  mass  m  bo  at  the  positions  X  =  d,  2rf,  •  •  • 
Nd,  and  those  of  mass  m'  at  x  =  d  +  rf',  2ii  +  d',  -  •  •  Nd  +  dr,  whoro 
d'  is  loss  than  d.  Wo  formulate  only  tho  problem  of  longitudinal  vibra- 
tions, understanding  that  tho  transverse  vibrations  can  bo  handled  in  a 
similar  way.  By  analogy  with  Sec.  1,  wo  assume  that  tho  foroos  on  oach 
atom  come  from  its  two  neighboring  atoms.  Those  neighboring  atoms 
are  both  of  tho  opposite  typo  to  the  one  we  are  considering,  but  are  at 


-  e  -  9  -  e  -  0  -  e  -  0  - 

XJ-1  XJ-!  XJ  XJ  Xj+l  X'j*' 

Flo.  XV-3.  —  Arrangement  of  atoms  in  ono-dimeiiHioritil  diatomic  molc-culnr  lattice. 

different  distances,  one  being  at  distance  d'  (in  tho  same  molecule)  the 
other  at  distance  d  —  dr  (in  an  adjacent  molecule).  The  arrangement  of 
atoms  will  be  clearer  from  Fig.  XV-3.  We  assume  two  force  constants: 
a  for  tho  interaction  between  atoms  in  different  molecules,  a1  for  inter- 
action between  atoms  in  tho  same  molecule.  Thus  the  equations  of 
motion  are 


m'%  =  -a(tf  -  UO  ~  a'U,'  -  fc),  (2.1) 

where  £/,  {J  represent  the  displacement  of  the  atoms  of  mass  m  and  m' 
respectively  in  the  jth  molecule. 

To  solve  Eq.  (2.1),  we  assume  sinusoidal  standing  waves  for  both  types 
of  atoms,  but  with  different  phases: 


248  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XV 

£;  =  A  sin  2Trvt  sin          » 

A 


£  =  /T  sin  27r^  sin  +  B'  sin  2r*<  cos          -  (2.2) 

A  A 

Substituting  Eq.  (2.2)  in  Eq.  (2.1),  we  find  that  the  factor  sin  2irvt  cancels 
from  all  forms.     The  remaining  equations  are 

[-4irVffi  +  (a  +  a')]  A  sin  ^~  =  .4'  a  sin  2*0'  -  1)^  +  a'  sin  ^ 

A  L  A  A 


A      J 

cos         ' 


[-4irVm'  +  (n  +  a')l/T  sin  +  B'  cos 


X       '  X    / 

(1 


=  A]  a  si 

I 


sin  2x(j  +  1)     +  «'  sin 


A  A 


(2.3) 


In  Eqs.  (2.3)  wo  expand  the  sines  of  2ir(j  ±  1)     by  the  formulas  for  the 

A 

sines  of  sums  and  differences  of  angles  and  collect  terms  in  sin  —  ^~  and 

A 

cos  -^-     Then  Eqs.  (2.3)  become 


sn 


+  (a  +  n')]  -  X  Ya  cos  -~  +  a'J 


sn,  -~ 


sin  -  ^a  cos          +  «'=  0, 


/A'[-4r*v*m'  +  (a  -f-  a'))  -  ^l(a  cos  ^  +  a'H 

'JB'f-igrVm'  +  (a  +  a')]  -  --ifa  sin  ^H  =  0.     (2. 


sn 

+  cos  B'f-igrVm'  +  (a  +  a')]  -  --ia  sin  =  0.     (2.4) 

If  our  assumptions  (2.2)  arc  to  furnish  solutions  of  Eq.  (2.1),  we  must 
have  Eqs.  (2.4)  satisfied  independent  of  j;  that  is,  for  each  atom  in  tin- 
chain.  The  only  way  to  do  this  is  for  each  of  the  four  coefficients  of  sin 

-^-  or  cos  -~  in  Eq.  (2.4)  equal  to  zero.     We  thus  have  four  simul- 

A  A 

taneous  equations  for  the  four  unknowns  A,  A',  B',  and  v.  Really  there 
are  only  three  unknowns,  however,  for  we  can  determine  only  the  ratios 
A'  /A,  B'  I  A,  and  not  the  absolute  values  of  the  three  amplitudes  A,  A', 


SBC.  2] 


THE  SPECIFIC  HEAT  OF  COMPOUNDS 


249 


Bf.  Thus  we  should  not  expect  to  find  solutions  for  our  equations,  since 
there  are  more  equations  than  unknowns,  but  it  turns  out  that  the  equa- 
tions are  just  so  set  up  that  they  have  solutions.  We  have  the  four 
equations 


J[~4irV7tt  +  (a  +  a')]  -  ;t'(a  cos  -^  +  a'j  -  B'(a  sin  —-)  =  0 

,,/      .     27r<A        ,J  2ird   .      \ 

A'la  Hiu  -r-  1  —  B  '(  a  cos  —  —  h  a  )  =  0, 

A'[-4irVm'  +  (a  +  a')]  -  A\a  cos  ~  +  a')  =  0, 
B'[-4**v*m'  +  (a  +  a')]  -  A\a  sin 


=  0. 

(2.5) 

To  solve  them,  we  first  determine  Br  in  terms  of  A'  from  the  second.  Sub- 
stituting in  the  other  thrco,  wo  have  equations  relating  A  and  A'.  We 
find,  however,  that  tho  third  and  fourth  equations  load  to  tho  same  rosult, 
so  that  there  arc  only  two  independent  equations  for  A  in  terms  of  A'. 
It  is  this  which  makes  solution  possible.  These  two  equations  arc  at 
once  found  to  be 


A[  -  47rVw  +  (a  +  a')]  -  A' 


(  2ird    .      \ 

(a  cos  -— — h  a  1 


,      , 
a  cos  ~  --  h  a 

A 


=  0, 


(2.6) 


From  oach  of  Eqs.  (2.6)  wo  can  solve  for  tho  ratio  A /A'.     Equating  those 
ratios  wo  get  an  equation  for  tho  frequency: 


A\a  cos  ~  +  a'J  -  A'(-Wv*m'  +  (a  +  a')]  =  0. 


1 

r               f  •  2 

9/7                        \  a  SU1  ~ 
af/^c       —  —     1     ft     -L.       ^ 

r)" 

-4rrVm  +  (a  +  a') 

X                              27rd 
a  cos  -r- 

A 

+  «' 

2m'  +  (a  +  a') 

a  cos 


(2.7) 


250  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XV 

From  Eq.  (2.7)  we  have  at  once 

[4*-Vm  -  (a  +  a')][4a-Vm'  -  (a  +  a')] 

/  2nd  ^    ,V   .    (      .    27r<A2 

=  I  a  cos  -r—  +  a'  ]    +  (a  am  —  ) 

=  a2  +  a'5  +  2aa'  cos  ^ 

A 

=  (a  +  a')2  -  2oa'(l  -  cos  ^\ 

=  (a  +  a')2  -  4aa'  sin2^-  (2.8) 

A 

Expanding  the  left  side  of  Eq.  (2.8), 

(27n/)4mm'  -  (2Trv)*(m  +  m')(a  +  a')  +  (a  +  a')2 

=  (a  +  a')2  -  4aa;  sin2  ^-     (2.9) 

A 

Equation  (2.9)  is  a  quadratic  in  (27rv)2.     Solving  it,  we  have 

" 


m 


mm 
where 


/  /\  if"  /  /  /\    ~~1  o 

m  /a  +  a  \          /  m  -f-  m  /a  +  a  \  r       4aa     .   „  ird 

I  _     1   _|_  ^  M  [ i  .....    i  i     _ sin    

'    \     2     /  ~   \  I     mm'    \     2     /  J        mm'          X 

_       .   ,  «[\ 
X/' 


mm 

'  ~  (2'1()) 

m  +  r/i'/fl 


(2.11) 
/ 

4mm' 


(a  +  a')2  (w  +  m')2 


Equation  (2.11)  is  the  desired  equation  giving  frequency  in  terms  of  wave 

length,  for  longitudinal  vibrations  of  a  chain  of  diatomic  molecules. 

As  in  Eq.  (1.7),  we  see  that  the  frequency  depends  on  the  quantity 

sin  — ,  so  that  we  go  through  all  possible  values  when  -  goes  from  zero 

A  A 

to  l/2d.  Thus  there  is  the  same  sort  of  periodicity  seen  in  Fig.  XV-1. 
Furthermore,  when  we  introduce  boundary  conditions  for  a  crystal  of  N 
molecules,  we  find  as  before  that  there  are  N  allowed  overtones  in  this 
fundamental  region  of  reciprocal  space.  In  the  present  case,  however, 
Eq.  (2.11)  has  two  solutions  for  each  value  of  1/X,  coming  from  the  ± 
sign.  That  is,  there  are  two  branches  to  the  curve,  two  allowed  types  of 
vibration  for  each  wave  length,  and  consequently  2N  vibrations  in  all. 
This  is  natural,  for  while  there  are  just  N  molecules,  each  of  these  has 
two  atoms,  so  that  there  are  2N  atoms  and  2N  degrees  of  freedom  for 
longitudinal  vibration.  In  Fig.  XV-4  we  give  curves  of  v2  vs.  2/X,  for 
several  different  values  of  the  constant  K.  From  Eq.  (2.11)  we  see  that 


SEC.  2] 


THE  SPECIFIC  HEAT  OF  COMPOUNDS 


251 


except  for  scale  the  curve  depends  on  only  one  constant  K.  This  constant, 
equals  unity  when  a  =  a',  m  =  m'.  Under  all  other  circumstances  it  is 
easy  to  show  from  its  definition  that  it  is  less  than  unity;  when  a  is  very 
different  from  a',  or  m  very  different  from  m',  or  both,  it  is  very  small. 
Thus  the  limit  K  =  0  corresponds  to  very  unlike  atoms  in  the  molecule, 
with  very  unlike  forces  between  the  atoms  in  the  molecule  and  atoms  in 
adjacent  molecules.  This  is  the  case  of  strong  molecular  binding,  with 
weak  interatomic  forces.  The  limit  K  =  1  corresponds  to  the  ease  where 
the  atoms  are  similar  and  the  binding  within 
the  molecule  is  almost  the  same  as  that  between 
molecules.  This  limit,  for  instance,  would  be 
found  approximately  in  the  case  of  an  ionic 
crystal  like  the  alkali  halides,  where  there  is 
no  molecular  structure  in  the  proper  sense  awl 
where  the  two  typos  of  atom  have  approxi- 
mately the  same  mass.  In  (a),  Fig.  XV-4, 
we  have  the  case  of  strong  molecular  binding. 
Here  one  branch  of  the  curve*  corresponds  to 
low  frequency  vibrations,  going  to  zero  fre- 
quency in  the  limit  of  infinite  wave  length. 
These  vibrations  are  those  in  which  the  mole- 
cules as  a  whole  vibrate,  and  they  are  acous- 
tical vibrations  of  the  same  sort  we  have 
discussed  in  the  preceding  chapter.  The  other 
branch,  however,  corresponds  to  a  much  higher 
frequency,  even  at  infinite  \\avci  length. 
These  vibrations  are  vibrations  of  the  two 


Vd 


0 
(a)  K  «  \ 


2/\ 


/wv 


0     l/d   2/d    2/X 

atoms  in   the  molecule  with  respect  to  each  (c)  K  Shgh%  Less  Thorn  Unity 
other,  almost  exactly  as  the  vibrations  would   FIG.  xv-4.— »>2  vs.  2/x,  for  Hia- 
occur  in   the   isolated   molecule.     These  are        tonne  molecular  lattice. 
often  called  optical  vibrations,  for  as  we  shall  see  later  they  can  be 
observed  in  certain  optical  absorptions  in  the  infrared. 

In  (6),  Fig.  XV-4,  there  is  an  intermediate  case  between  strong  and 
weak  binding.  The  forces  between  molecules  are,  here  not  much  greater 
than  those  within  a  molecule,  and  the  result  is  that  the  optical  branch 
of  the  spectrum  is  not  at  much  greater  frequency  than  the  acoustical 
branch,  Finally  in  (c)  we  show  almost  the  limiting  case  of  equal  atoms 
and  equal  binding.  The  exactly  equal  case  would  correspond  to  a  crystal 
with  2N  equal  atoms  with  a  spacing  of  rf/2.  This  is  the  case  of  Sec.  1,  and 
we  should  expect  the  curve  of  v*  against  2/X  to  correspond  to  Fig.  XV-1, 
except  that  the  first  maximum  of  the  curve  should  come  at  2/d  instead  of 
l/d,  on  account  of  the  spacing  of  d/2.  In  Fig.  XV-4  it  is  shown  how 
this  limit  is  approached.  We  have  already  pointed  out  that  on  account 


252  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XV 

of  periodicity  we  repeat  all  the  overtones  found  between  2/X  =  0  and 
2/X  =  l/d  in  the  next,  period,  from  l/d  to  2/d.  Let  us  then  take  the 
lower  branch  of  the  curve  in  the  range  0  to  1  A/,  and  the  upper  branch  from 
l/d  to  2/rf,  as  shown  by  the  heavy  line  in  (c),  Fig.  XV-4.  If  K  were  exactly 
unity,  these  two  branches  would  join  smoothly,  and  would  give  exactly 
the  curve  we  expect.  When  K  is  slightly  less  than  unity,  so  that  the  two 
types  of  atom  arc  slightly  different  from  each  other  or  there  is  a  slight 
tendency  to  form  diatomic  molecules,  there  is  a  slight  discontinuity  in 
the  curve  at  1/rf,  as  shown  in  the  figure,  but  still  it  is  better  to  regard 
them  as  both  parts  of  a  single  curve.  It  is  clear  from  this  discussion  how 
the  case  of  the  present  section  reduces  to  that  of  Sec.  1  as  the  molecular 
lattice  reduces  to  an  atomic  lattice. 

3.  Vibration  Spectra  and  Specific  Heats  of  Polyatomic  Crystals. — 
The  simplified  one-dimensional  model  which  wre  have  taken  up  in  the  last 
two  sections  is  enough  so  that  we  can  readily  understand  what  happens  in 
the  case  of  real  crystals.  First,  we  consider  definitely  molecular  crystals, 
with  strong  binding  within  the  molecules,  and  relatively  weak  inter- 
molecular  forces.  The  vibration  spectrum  in  this  case  breaks  up 
definitely  into  two  parts.  First  there  is  the  acoustic  spectrum,  connected 
with  vibrations  of  the  molecules  as  a  whole.  This  reduces  to  the  ordinary 
elastic  vibrations  at  very  low  frequencies,  and  extends  to  a  high  enough 
frequency  in  the  infrared  to  include  3N  modes  of  oscillation,  where  there 
are  N  molecules  in  the  crystal.  If  the  intermolecular  forces  are  weak,  so 
that  the  compressibility  of  the  crystal  is  large,  this  limiting  frequency  will 
be  low,  in  the  far  infrared.  Then  there  is  the  optical  spectrum,  connected 
with  vibrations  within  the  molecules.  As  we  see  from  Fig.  XV-4  (a), 
these  vibrations  in  different  molecules  are  coupled  together  to  some 
extent,  so  that  their  frequencies  depend  on  the  particular  way  in  which 
the  vibrations  combine  to  form  a  standing  wave  in  the  crystal.  Never- 
theless, this  dependence  of  frequency  on  wave  length  is  relatively  small,  as 
the  upper  branch  of  Fig.  XV-4  (a)  shows.  The  important  point  is  that 
the  optical  vibrations  in  molecular  crystals  come  at  a  good  deal  higher 
frequency  than  the  acoustical  vibrations,  lying  in  the  part  of  the  infrared 
near  the  visible.  And  these  optical  frequencies  are  only  slightly  different 
from  what  they  would  be  in  isolated  molecules.  This  can  be  seen  from 
our  simple  model.  Thus  in  Eq.  (2.12)  let  a',  the  force  constant  for 
molecular  vibration,  be  large  compared  to  «,  the  force  constant  between 
molecules.  Then  K  will  be  very  small,  and  if  we  neglect  a  compared  to  a' 
we  have  (2irv)2  =  [(m  +  m')/mm']a',  just  the  value  for  a  diatomic  mole- 
cule. We  may  see  this,  for  instance,  from  Chap.  IX,  Eqs.  (2.5)  and 
(4.4),  where  we  obtained  the  same  result.  Of  course,  with  complicated 
molecules,  there  will  be  many  optical  vibration  frequencies  of  the  isolated 
molecule,  and  all  of  these  will  appear  in  the  spectrum  of  the  crystal  with 
slight  distortions  on  account  of  molecular  interaction. 


SEC.  3]  THE  SPECIFIC  HEAT  OF  COMPOUNDS  253 

With  a  spectrum  of  this  sort,  it  is  clear  how  to  handle  the  specific  heat 
of  such  a  molecular  crystal.  The  acoustical  vibrations  can  be  handled 
by  a  Debye  curve,  using  the  elastic  constants,  or  an  empirical  charac- 
teristic temperature,  and  using  the  number  of  molecules  as  N.  Then  we 
add  a  number  of  Einstein  terms  to  take  care  of  the  molecular  vibrations. 
These  again  can  be  found  empirically,  or  they  can  be  deduced  from 
vibration  frequencies  observed  in  the  spectrum.  Those  frequencies 
would  be  exported  to  bo  approximately  the  samo  as  in  tho  molecules,  so 
that  this  part  of  tho  specific  heat  should  agree  with  the  vibrational  part 
of  the  .specific  heat  of  the  corresponding  gas.  In  some  eases,  as  wo  shall 
mention  later,  the;  vibration  frequencies  can  be  found  directly  by  optical 
investigation  of  the  solid.  In  addition  to  the  molecular  vibrations, 
corresponding  to  tho  Einstein  terms,  and  tho  molecular  translation,  cor- 
responding to  tho  Debye  terms,  in  tho  specific  heat  of  the  crystal,  there 
must  be  something  corresponding  to  the  molecular  rotation.  In  most 
solids,  the  molecules  cannot  rotate  but  are  only  free  to  change  their 
orientation  through  a  small  angle,  being  held  to  a  particular  orientation  by 
linear  restoring  forces.  In  their  vibration  spectrum,  this  will  lead  to 
vibwtional  terms  like  the  upper  branch  of  Fig.  XV-4  (a),  and  there  will 
be  additional  Einstein  terms  in  the  specific  heat  coming  from  thi>s  hindered 
rotation.  These  terms  of  course  cannot  bo  predicted  from  tho  properties 
of  the  separate  molecules,  but  ordinarily  must  be  found  empirically  to  fit 
the  observed  specific  heat  curves.  There  are  certain  cases,  on  the  other 
hand,  in  which  the  molecules  at  high  temperatures  really  can  rotate  in 
crystals,  though  at  low  temperatures  they  cannot.  In  such  a  case,  at 
high  temperature,  there  would  be  a  term  in  the  specific  heat  of  the  solid 
much  like  tho  rotational  term  in  a  gas.  At  low  temperatures  whore  the 
rotation  changes  to  vibration,  the  transition  is  more  complicated  than  any 
that  we  have  taken  up  so  far  and  is  really  more  like  a  change  of  phase. 

A  crystal  such  as  those  of  tho  alkali  halides.  formed  from  a  succession 
of  equally  spaced  ions  of  alternating  sign,  is  quite  different,  from  JLho 
molecular  crystals.  The  spectrum,  as  indicated  in  Fig.  XV-4  (<:).  is  much 
more  like  that  of  an  element,  the  distinction  between  the  two  types  of 
ions  being  unimportant.  Thus  we  can  treat  it  bv  methods  of  tho  proced- 
ing  chapter,  using  only  a  Dobye  curve,  but  taking  N  to  bo  tho  total 
number  of  atoms,  not  the  total  number  of  molecules.  This  is  commonly 
done  for  the  alkali  halides,  and  it  is  found  that  one  gets  as  good  agreement 
between  theory  and  experiment  as  for  the  metals.  For  more  complicated 
ionic  crystals,  such  as  carbonates  or  nitrates,  which  aro  formed  of  positive 
metallic  ions  and  negative  carbonate  or  nitrate  radicals,  the  situation  is 
midway  between  the  ionic  and  molecular  cases.  In  CaCOar  for  instance, 
we  should  expect  a  Debye  curve  coming  from  vibrations  of  tho  Cn.4"*-  a.nrl 
CO»—  ions  as  a  whole,  and  also  Einstein  terms  from  the  internal  vibrations 
of  the  carbonate  ions. 


254  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XV 

4.  Infrared  Optical  Spectra  of  Crystals. — Though  we  have  said 
nothing  about  the  interaction  of  molecules  and  light,  we  may  mention 
the  infrared  optical  spectra  of  crystals,  related  to  their  optical  vibrations. 
Light  can  be  emitted  or  absorbed  by  an  oscillating  dipole;  that  is,  by  two 
particles  of  opposite  charge  oscillating  with  respect  to  each  other.  We 
should  then  expect  that  such  vibrations  as  involved  the  relative  motion 
of  different  charges  could  bo  observed  in  the  spectrum.  This  is  never 
tho  case  with  the  acoustic  branch  of  the  vibration  spectrum,  for  there 
we  have  molecules  vibrating  as  a  whole,  and  they  are  necessarily 
uncharged.  But  in  the  optical  branches,  for  instance  with  the  alkali 
halides,  wo  have  just  tho  nooossary  circumstances.  The  rase  1/X  =  0,  in 
the  optical  branch  of  an  alkali  halide,  corresponds  to  a  rigid  vibration 
of  all  the  positive  ions  with  respect  to  all  tho  negative  ions,  so  that  each 
pair  of  positive  and  negative  ions  in  the  crystal  can  radiate  light,  and  all 
these  sources  of  radiation  are  in  phase  with  each  other.  If  such  an 
oscillation  were  excited,  then,  the  crystal  would  emit  infrared  radiation  of 
tho  frequency  of  the  vibration.  Only  the  case  1/X  -  0  corresponds 
to  radiation,  for  if  1/X  were  different  from  zero,  different  parts  of  the 
crystal  would  be  vibrating  in  different  phases  and  tho  emitted  radiation 
from  the  various  atoms  would  cancel  by  interference.  Ordinarily  the 
radiation  is  not  observed  in  emission  but  in  absorption,  for  there  is  no 
available  way  to  excite  this  typo  of  vibration  strongly.  There  is  a  general 
law  of  optics,  however,  stating  that  any  frequency  that  can  bo  omitted  by 
a  system  can  also  be  absorbed  by  the  same  system.  Thus  light  of  this 
particular  infrared  wave  length  can  be  absorbed  by  an  alkali  halide 
crystal.  A  still  further  optical  fact  is  that  light  of  a  frequency  which  is 
very  strongly  absorbed  is  also  strongly  reflected.  Hence  alkali  halide 
crystals  have  abnormally  high  reflecting  power  for  this  particular  wave 
length.  This  is  observed  in  the  experiment  to  measure  residual  rays,  or 
"  Reststrahlon."  In  this  experiment,  infrared  light  with  a  continuous 
distribution  of  frequencies,  as  from  a  hot  body,  i.s  reflected  many  times 
from  the  surfaces  of  alkali  halide  crystals.  For  most  frequencies,  the 
reflection  coefficient  is  so  low  that  practically  all  the  light  is  lost  after 
the  multiple  reflection.  Tho  characteristic  frequencies  have  such  high 
reflecting  power,  however,  that  a  good  deal  of  the  light  of  these  wave 
lengths  is  reflected,  and  the  emergent  beam  is  almost  monochromatic, 
corresponding  to  the  absorption  frequency.  These  beams  which  are  left 
over,  called  residual  rays,  form  a  convenient  way  .of  getting  approxi- 
mately monochromatic  light  in  the  far  infrared. 

By  measurement  of  the  wave  length  of  tho  residual  rays,  for  instance 
with  a  diffraction  grating,  it  is  possible  to  get  a  direct  measurement  of  the 
maximum  frequency  in  the  optical  band  of  the  vibration  spectrum  of  an 
alkali  halide.  If  we  treat  the  spectrum  by  the  Debye  method,  regarding 


SEC.  4]  THE  SPECIFIC  HEAT  OF  COMPOUNDS  255 

the  crystal  as  an  atomic  rather  than  a  molecular  crystal,  this  frequency 
should  agree  approximately  with  tho  Dehye  characteristic  frequency, 
which  can  be  found  from  the  characteristic  temperature  9#.  Such  an 
agreement  is  in  fact  found  fairly  accurately,  as  will  be  shown  in  a  later 
chapter  on  ionic  crystals,  in  which  we  shall  make  comparison  with  experi- 
ment. For  molecular  lattices,  it  is  also  possible  to  got  residual  ray 
frequencies,  in  case  the  molecules  contain  ions  which  can  vibrate  with 
respect  to  each  others.  These  frequencies  have  no  connection  with  the 
Debye  frequencies,  however,  and  they  have  been  much  loss  studied  than 
in  the  case  of  ionic  crystals. 


CHAPTER  XVI 
THE  LIQUID  STATE  AND  FUSION 

For  several  chapters  wo  have  been  taking  up  the  properties  of  solids. 
Earlier,  in  Chap.  XI,  we  discussed  the  equilibrium  of  solids,  liquids,  and 
gases  in  a  general  way  but  without-  using  much  information  about  the 
liquid  or  solid  states.  In  Chap.  XII,  discussing  imperfect  gases  and  Van 
der  Waals1  equation,  we  again  touched  on  the  properties  of  liquids  but 
again  without  much  detailed  study  of  them.  Now  that  we  understand 
solids  better,  we  can  again  take  up  the  problem  of  liquids  and  of  melting. 
The  liquid  phase1  forms  a  sort  of  bridge  between  the  solid  and  the  gaseous 
phases.  It  is  hard  to  treat,  because  it  has  no  clear-cut  limiting  cases,  such 
as  the  crystalline  phase  of  the  solid  at  the  absolute  zero  and  the  perfect 
gas  as  a  limiting  case  of  the  real  gas.  The  best  we  can  do  is  to  handle  it 
as  an  approximation  to  a  gas  or  an  approximation  to  a  crystalline  solid. 
The  first  is  essentially  the  approach  made  in  Van  der  Waals'  equation, 
which  we  have  already  discussed.  The  second,  the  approach  through  the 
solid  phase  and  through  fusion,  is  the  subject  of  the  present  chapter. 

1.  The  Liquid  Phase.--  We  ordinarily  deal  with  liquids  at  tem- 
peratures and  pressures  well  below  the  critical  point,  and  it  is  here  that 
they  resemble  solids.  When  studied  by  x-ray  diffraction  methods,  it  is 
found  that  the  immediate  neighbors  of  a  given  atom  or  molecule  in  a 
liquid  are  arranged  very  much  as  they  are  in  the  crystal,  but  more*  distant 
neighbors  do  not  have  the  same  regular  arrangement.  We  shall  meet  a 
particularly  clear  case  of  this  later,  in  discussing  glass.  A  glass  is  simply 
a  supercooled  liquid  of  a  silicate,  or  other  similar  material,  held  together 
by  bonds  extending  throughout  the  structure,  just  as  in  the  crystalline 
form  of  the  same  substance.  These  materials  supercool  particularly 
easily,  presumably  because  the  atoms  or  ions  of  the  liquid  are  so  tightly 
bound  together  that  they  do  not  rearrange  themselves  easily  to  the 
positions  suitable  for  the  crystal.  Thus  we  can  observe  their  liquid 
phases  at  temperatures  low  enough  so  that  they  take  on  most  of  the 
elastic  properties  of  solids.  They  acquire  rigidity,  resistance  to  torque. 
Nevertheless  they  never  lose  entirely  their  properties  of  fluidity.  A  rod 
of  glass  at  room  temperature,  subjected  to  a  continuous  stress  which  is  not 
great  enough  to  break  it,  will  gradually  deform  or  flow  over  long  periods  of 
time.  The  study  of  materials  which  behave  in  this  way,  showing  both 
fluidity  and  elasticity,  is  called  rheology,  and  it  shows  that  such  a  com- 

256 


SEC.  1]  THE  LIQUID  STATE  AND  FUSION  257 

bination  of  properties  is  very  widespread.  The  fluidity  of  the  glasses 
is  a  result  of  the  fact  that  there  is  no  unique  arrangement  of  the  atoms,  as 
there  is  in  a  perfect  crystal.  Certain  atoms  may  be  in  a  situation  where 
there  are  two  possible  positions  of  equilibrium,  near  to  each  other.  That 
is,  the  atom  is  froe  to  move  from  tho  position  whore  it  is  io  an  adjacent 
hole  in  the  structure,  with  only  a  small  expenditure  of  energy.  Tho 
maximum  of  potential  energy  between  tho  minima  is  closoly  analogous 
to  the  energy  of  activation  in  chemical  reactions,  mentioned  in  C -hap.  X, 
Sec.  3.  In  Fig.  XVI-1  (a)  we  show  schematically  how  the  potential 
energy  acting  on  this  atom  might  appear,  as  we  pass  from  one  position  of 
equilibrium  to  the  other.  There  will  ordinarily  not  be  much  tendency 
for  the  atom  to  go  from  one  position  to  the  other.  But  if  the  material  is 
under  stress  and  one  of  the  positions  tends  to  relieve  tho  stress,  the  other 


(oO 

Irir,.  XVI-1.     Schematic  potential  energy  acting  on  an  atom  in  n  KhiHH.      (a)   Unstressed 
material,  (//)  material  under  stress 

not,  the  energy  relations  will  be  shifted  so  that  the  position  which  relieves 
tho  stress,  as  shown  in  (16),  Fig.  XVI-1,  will  have  lower  energy  than  (26), 
which  does  not.  Then  if  the  atom  is  in  position  (2),  it  will  have;  a  good 
chance  of  acquiring  the  energy  c2,  enough  to  pass  over  the;  potential  hill 
and  fall  to  the  position  (1),  simply  by  thermal  agitation.  By  tho  Max- 
vvell-Boltzmaim  relation,  wo  should  expect  the  probability  of  finding 
an  atom  with  this  energy  to  bo  proportional  to  the  quant  ity  exp  ( —  €2/&77), 
increasing  rapidly  with  increasing  temperature.  On  tho  other  hand,  the 
probability  that  an  atom  in  (1)  should  have  the  energy  t\  necessary  to 
pass  over  the  hill  to  position  (2)  would  contain  tho  much  smaller  factor 
exp  (  —  €\/kT).  The  net  result  would  bo  that  atoms  in  positions  like  (2) 
would  move  to  positions  like  (1),  relieving  the  stress,  but  tho  opposite; 
type  of  transition  would  not  occur.  This  would  amount  to  a  flow  of 
the  material,  if  many  such  transitions  took  place.  Furthermore,  the  rate; 
of  the  process  woujd  be  proportional  to  exp  (  —  e2/M7),  anel  this  woulel 
be  expected  to  be^roportional  to  the  rate  of  flow  under  fixed  stress,  or  to 
the  coefficient  of"  yisoosity.  The*  actual  viscosities  of  glasses  show  a 
dependence  on  temperature  of  this  general  nature,  the  flow  becoming 
very  slow  at  low  temperatures,  so  slow  that  it  is  ordinarily  not  observed 
ut  all.  But  there  is  no  sudden  change  between  a  fluid  and  a  solid  state. 


258  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XVI 

The  glasses  are  particularly  informing  fluids,  because  we  can  observe 
them  over  such  wide  temperature  ranges.  Other  types  of  liquids  do  not 
supercool  to  any  such  extent,  so  that  ordinarily  they  cannot  be  observed 
at  temperatures  low  onough  so  that  they  have  begun  to  lose  their  charac- 
teristically fluid  properties.  We  may  infer,  however,  that  the  process  of 
flow  in  all  liquids  is  similar  to  what  we  have  described  in  the  glasses. 
That  is,  the  atoms  surrounding  a  given  atom  arc  approximately  in  the 
arrangement  correct  for  a  crystal,  but  there  arc  so  many  flaws  in  the 
structure  that  there  are  many  atoms  which  can  move  almost  freely  from 
the  position  where  they  happen  to  bo,  to  a  nearby  vacant  place  where 
they  would  fit  into  the  structure  almost  as  well.  One  rather  informing 
approximation  to  the  liquid  state  treats  it  straightforwardly  as  a  mixture 
of  atoms  and  holes,  the  holes  simply  representing  atoms  missing  from  the 
structure,  and  it  treats  flow  as  the  motion  of  atoms  next  the  holes  into  the 
holes,  leaving  in  turn  empty  holes  from  which  they  move.  The  essential 
point  is  that  the  structure  is  something  like  the  solid,  but  like  an  extremely 
imperfect,  disordered  solid.  And  it  is  this  disorder  that  gives  the  possi- 
bility of  flow.  A  perfect  crystal  cannot  flow,  without  becoming  deformed 
and  imperfect  in  the  very  process  of  flow. 

2.  The  Latent  Heat  of  Melting.—  With  this  general  picture  of  the 
liquid  state,  let  us  ask  what  we  expect  to  find  for  the  thermodynamic 
properties  of  liquids.  We  shall  consider  two  thermodynamic  functions, 
the  internal  energy  and  the  entropy,  and  shall  ask  how  we  expect  these 
quantities  to  differ  from  the  corresponding  quantities  for  the  solid.  This 
information  can  be  found  experimentally  from  the  latent  heat  of  fusion 
Lw,  which  gives  directly  the  change  in  internal  energy  between  the  two 
phases  (for  solids  and  liquids  the  small  quantity  P(Vi  —  V8)  by  which 
this  should  be  corrected  is  negligible),  and  from  the  melting  point  Tm,  for 
the*  entropy  of  fusion  is  given  by  Lm/Tm.  As  a  matter  of  general  orienta- 
tion, we  first  give  in  Table  XVI-1  the  necessary  information  about  a 
number  of  materials.  In  the  first  column  we  give  the  latent  heat  of 
fusion.  We  shall  find  it  interesting  to  compare  it  with  the  latent  heat  of 
vaporization;  therefore  we  give  that  quantity  in  the  next  column,  and  the 
ratio  of  heat,  of  fusion  to  heat  of  vaporization  in  the  third.  Next  wo 
tabulate  the  molting  point  and  finally  the  entropy  of  fusion. 

Referring  to  Table  XVI-1,  let  us  first  consider  the  latent  heat  of  fusion. 
We  observe  that  in  practically  every  case  it  is  but  a  small  fraction  of  the 
boat  of  vaporization.  That  is,  the  atoms  or  molecules  are  pulled  apart 
only  slightly  in  the  liquid  state  compared  with  the  solid,  while  in  the 
vapor  they  are  completely  separated.  Of  course,  this  holds  only  for 
pressures  low  compared  to  the  critical  pressure;  near  the  critical  point,  the 
heat  of  vaporization  reduces  to  zero.  To  be  more  specific,  we  notice 
that  in  the  metals  the  heat  of  fusion  is  generally  three  or  four  per  cent  of 


2]  THE  LIQUID  STATE  AND  FUSION 

TABLE  XVI-1. — DATA  REGARDING  MELTING  POINT 


259 


Lm,  kg.-cal. 
per  mole 

Lv 

Lm 

Z7 

Tm,  °  abs. 

ASm,  cal.  per 
degree 

Metals: 
Na  

0  63 

26  2 

0  024 

371 

1  70 

Mg                    

1  16 

34  4 

0  034 

923 

1  26 

Al      

2  55 

67  6 

0  038 

932 

2  73 

K  

0  58 

21  9 

0  020 

330 

1  72 

Cr          

3  93 

89  4 

0  044 

1823 

2   15 

Mn 

3  45 

G9  7 

0  050 

14'J3 

2  31 

Fe     

3  50 

96  5 

0  037 

1802 

1  97 

Co        

3  GO 

1763 

2  08 

Ni  

4  20 

98  1 

0  043 

1725 

2  44 

Cu  

3  11 

81  7 

0  038 

1357 

2  29 

Zn                  .... 
Ga  
Se        

1  00 
1  34 
1  22 

31   4 

0  051 

(192 
303 
490 

2  32 
4  42 
2  49 

Rb  

0  53 

20  6 

0  020 

312 

1  70 

Ag               ... 

2  70 

G()  4 

0  039 

1234 

2  19 

Cd 

1  40 

°7  0 

0  054 

594 

2  46 

In          
Sn              
Sb          

0  78 
1   72 
4  77 

08  0 
54  4 

0  025 
0  088 

429 
505 
903 

1  82 
3  40 
5  29 

Cs  

0  50 

18  7 

0  027 

302 

1   GO 

Pt          

5  33 

125 

0  043 

2028 

2  03 

Au   
Hg  

3  03 
0  58 

90  7 
15  5 

0  033 
0  037 

1330 
234 

2  27 
2  48 

Ti     :    ::    :. 

Pb 
Bi 

Ionic  substances 
NaF 
NaCl 
KF 
KC1 
KBr 
AgCl 

asr  . 

TlBr 
LiNOs       . 
NaNOs 
KN03 
A«N03 
NaClOa 
NaOH     .  . 
KOH 
K2Cr2O; 
BaCli 
CaCh 
PbCh 
PbBrz 
Pblj 
HgBrz 
HgI2 
Molecular  substances. 
H, 
NO 
H:0 
O2 
A 
NHj        .    . 

0  70 
1   22 
2  51 

7  81 
7  22 
(\  28 
G  41 
2  84 
3   15 
2  18 
4  2(1 
3  '.)!) 
G  00 
3  70 
2  57 
2  70 
5  29 
1  00 
1  01 
8  77 
5  75 
G  03 
5  05 
4  29 
5  18 
4  02 
4  50 

0  028 
0  551 
1  43 
0  096 
0.280 
1  84 

43  0 
40  7 
47  8 

213 
183 
190 
105 
159 

0  22 
3  82 
11  3 
2  08 
1.88 
7  )4 

0  018 
0  02G 
0  053 

0  037 
0  039 
0  033 
0  039 
0  018 

0  13 
0  14 
0  13 
0  05 
0  15 
0  20 

570 
001 
544 

1  205 
1073 
1133 
1043 
Oil 

728 
703 
700 
733 
523 
583 
581 
481 
528 
591 
033 
071 
1232 
1047 
771 
761 
648 
508 
523 

14 
110 
273 
54 
83 
198 

I   32 
2  03 
4  01 

0  19 
0  72 
5  53 
0  15 
4  05 
4  33 
3   10 
0  09 
8  18 
11    G 
6  45 
4  42 
5  72 
9  90 
2  71 
2  54 
13  07 
4  65 
5  77 
7  32 
5  63 
8  00 
9  09 
8  60 

2  0 
5  02 
5  25 
1  78 
3  38 
9  30 

Na                          

0  218 

1  69 

0  13 

63 

3  46 

CO                  

0  200 

1  90 

0  11 

68 

2  04 

HC1 

0  506 

4  85 

0  10 

159 

3  20 

CO* 

1  99 

6  44 

0  31 

217 

9  16 

CH4     

0  224 

2  33 

0  10 

90 

2  49 

HBr       

0  620 

5.79 

0  11 

187 

3  31 

Ch  

1  63 

7  43 

0  22 

170 

9  59 

ecu       

0  577 

8  0 

0  07 

250 

2  30 

CHsOH  

0  525 

0  2 

0  06 

176 

2  08 

CzH6OH     

1   10 

10  4 

0  11 

156 

7  10 

CHsCOOII 

2  64 

20  3 

0  13 

290 

9  21 

C«H«      

2  35 

8  3 

0  28 

278 

8  45 

Data  are  from  Landolt's  Tables.     The  heats  of  vaporization  tabulated  for  alkali  hulide*  are  the 
energies  required  to  break  the  crystal  up  into  ions,  rather  than  into  atoms. 


INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XVI 

the  heat  of  vaporization.  This  receives  a  ready  explanation  in  terms  of 
our  model  of  a  liquid  as  a  crude  approximation  to  a  solid  but  with  many 
holes  or  vacant  places.  Suppose  there  are  three  or  four  chances  out  of  a 
hundred  that  there  will  be  a  vacant  space  instead  of  an  atom  at  a  given 
point  in  the  imperfect  lattice.  Then  the  energy  of  the  substance  will  bo 
a,  corresponding  amount  less  than  if  all  points  wore  occupied,  for  each 
atom  will  have  a  correspondingly  smaller  number  of  neighbors  on  the 
average,  and  the  energy  of  the  crystal  comes  from  the  attraction  of  neigh- 
boring atoms  for  each  other.  This  is  about  the  right  order  of  magnitude 
to  account  for  the  latent  heat  of  fusion,  then,  and  at  the  same  time  it 
indicates  a  density  three  or  four  per  cent  less  for  the  liquid  than  for  the 
solid,  which  is  about  the  right  order  of  magnitude.  From  Table  XVi-1 
we  see  that  the  situation  is  about  the  same  for  the  alkali  halides  (and 
presumably  for  other  ionic  crystals)  as  for  the  metals,  if  we  understand  by 
the  latent  heat  of  vaporization  the  energy  required  to  break  up  the  sub- 
stance into  a  gas  of  ions. 

In  the  case  of  molecular  substances,  the  latent  heat  of  fusion  is  a 
larger  fraction  of  the  latent  heat  of  vaporization,  from  10  to  20  per  cent 
or  even  higher,  so  that  it  seems  clear  that  the  liquid  differs  from  the  solid 
more  in  these  cases  than  with  metals  and  ionic  substances.  In  many  of 
the  molecular  crystals,  the  molecules  are  fitted  together  in  a  regular 
arrangement,  whereas  in  the  liquid  there  is  presumably  more  tendency  to 
rotation  of  the  molecules,  and  they  do  not  fit  so  perfectly.  This  tendency 
would  make  considerable  change  in  the  energy  and  in  the  volume,  and 
would  represent  a  feature  which  is  absent  with  metals,  where  the  atoms 
are  effectively  spherical.  For  example,  in  ice,  as  we  shall  see  later,  the 
triangular  water  molecules  are  arranged  in  a  definite  structure,  each 
oxygen  being  surrounded  by  four  others,  with  the  hydrogens  in  such  an 
arrangement  that  the  dipoles  of  adjacent  molecules  attract  each  other. 
In  liquid  water,  on  the  other  hand,  the  arrangement  is  far  less  perfect 
and  precise,  the  molecules  are  farther  apart  on  centers,  and  one  can 
understand  the  latent  heat  of  fusion  simply  as  the  work  necessary  to 
increase  the  average  distance  between  the  dipoles,  against  their  attrac- 
tions. Presumably  in  all  the  molecular  substances,  it  is  more  accurate  to 
think  simply  of  the  increased  interatomic  distance  as  leading  to  the 
heat  of  fusion,  rather  than  postulating  holes  in  the  structure  as  definitely 
as  one  does  with  a  metal.  However  one  looks  at  it,  the  liquid  is  a  more 
open,  less  well-ordered  structure  than  the  solid,  and  the  latent  heat 
represents  the  work  necessary  to  pull  the  atoms  or  molecules  apart  to  this 
open  structure. 

3.  The  Entropy  of  Melting. — When  we  look  at  the  entropies  of 
melting,  in  Table  XVI-1,  we  see  that  there  is  a  certain  amount  of  regu- 
larity in  the  table.  For  most  of  the  metals,  the  entropy  of  fusion  is 


ftEC. 


TUB 


STATE  AMU  FUS1UJN 


between  two  and  three  calories.  For  the  diatomic  ionic  crystals  it  is 
about  twice  as  much,  so  that  if  we  figure  the  entropy  per  atom  instead  of 
per  molecule  it  is  about  the  same  as  for  the  metals.  As  a  first  step  toward 
understanding  the  entropy  of  molting,  we  might  use  a  rough  argument 
similar  to  that  of  Chap.  XIII,  Soc.  5,  where  we  discussed  the  entropy 
changes  in  polymorphic  transitions.  We  were  them  interested  in  tho 
slope  of  the  transition  curves  between  phases,  but  the  calculation  we  made 
was  one  of  A$/AF,  the  change  of  entropy  between  two  phases  divided  by 
the  change  of  volume,  and  we  assumed  that  the  change  of  entropy  with 
volume  in  going  from  one  phase  to  another  was  the  same  as  in  changing 
the  volume  of  a  single  phase.  Tn  this  case,  using  the  thermodynamic 
relation 


(es\      _ 

\dVjr  7 


(ev\ 
Wi- 


dV 


we  have 


AS 
AV 


thermal  expansion 
compressibility 


(3.1) 


(3.2) 


From  the  relation  (3.2),  and  tho  observed  change  of  volume  on  molting, 
we  can  compute  the  change  of  entropy.  In  Table  XVI-2  we  give  values 
of  volume  of  the  solid  per  mole  (extrapolated  from  room  temperature  to 
tho  melting  point  by  use  of  the  thermal  expansion),  volume4  of  the  liquid 

TABLE  XVI-2.     CALCULATION  OF  ENTKOPY  OK  MELTINCJ 


Molecular 
volume  of 
solid,  cc. 

Molecular 
volume  oi 
liquid 

,.              Thermal             A/S',,, 
expansion       computed 

A&. 
observed 

Na 

24  2 

24  6 

0  4 

21   f>  X  10  *  '       0  13 

I  70 

Mg 

14  6              15  2 

0  6 

7  3                0  36 

1  26 

Al 

10  6 

11  0 

0.4 

6  8 

0.49 

2.73 

K 

46.3       !       47.2 

0  9 

23.0 

0  13 

1.72 

Fo 

7  50 

8   12 

0  6 

3  36 

0  86 

1.97 

A*~ 

10  8 

11  3 

0  f> 

3  7 

0  69 

2.19 

Cd 

13  4 

14  0 

0  6 

9  3 

0.93 

2.46 

NaCl 

29.6 

37  7 

8  1 

12  1 

5.6 

6.72 

KC1 

40  5 

48.8 

8.3 

11   4 

4  0 

6.15 

KBr 

45.0 

56.0 

11  0 

12  6 

4.9 

4  65 

AgCl 

27.0 

29.6 

2  6 

9.9 

2.6 

4  33 

AgBr 

30.4 

33  6 

3  2 

10.4 

3.2 

3  10 

Molecular  volumes  of  the  solid  are  calculated  from  observed  densities  at  room  temperature  (as 
tabulated  in  Landolt's  Tables),  extrapolated  to  the  melting  point  by  using  the  thermal  expansion.  For 
the  ionic  ciystals,  data  on  densities  of  liquids  and  solids  are  taken  from  Lorenz  and  Herz,  X  anory. 
allffem  Chetn.,  145,  88  (1025). 


262  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XVI 

at  the  melting  point  per  mole,  change  of  volume,  thermal  expansion,  and 
AS  as  computed  from  them  by  Eq.  (3.2),  for  a  number  of  solids  for  which 
the  required  quantities  are  known,  and  we  compare  with  the  values  of  A/S 
as  tabulated  in  Table  XVI-1,  which  we  repeat  for  convenience.  We 
see  that  the  calculated  entropies  of  melting  are  of  the  right  order  of 
magnitude,  but  that  in  most  cases  they  are  decidedly  smaller  than  tho 
observed  ones.  In  other  words,  wo  must  assume  that  the  entropy  actu- 
ally increases  in  melting  more  than  we  should  assume  simply  from  the 
change  of  volume,  though  as  a  first  approximation  that  gives  a  useful  and 
partially  correct  picture  of  what  happens.  The  fact  that  the  ratio  of 
change  of  entropy  to  change  of  volume  is  approximately  given  by  this 
simple  picturo  shows  that  the  calculation  of  Chap.  XIII,  Sec.  5,  on  the 
slope  of  transition  lines,  will  apply  roughly  to  the  slope  of  the  melting 
curves,  and  this  is  found  to  be  true  experimentally.  Of  course,  as  with 
transitions  between  solids,  there  is  great  variation  from  one  material  to 
another,  arid  occasional  materials,  of  which  water  is  the  most  conspicuous 
example,  actually  have  a  decrease  of  volume  on  melting,  though  their 
entropy  increases,  so  that  in  such  cases  relation  (3.2)  is  obviously  entirely 
incorrect. 

The  simple  argument  we  have  given  so  far  does  not  give  a  very  ade- 
quate interpretation  of  the  entropy  of  melting,  and  as  a  matter  of  fact  no 
very  complete  theory  is  available.  We  can  analyze  the  problem  a  little 
better,  however,  by  considering  it  more  in  detail.  We  can  imagine  that 
the  entropy  of  the  liquid  should  be  greater  than  that  of  the  solid  for  two 
reasons.  First,  at  the  absolute  zero,  the  solid  has  zero  entropy.  If  we 
could  carry  the  liquid  to  the  absolute  zero  by  supercooling,  however,  we 
should  imagine,  at  least  by  elementary  arguments,  that  its  entropy  would 
be  greater  than  zero.  The  reason  is  that  the  atoms  or  molecules  of  the 
liquid  are  arranged  in  a  much  more  random  way  than  in  the  crystal,  and 
since  entropy  measures  randomness,  this  must  lead  to  a  positive  entropy 
for  the  liquid.  We  shall  be  able  to  estimate  this  contribution  to  the 
entropy  in  the  next  paragraph  and  shall  see  that  while  it  is  appreciable,  it 
is  not  great  enough  to  account  for  nearly  all  of  the  entropy  of  fusion. 
Secondly,  there  aro  good  reasons  for  thinking  that  the  specific  hoat  of  the 
liquid,  at  temperatures  between  the  absolute  zero  and  the  melting  point, 


would  be  greater  than  that  of  the  solid.     Thus  the  integral    I      -=•  dT 

Jo     * 

measuring  the  difference  of  entropy  between  absolute  zero  and  the  melting 
point  will  be  greater  for  the  liquid  than  for  the  solid,  giving  an  additional 
reason  why  the  entropy  of  the  liquid  should  be  greater  than  that  of  the 
solid.  It  is  reasonable  to  think  that  this  effect  is  fairly  large,  and  the 
whole  entropy  of  fusion  can  be  regarded  as  a  combination  of  the  two 
effects  we  have  mentioned. 


SEC.  3]  THE  LIQUID  STATE  AND  FUSION  263 

Let  us  try  first  to  estimate  the  contribution  to  the  entropy  of  fusion  on 
account  of  the  randomness  of  arrangement  of  the  atoms  in  a  liquid.  This 
calculation  can  be  carried  out  at  the  absolute  zero,  and  we  can  get  some- 
thing of  the  right  order  of  magnitude  by  taking  our  simple  picture  of  the 
liquid  as  a  mixture  of  atoms  and  holes.  Let  us  assume  N  atoms  and  Na 
holes,  where  a  is  a  small  fraction  of  unity.  We  may  consider  that  these 
form  a  lattice  of  N(l  +  a)  points  and  may  say  very  crudely  that  any 
arrangement  of  the  N  atoms  and  the  Na  holes  on  these  N(\  +  a)  points 
will  constitute  a  possible  arrangement  of  the  substance  having  the  same, 
lowest  energy  Vo.  By  elementary  probability  theory,  the  number  of 
ways  of  arranging  N  things  of  one  sort,  and  Na  of  another,  in  N(l  +  «) 
spaces,  one  to  a  space,  is 

[JVQ  +«)]!  ,,,. 

W  ~     Nl(Na)\"  (At6) 

Using  Stirling's  formula,   AM  =  (N/e)N  approximately,   the  expression 
(3.3)  becomes 

w  _ 


the  powers  of  N  and  e  canceling  in  numerator  and  denominator.  Now 
we  can  calculate  the  entropy  by  Boltzmann's  relation  S  =  k  In  W,  of 
Eq.  (1.3),  Chapter  III.  We  have  immediately 

S  =  Nk[(l  +  a)  In  (1  +  a)  -  a  In  a].  (3.5) 

In  the  expression  (3.5)  let  us  put  a.  —  0.04,  the  value  which  we  roughly 
estimated  from  the  latent  heat.  Then  calculation  gives  us  at  once 

S  =  O.l&SNk  =  0.33  cal.  per  degree  per  mole.  (3.6) 

The  value  (3.6),  while  appreciable  compared  with  the  values  of  Table 
XVI-1,  which  are  of  the  order  of  magnitude  of  two  or  more  calories 
per  degree,  is  definitely  less,  so  much  less  that  it  cannot  possibly  account 
for  the  whole  entropy  of  fusion.  Let  us  see  what  value  of  a  we  should 
have  to  take  to  get  the  whole  entropy  of  fusion  from  this  term.  If  we  set 
a  =  1,  for  instance,  we  have 

S  =  l.38Nk  =  2.75  cals.  per  degree  per  mole,  (3.7) 

about  the  right  value.  But  this  would  correspond  to  an  equal  mixture 
of  atoms  and  holes,  a  substance  with  a  density  only  half  that  of  the  solid, 
which  is  clearly  impossible.  It  is  unlikely  that  the  crudity  of  our  calcula- 
tion could  make  a  very  large  difference  in  the  result,  so  that  we  may 
conclude  that  the  effect  of  randomness  on  the  entropy  of  fusion  is  impor- 


264  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XVI 

tant,  but  not  the  only  significant  effect.  In  this  connection  it  is  interest- 
ing to  note  that  in  one  or  two  cases  supercooled  liquids  have  been  carried 
down  practically  to  the  absolute  zero  and  their  specific  heats  measured,  so 
that  the  entropy  could  be  determined  at  the  absolute,  zero.  To  within 
an  error  of  a  fow  tenths  of  a  unit,  tho  entropy  was  found  to  be  zero.  This 
would  not  exclude  an  entropy  at  tho  absolute  zero  of  tho  order  of  magni- 
tude of  Kq.  (3.5),  which  seems  possible,  but  it  definitely  would  show  that 
tho  out  ropy  difference  between  solid  and  liquid  comes  mostly  at  tempera- 
tures above  the  absolute1  zero. 

It  appears  from  the  previous  paragraph  that  the  larger  part  of  tho 
entropy  of  fusion  must  arise  because  the  liquid  has  a  larger  specific  heat 
than  tho  solid  in  the  hypothetical  state  of  supercooling,  so  that  its  entropy 
difference1  between  absolute  zero  and  the  melting  point  is  greater  than  for 
tho  solid.  An  indication  as  to  why  this  should  bo  tnio  is  soon  in  the 
preceding  paragraphs  of  the  present  section,  whore  we  have  discussed 
tho  change  of  entropy  with  volume.  This  is  represented  graphically  in 
Fig.  XIII-3,  whore  tho  entropy  is  shown  as  a  function  of  temperature,  for 
several  volumes.  At  larger  volumes  (as  V  =  VQ  in  Fig.  XIII-3),  the 
natural  frequencies  of  molecular  vibration  are  lower,  so  that  tho  specific 
heat  rises  to  its  classical  value  at  lower  temperatures,  and  the  specific, 
heat,  and  consequently  the  entropy,  are  greater  than  at  the  smaller 
volume  (V  —  0.7Fo  in  the  figure,  for  instance).  In  tho  particular  case 
shown  in  tho  figure,  the  entropy  difference  between  tho  two  volumes 
shown,  which  differ  by  30  per  cent,  amounts  to  about  throe  entropy  units 
at  temperatures  above  300°  abs.  Something  of  the  same  effect,  though  on 
a  smaller  scale,  would  be  expected  in  comparing  solids  and  liquids,  as  wo 
have  mentioned  at  the  beginning  of  this  section.  The  liquid  is  n  more 
open  structure,  having  therefore  lower  frequencies  of  molecular  vibration 
and  a  more  nearly  classical  specific  heat  at  low  temperatures.  Thus  its 
entropy  difference  between  the  absolute  zero  zmd  the  melting  point,  if 
the  liquid  really  could  bo  carried  to  absolute  zero,  would  bo  greater  than 
for  the  solid.  By  itself,  however,  us  we  ran  judge  from  Table  XVI-2,  it 
seems  unlikely  that  this  effect  would  be  large1  enough.  If  30  per  cent 
difference  in  volume  amounts  to  throe  entropy  units,  wo  should  need 
something  like  15  per  cent  difference  in  volume  to  account  for  the  approxi- 
mately 1.5  entropy  units  needed,  when  we  take  account  of  possible 
entropy  of  the  liquid  coming  from  randomness.  And  this  is  more  than 
the  actual  difference  in  volume,  in  most  cases.  Nevertheless,  there  is  an 
additional  feature  of  difference  between  the  liquid  and  solid  that  might 
lead  to  still  higher  specific  heat  and  entropy  for  the  liquid.  In  Fig.  XVI-1 
we  have  seen  the  type  of  potential  to  be  expected  for  an  appreciable 
number  of  atoms,  those  that  are  capable  of  shifting  to  a  near-by  position 
of  equilibrium  with  small  expenditure  of  energy.  This  is  so  far  from  the 


SBC.  4]  THE  LIQriD  STATE  AND  FUSION  265 

potential  of  a  linear  restoring  force  that  our  whole  discussion  of  specific 
heats,  which  rests  on  simple  harmonic  motion,  does  not  apply  to  it.  As  a 
matter  of  fact,  tho  energy  levels  in  a  potential  of  the  type  shown  in  Fig. 
XVI-1  lie  closer  together  than  we  should  suppose  from  our  study  of  linear 
restoring  forces.  But  in  general,  the  closer  together  the  energy  levels  of 
any  problem  are,  the  lower  the  temperature  at  which  its  specific  heat 
becomes  approximately  classical.  This  reason  to  expect  a  high  specific 
heat  for  a  supercooled  liquid,  in  addition  to  those  already  discussed,  is 
probably  enough  to  account  for  the  entropy  difference  between  the  liquid 
and  tho  solid.  It  is  very  hard  to  get  a  satisfactory  way  of  calculating  tho 
magnitude  of  this  entropy  difference,  however,  and  we  must  remain  con- 
tent with  a  qualitative  explanation  of  the  values  of  Table  XVI-1.  One 
thing  is  clear  from  our  discussion:  it  will  hardly  bo  possible  to  understand 
fusion  without  studying  the  liquid  state  as  woll  as  tho  solid  state  from 
the  standpoint  of  the  quantum  theory,  and  this  is  a  field  that  has  hardly 
been  explored  at  all.  No  treatment  based  puroly  on  classical  theory  can 
be  expected  to  be  very  good. 

4.  Statistical  Mechanics  and  Melting. — Objection  might  bo  made  to 
our  argument  of  tho  preceding  section,  in  which  we  considered  a  hypo- 
thetical supercooled  state  of  tho  liquid  down  to  tho  absolute  zero,  on  tho 
ground  lhat  that  state  is  not  one  of  thermal  equilibrium  and  that  wo  can- 
not proporly  consider  it  by  itself  at  all.  A  correct  statistical  treatment 
should  yield  the  equilibrium  state  at  any  temperature;  that  is,  below 
the  melting  point  it  should  give  the  solid,  above  the  molting  point  tho 
liquid,  with  a  discontinuous  change  in  properties  at  that  point.  Wo 
shall  now  show  by  a  simple  example  that  the  statistical  treatment  really 
will  give  such  a  discontinuous  change,  but  that  nevertheless  our  method 
of  treatment  was  entirely  justified.  We  shall  calculate  the  partition 
function,  and  from  it  the  free  energy,  of  a  simple  model  of  solid  and  liquid, 
and  shall  show  that  the  free  energy  as  a  function  of  temperature  is  a 
function  with  practically  a  discontinuous  slope  at  a  given  temperature,  the 
molting  point,  below  which  one  phase,  the  solid,  is  stable,  and  abovo 
which  another,  the  liquid,  is  tho  stable  phase. 

To  describe  our  model,  we  shall  give  its  energy  levels,  so  that  we  can 
calculate  the  partition  function  directly.  The  simplest  model  that  shows 
the  properties  we  wish  is  the  following.  We  assume  a  single  level,  at 
energy  NE8,  whore  N  is  the  number  of  atoms,  corresponding  to  tho  solid. 
At  a  higher  energy,  NEi,  we  assume  a  multiple  level  corresponding  to  tho 
liquid.  The  energy  is  higher  on  account  for  example  of  the  greater 
interatomic  distance  in  liquids.  The  multiplicity  of  the  level  arises,  for 
example,  on  account  of  the  randomness  of  molecular  arrangement.  Wo 
assume  that  the  multiple  energy  level  at  NEi  really  consists  of  WN  coin- 
cident levels,  where  w  is  a  constant.  Then  the  partition  function  is 


266 


INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XVI 


NE. 
~  kT  , 


NEi 

~  kT' 


(4.1) 


and  the  Helmholtz  free  energy,  which  is  equal  to  the  Gibbs  free  energy 
in  this  case  where  there  is  no  pressure,  is 


/       NE,  NEj\ 

G  =  -kTln  (e    kT+w"e~'  kT). 


(4.2) 


If  we  had  only  the  solid,  or  only  the  single  level  at  NEa,  the  partition 
function  would  have  contained  only  the  first  term  in  Eq.  (4.2),  and  the 
free  energy  would  have  been 

G,  =  NE..  (4.3) 

If  we  had  only  the  liquid,  or  the  multiple  level  at  NEi,  we  should  have  had 
only  the  second  term  in  Eq.  (4.2),  and  the  free  energy  would  have  been 


-  (Nklnw)T, 


(4.4) 


in  which  the  first  term,  NEi,  is  the  internal  energy,  and  Nk  In  w  is  the 
entropy,  exactly  analogous  to  Kq.  (1.3),  Chap.  Ill  (the  number  of  states 


NE, 


NE, 


Tm  T 

FIG.  XVI-2. — Gibbs  free  energy  as  function  of  temperature,  for  simplified  model  of  solid 
and  liquid,  illustrating  change  of  phase  on  melting. 

is  here  WN,  so  that  the  entropy  should  be  k  In  (WN)  =  Nk  In  w).  For  free 
energy  as  a  function  of  temperature  we  should  then  have  the  two  straight 
lines  of  Fig.  XVI-2,  the  horizontal  one  representing  the  solid,  the  sloping 
one  the  liquid.  The  slope  of  the  curve  measures  the  negative  of  the 
entropy,  as  we  see  at  once  from  Eqs.  (4.3)  and  (4.4),  where  the  solid  has 
zero  entropy,  the  liquid  the  positive  entropy  Nk  In  w.  This  accords  at 
once  with  the  thermodynamic  relation  S  =  —(dA/dT)v  =  —  (dG/dT)P. 
From  Fig.  XVI-2,  we  see  that  the  solid  has  the  lower  free  energy  at  tem- 
peratures below  the  intersection,  on  account  of  its  lower  internal  energy, 
while  the  liquid  has  lower  free  energy  above  the  intersection,  its  greater 
entropy  resulting  in  a  downward  slope  which  counteracts  its  greater 
internal  energy.  The  melting  point  comes  at  the  intersection,  at  the 


SBC.  4]  THE  LIQUID  STATE  AND  FUSION  267 

temperature  where  G6  =  Gi,  or  at 


m       ~T-\  -  —   A  0   >  'W 

kin  w         ASm  x      ' 

where  the  latent  heat  of  melting,  Lm,  equals  N(Ei  —  Et)  and  the  entropy 
of  melting,  A/SW,  equals  JV/fc  In  w. 

The  calculation  we  have  just  made,  considering  the  solid  and  liquid 
separately,  drawing  a  free  energy  curve  for  oach,  for  all  temperatures, 
whether  they  are  stable  or  not,  and  finding  which  free  energy  curve  is 
lower  at  any  given  temperature,  is  analogous  to  the  method  used  in 
this  chapter  to  discuss  fusion  and  also  to  the  method  used  in  Chap.  XII, 
for  example  in  Fig.  XII-4,  in  discussing  vaporization  by  Van  der  Waals' 
equation.  Properly,  however,  we  should  have  used  directly  the  single 
free  energy  formula  (4.2),  and  plotted  it  as  a  function  of  temperature. 
This  almost  precisely  equals  the  function  G9  when  T  <  Tm,  and  GI  when 
T  >  Tm.  For  if  T  «  Tm,  the  first  term  in  the  bracket  of  Eq.  (4.2) 
is  much  larger  than  the  second,  and  Eq.  (4.3)  is  a  good  approximation  to 
Eq.  (4.2),  while  if  T  >  >  Tm  the  second  term  is  much  larger  than  the 
first,  and  Eq.  (4.4)  is  the  correct  approximation.  The  formula  (4.2), 
however,  represents  a  curve  which  joins  these  two  straight  lines  con- 
tinuously, bending  sharply  but  not  discontinuously  through  a  small  range 
of  temperatures,  in  which  the  two  terms  of  Eq.  (4.2)  are  of  the  same 
order  of  magnitude.  For  practical  purposes,  this  rango  of  temperatures  is 
so  small  that  it  can  be  neglected.  Let  us  compute  it,  by  finding  the 
temperature  T  at  which  the  second  term  in  the  bracket  of  Kq.  (4.2)  has 
a  certain  ratio,  say  c,  to  the  first  term.  That  is,  we  have 

-JVEi 

kT 


c  kT 

,            .,  .             N(Ei  -  E.) 
In  c  =  N  In  w pr 

=  JV  hi  w(  } ^J 

=  N  In  wf — T^"/'  (4'6) 

using  Eq.  (4.5).     Thus  we  have 

T  -  Tm         In  c 


Nlnw 


Here  In  u;  is  of  the  order  of  magnitude  of  unity.     If  we  ask  for  the  tem- 
perature when  the  second  term  of  Eq.  (4.2)  is,  say,  ten  times  the  first,  or 


268  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XVI 

one  tenth  the  first,  we  have  c  equal  respectively  to  10  or  -fa,  so  that  In  c 
is  of  the  order  of  magnitude  of  unity,  and  positive  or  negative  as  the  case 
may  be.  Thus  (T  -  Tm)/T  is  of  the  order  of  I/TV,  showing  that  the 
range  of  temperature  in  which  the  correct  free  energy  departs  from  the 
lower  of  the  two  straight  lines  in  Fig.  XVI-2  is  of  the  order  of  1/N  of 
the  melting  temperature,  a  wholly  negligible  range  if  N  is  of  the  order  of 
the  number  of  molecules  in  an  ordinary  sample.  Thus,  our  procedure 
of  finding  the  intersection  of  the  two  curves  is  entirely  justified. 

The  real  situation,  of  course,  is  much  more  complicated  than  this. 
There  are  many  stationary  states  for  the  solid,  corresponding  to  different 
amounts  of  excitation  of  vibrational  energy;  when  we  compute  the  parti- 
tion function  considering  all  these  states,  and  from  it  the  free  energy,  we 
get  a  curve  like  those  of  Fig.  XIII-6,  curving  down  as  the  temperature 
increases  to  indicate  an  increasing  entropy  and  specific  heat.  Similarly 
the  liquid  will  have  not  merely  one  multiple  level  but  a  distribution  of 
stationary  states.  There  is  even  good  reason,  on  the  quantum  theory,  to 
doubt  that  the  lowest  level  of  the  liquid  will  be  multiple  at  all;  it  is  much 
more  likely  that  it  will  be  spread  out  into  a  group  of  closely  spaced,  but 
not  exactly  coincident,  levels,  so  that  the  entropy  will  really  be  zero  at 
the  absolute  zero,  but-  will  rapidly  increase  to  the  value  characteristic 
of  the  random  arrangement  as  the  temperature  rises  above  absolute  zero. 
However  this  may  be,  there  will  surely  be  a  great  many  more  levels  for 
the  liquid  than  for  the  solid,  in  a  given  range  of  energies,  and  the  liquid 
levels  will  lie  at  definitely  higher  energies  than  those  of  the  solid.  This  is 
all  we  need  to  have  a  partition  function,  like  Eq.  (4.1),  consisting  of  two 
definite  parts :  a  set  of  low-lying  levels,  which  arc  alone  of  importance  at 
low  temperatures,  and  a  very  much  larger  set  of  higher  levels,  which  are 
negligible  at  low  temperatures  on  account  of  the  small  exponential  Boltz- 
mann  factor  but  which  become  much  more  important  than  the  others  at 
high  temperature,  on  account  of  their  great  number.  In  turn  this  will 
lead  to  a  free  energy  curve  of  two  separate  segments,  joined  almost  with 
discontinuous  slope  at  the  melting  point. 

It  is  not  impossible,  as  a  matter  of  fact,  to  imagine  that  the  two 
groups  of  levels,  those  for  the  solid  and  for  the  liquid,  should  merge 
partly  continuously  into  each  other.  An  intermediate  state  between  solid 
and  liquid  would  be  a  liquid  with  a  great  many  extremely  small  particles 
of  crystal  in  it,  or  a  solid  with  many  amorphous  flaws  in  it  that  simulated 
the  properties  of  the  liquid.  Such  states  are  dynamically  possible  and 
would  give  a  continuous  series  of  energy  levels  between  solid  and  liquid. 
If  there  were  enough  of  these,  they  could  affect  our  conclusions,  in  the 
direction  of  rounding  off  the  curve  more  than  we  should  otherwise  expect, 
so  that  the  melting  would  not  be  perfectly  sharp.  We  can  estimate 
very  crudely  the  temperature  range  in  which  such  a  hypothetical  gradual 


SEC.  4]  THE  LIQUID  STATE  AND  FUSION  269 

change  might  take  place,  from  our  formula  (4.7).  Suppose  that  instead 
of  considering  the  melting  of  a  large  crystal,  we  consider  an  extremely 
small  crystal  containing  only  a  few  hundred  atoms.  Then,  by  Eq.  (4.7), 
the  temperature  range  in  which  the  gradual  change  was  taking  place 
might  be  of  the  order  of  a  fraction  of  a  per  cent  of  the  melting  temperature, 
or  a  degree  or  so.  Crystals  of  this  size,  in  other  words,  would  not  have  a 
perfectly  sharp  melting  point,  and  if  the  material  breaks  up  into  as  fine- 
grained a  structure  as  this  around  the  melting  point,  even  a  large  crystal 
might  melt  smoothly  instead  of  discontinuously.  The  fact,  however,  that 
observed  melting  points  of  pure  materials  are  as  sharp  as  they  are,  shows 
that  this  effect  cannot  be  very  important  in  a  large  way.  In  any  case,  it 
cannot  affect  the  fundamental  validity  of  the  sort  of  calculation  which 
we  have  made,  finding  the  melting  point  by  intersection  of  free  energy 
curves  for  the  two  phases;  for  mathematically  it  simply  amounts  to  an 
extremely  small  rounding  off  of  the  intersection.  We  shall  come  back  to 
such  questions  later,  in  Chap.  XVIII,  where  we  shall  show  that  in  certain 
cases  there  can  be  a  large  rounding  off  of  such  intersections,  with  corre- 
sponding continuous  change  in  entropy.  This  is  not  a  situation  to  be 
expected  to  any  extent,  however,  in  the  simple  problem  of  melting. 


CHAPTER  XVII 
PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS 

In  the  preceding  chapter  wo  have  been  considering  the  equilibrium 
of  two  phases  of  the  same  substance.  Some  of  the  most  important  cases 
of  equilibrium  come,  however,  in  binary  systems,  systems  of  two  com- 
ponents, and  we  shall  take  thorn  up  in  this  chapter.  We  can  best  under- 
stand what  is  meant  by  this  by  some  examples.  The  two  components 
mean  simply  two  substances,  which  may  bo  atomic  or  molecular  and 
which  may  mix  with  each  other.  For  instance,  they  may  be  substances 
like  sugar  and  water,  one  of  which  is  soluble  in  the  other.  Then  the  study 
of  phase  equilibrium  becomes  the  study  of  solubility,  the  limits  of  solu- 
bility, the  effect  of  the  solute  on  the  vapor  pressure,  boiling  point,  molting 
point,  etc.,  of  the  solvent.  Or  the  components  may  be  metals,  like  copper 
and  zinc,  for  instance.  Then  we  meet  the  study  of  alloys  and  the  whole 
field  of  metallurgy.  Of  course,  in  metallurgy  one  often  has  to  deal  with 
alloys  with  more  than  two  components — ternary  alloys,  for  instance,  with 
three  components — but  they  are  considerably  more  complicated,  and  we 
shall  not  deal  with  them. 

Binary  systems  can  ordinarily  exist  in  a  number  of  phases.  For 
instance,  the  sugar-water  system  can  exist  in  the  vapor  phase  (practically 
pure  water  vapor),  the  liquid  phase  (the  solution),  and  two  solid  phases 
(pure  solid  sugar  and  ice).  The  copper-zinc  system  (the  alloys  that  form 
brasses  of  various  compositions),  can  exist  in  vapor,  liquid,  and  five  or 
more  solid  phases,  each  of  which  can  exist  over  a  range  of  compositions. 
Our  problem  will  be  to  investigate  the  equilibrium  between  these  phases. 
We  notice  in  the  first  place  that  we  now  have  three  independent  variables 
instead  of  the  two,  which  we  have  ordinarily  had  before.  In  addition  to 
pressure  and  temperature,  we  have  a  third  variable  measuring  the  com- 
position. We  ordinarily  take  this  to  be  the  relative  concentration  of  ono 
or  the  other  of  the  components,  N\/(N\  +  AT2)  or  N2/(Ni  +  N2),  as 
employed  in  Chap.  VIII,  Sec.  2;  since  these  two  quantities  add  to  unity, 
only  one  of  them  is  independent.  Then  we  can  express  any  thermo- 
dynamic  function,  as  in  particular  the  Gibbs  free  energy,  as  a  function 
of  the  three  independent  variables  pressure,  temperature,  and  composi- 
tion. We  shall  now  ask,  for  any  values  of  pressure,  temperature,  and 
composition,  which  phase  is  the  stable  one;  that  is,  which  phase  has  the 
smallest  Gibbs  free  energy.  In  some  cases  we  shall  find  that  a  single 

270 


SEC.  1]  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  271 

phase  is  stable,  while  in  other  cases  a  mixture  of  two  phases  is  more  stable 
than  any  single  phase.  Most  phases  are  stable  over  only  a  limited  range 
of  compositions,  as  well  as  of  pressure  and  temperature.  For  instance,  in 
the  sugar  and  water  solution,  the  liquid  is  stable  at  a  given  temperature, 
only  up  to  a  certain  maximum  concentration  of  sugar.  Above  this 
concentration,  the  stable  form  is  a  mixture  of  saturated  solution  and 
solid  sugar.  The  solid  phases  in  this  case,  solid  sugar  and  solid  ice,  are 
stable  only  for  quite  definite  compositions;  for  any  other  composition, 
that  is  for  any  mixture;  of  sugar  and  water,  the  stable  form  of  the  solid  is  a 
mixture  of  sugar  and  ice.  On  the1  other  hand,  we  have  stated  that  the 
solid  phases  of  brass  are  stable  ovor  a  considerable  range  of  compositions, 
though  for  intermediate  compositions  mixtures  of  two  solid  phases  are 
stable.  In  studying  these  subjects,  the  first  thing  is  to  get  a  qualitative 
idea  of  the  various  sorts  of  phases  that  exist,  and  we  proceed  to  that  in 
the  following  section. 

1.  Types  of  Phases  in  Binary  Systems. — A  two-component  system, 
like  a  system  with  a  single  component,  can  exist  in  solid,  liquid,  and 
gaseous  phases.  The  gas  phase,  of  course,  is  perfectly  simple:  it  is  simply 
a  mixture  of  the  gas  phases  of  the  two  components.  Our  treatment  of 
chemical  equilibrium  in  gases,  in  Chap.  X,  includes  this  as  a  special 
case.  Any  two  gases  can  mix  in  any  proportions  in  a  stable  way,  so  long 
as  they  cannot  react  chemically,  and  we  shall  assume  only  the  simple  case 
where  the  two  components  do  not  react  in  the  gaseous  phase;. 

The  liquid  phase  of  a  two-component  system  is  an  ordinary  solution. 
The  familiar  solutions,  like  that  of  sugar  in  water,  exist  only  when  a 
relatively  small  amount  of  one  of  the  components,  called  the  solute  (sugar 
in  this  case)  is  mixed  with  a  relatively  large  amount  of  the  other,  the* 
solvent.  But  this  is  the  case  mostly  with  components  of  very  different 
physical  properties,  like  sugar  and  water.  Two  similar  components  often 
form  a  liquid  phase  stable  over  large  ranges  of  composition,  or  even  for  all 
compositions.  Thus  water  and  ethyl  alcohol  will  form  a  solution  in  any 
proportion,  from  pure  water  to  pure  alcohol.  And  at  suitable  tempera- 
tures, almost  any  two  metals  will  form  a  liquid  mixture  stable  at  any 
composition .  A  liquid  solution  is  similar  in  physical  properties  to  any 
other  liquid.  We  have;  seen  that  an  atom  or  molecule  in  an  ordinary 
liquid  is  surrounded  by  neighboring  atoms  or  molecules  much  as  it  would 
be  in  a  solid,  but  the  ordered  arrangement  does  not  extend  beyond  nearest 
neighbors.  When  we  have  a  mixture  of  components,  it  is  obvious  that 
each  atom  or  molecule  will  be  surrounded  by  some  others  of  the  same; 
component  but  some  of  the  other  component.  If  an  atom  or  molecule  of 
one  sort  attracts  an  unlike  neighbor  about  as  well  as  a  like  neighbor,  and 
if  atoms  or  molecules  of  both  kinds  fit  together  well,  the  solution  may  well 
have  an  energy  as  low  as  the  liquids  of  the  individual  components.  In 


272  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XVII 


addition,  A  golntjon.  like  p,  mixture  of  gases  t  has  ail  entropy  of  mixing,  so 
that  the  entropy  of  the  solution  will  be  greater  than  that  of  the  com- 
ponents. In  such  a  case,  then,  the  Gibbs  free  energy  of  the  solution  will 
he  lower  than  for  the  pure  liquids  and  it  will  be  the  stable  phase.  On 
1  he  other  hap^r  if  thn  fttoms  or  molecules  of  the  two  sorts  fit  together  badlv 
or  do  not  attmr.fr  ^q-eh  nf.hor,  the  rmerfiy  of  the  solution  may  well  be  greater 
1  1)  an  that  of  the  j>ure  liquids,  enough  greater  to  make  the  free  energy 
greater  in  spite  of  the  entropy  of  mixing,  so  that  the  stable  situation  will 
be  a  mixture  of  the  pure  liquids,  or  a  liquid  and  solid,  depending  on  the 
temperature.  Thus  oil  and  water  do  not  dissolve  in  each  other  to  any 
extent.  Their  molecules  are  of  very  different  sorts,  and  the  energy  is 
lower  if  water  molecules  are  all  close  together  and  if  the  oil  molecules 
are  congregated  together  in  another  region.  That  is,  the  stable  situation 
will  be  a  mixture  of  the  two  phases,  forming  separate  drops  of  oil  and 
water,  or  an  emulsion  or  suspension.  A  very  little  oil  will  presumably  be 
really  dissolved  in  the  water  and  a  very  little  water  in  the  oil,  but  the  drops 
will  be  almost  pure. 

As  we  have  just  seen,  the  condition  for  the  existence  of  a  liquid  phase 
stable  for  a  wide  range  of  concentrations  (that  is,  for  a  large  solubility 
of  one  substance  in  another),  is  that  the  forces  acting  between  atoms  or 
molecules  of  one  component  and  those  of  the  other  should  be  of  the  same 
order  of  magnitude  as  the  forces  between  pairs  of  atoms  or  molecules  of 
either  component,  so  that  the  internal  energy  of  the  solution  will  be  at 
least  as  low  as  that  of  the  mixture,  and  the  entropy  of  mixing  will  make 
the  free  energy  lower  for  the  solution.  Let  us  consider  a  few  specific 
cases  of  high  solubility.  In  the  first  place,  we  are  all  familiar  with  the 
large1  solubility  of  many  ionic  salts  in  water.  The  crystals  break  up  into 
ions  in  solution,  and  these  ions,  being  charged,  orient  the  electrical 
dipoles  of  the  water  around  them,  a  positive  ion  pulling  the  negatively 
charged  oxygen  end  of  the  water  molecule  toward  it,  a  negative  ion  pulling 
the  positively  charged  hydrogen.  This  leads  to  a  large  electrical  attrac- 
tion, and  a  low  energy  and  free  energy.  The  resulting  free  energy  will  be 
lower  than  for  the  mixture  of  the  solid  salt  and  water,  unless  the  salt  is 
very  strongly  bound.  Water  is  not  the  only  solvent  that  can  form  ionic 
solutions  in  this  way.  Liquid  ammonia,  for  instance,  has  a  large  dipole 
moment  and  a  good  many  of  the  same  properties,  and  the  alcohols,  also 
with  considerable  dipole  moments,  are  fairly  good  solvents  for  some  ionic 
crystals. 

Different  dipole  liquids,  similarly,  attract  each  others'  molecules  by 
suitable  orientation  of  the  dipoles  and  form  stable  solutions.  We  have 
already  mentioned  the  case  of  alcohol  and  water.  In  ammonia  and 
water,  the  interaction  between  neighboring  ammonia  and  water  molecules 
is  so  strong  that  they  form  the  ammonium  complex,  leading  to  NH4OH 


SEC.  1]  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  273 

if  the  composition  is  correct,  and  a  solution  of  NH4OH  in  either  water  or 
ammonia  if  there  is  excess  of  either  component.  The  substance  NH^)!! 
is  generally  considered  a  chemical  compound;  but  it  is  probably  more 
correct  simply  to  recognize  that  a  neighboring  water  and  ammonia  mole- 
cule will  organize  themselves,  whatever  may  be  the  composition  of  the 
solution,  in  such  a  way  that  one  of  the  hydrogens  from  the  water  and 
the  three  hydrogens  from  the  ammonia  form  a  fairly  regular  tetrahedral 
arrangement  about  the  nitrogen,  as  in  the  ammonium  ion.  There  are  no 
properties  of  the  ammonia- water  system  which  behave  strikingly  differ- 
ently at  the  composition  NH<OH  from  what  they  do  at  neighboring 
concentrations. 

Solutions  of  organifi  substances  are  almost  invariably  made  in  organi( 
solvents^  simply  because  here  again  the  attractive  forces  between  twi 
different  types  of  molecule  are  likely  to  be  large  if  the  molecules  an 
similar.  Different  hydrocarbons,  for  instance,  mix  in  all  proportions,  ai- 
one  is  familiar  with  in  the  mixtures  forming  kerosene,  gasoline,  etc.  The 
forces  between  an  organic  solvent  and  its  solute,  as  between  the  molecule? 
of  an  organic  liquid,  are  largely  Van  der  Waals  forces,  though  in  some 
cases,  as  the  alcohols,  there  are  dipole  forces  as  well. 

In  the  metals,  the  same  type  of  interatomic,  force1  acts  bet  ween  atom,- 
of  different  metals  that  acts  between  atoms  of  a  single  element.  We  have 
already  stated  that  for  this  reason  liquid  solutions  of  many  metals  with 
each  other  exist  in  wide  ranges  of  composition.  There,  are  many  other 
cases  in  which  two  substances  ordinarily  solid  at  room  temperature  are 
soluble  in  each  other  when  liquefied.  Thus,  a  great  variety  of  molten 
ionic  crystals  are  soluble  in  each  other.  And  among  the  silicates  and 
other  substances  held  by  valence  bonds,  the  liquid  phase  permits  a  wide 
range  of  compositions.  This  is  familiar  from  the  glasses,  which  can  have 
a  continuous  variability  of  composition  and  which  can  then  supercool  to 
essentially  solid  form,  still  with  quite  arbitrary  compositions,  and  yet 
perfectly  homogeneous  structure. 

Solid  phases  of  binary  systems,  like  the  liquid  phases,  are  very  com- 
monly of  variable  composition.  Here,  as  with  the  liquid,  the  stable  range 
of  composition  is  larger,  the  more  similar  the  two  components  are.  This 
of  course  is  quite  contrary  to  the  chemists'  notion  of  definite  chemical 
composition,  definite  structural  formulas,  etc.,  but  those  notions  are 
really  of  extremely  limited  application.  It  happens  that  the  solid  phases 
in  the  system  water — ionic  compound  are  often  of  rather  definite  com- 
position, and  it  is  largely  from  this  rather  special  case  that  the  idea  of 
definite  compositions  in  solids  has  become  so  firmly  rooted.  In  such  a 
system,  there  are  normally  two  solid  phases:  ice  and  the  crystalline  ionic 
compound.  Ice  can  take  up  practically  none  of  any  ionic  compound,  so 
that  it  has  practically  no  range  of  compositions.  And  many  ionic  crystals 


274  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 

take  up  practically  no  water  in  their  crystalline  form.  But  there  are 
many  ionic  crystals  which  are  said  to  have  water  of  crystallization. 
Water  molecules  form  definitely  a  part  of  the  structure.  And  in  some  of 
these  the  proportion  of  water  is  not  definitely  fixed,  so  that  they  form 
mixed  phases  of  variable  composition. 

Water  and  ionic  compounds  are  very  different  types  of  substances,  and 
it  is  not  unnatural  that  they  do  not  form  solids  of  variable  composition. 
The  reason  why  water  solutions  of  ionic  substances  exist  is  that  the  water 
molecules  can  rotate  so  as  to  be  attracted  to  the  ions;  this  is  not  allowed 
in  the  solid,  where  the  ice  structure  demands  a  fairly  definite  orientation 
of  the  molecule.  But  as  soon  as  we  think  about  solid  phases  of  a  mixture 
of  similar  components,  we  find  that  almost  all  the  solid  phases  exist  over 
quite  a  range.  Such  phases  are  often  called  by  the  chemists  solid  solu- 
tions, to  distinguish  them  from  chemical  compounds.  This  distinction 
is  valid  if  we  mean  by  a  chemical  compound  a  phase  which  really  exists 
at  only  a  quite  definite  composition.  But  the  chemists,  and  particularly 
the  metallurgists,  are  not  always  careful  about  making  this  distinction;  for 
this  reason  the  notation  is  misleading,  and  we  shall  not  often  use  it. 

Solid  phases  of  variable  composition  exist  in  the  same  cases  that  we 
have  already  discussed  in  connection  with  liquids.  Thus  mixtures  of 
ionic  substances  can  often  form  crystals  with  a  range  of  composition. 
The  conditions  under  which  this  range  of  composition  is  large  are  what  we 
should  naturally  suppose:  the  ions  of  the  two  components  should  be  of 
about  the  same  size  and  valence,  and  the  two  components  should  be 
capable  of  existing  in  the  same  crystal  structure.  We  shall  meet  many 
examples  of  such  solids  of  variable  composition  later,  when  we  come  to 
study  different  types  of  materials.  The  best-explored  range  of  solid 
compounds  of  variable  composition  comes  in  metallurgy.  Here  an  atom 
can  replace  another  of  the  same  size  quite  freely  but  not  another  of  rather 
different  size.  Thus  the  copper  and  nickel  atoms  have  about  the  same 
size;  they  form  a  phase  stable  in  all  proportions.  On  the  other  hand, 
calcium  and  magnesium  have  atoms  of  quite  different  sizes,  normally 
existing  in  different  crystal  structures,  and  they  cannot  bo  expected  to 
substitute  for  each  other  in  a  lattice.  They  form,  as  a  matter  of  fact,  as 
close  an  approach  to  phases  of  definite  chemical  composition  as  wre  find 
among  the  metals.  They  form  three  solid  phases:  pure  magnesium,  pure 
calcium,  and  a  compound  CasMg4,  and  no  one  of  these  is  soluble  to  any 
extent  in  any  of  the  others;  that  is,  each  exists  with  almost  precisely  fixed 
composition.  Most  pairs  of  elements  are  intermediate  between  these. 
They  form  several  phases,  each  stable  for  a  certain  range  of  compositions, 
and  often  each  will  be  centered  definitely  enough  about  some  simple 
chemical  composition  so  that  it  has  been  customary  to  consider  them  as 
being  chemical  compounds,  though  this  is  not  really  justified  except  in 


SEC.  2]  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  275 

such  a  definite  case  as  CaaMg4.  Each  of  the  phases  in  general  has  a 
different  crystal  structure.  Of  course,  the  crystal  cannot  be  perfect,  for 
ordinarily  it  contains  atoms  of  the  two  components  arranged  at  random 
positions  on  the  lattice.  It  is  the  lattice  that  determines  the  phase,  not 
the  positions  of  the  metallic  atoms  in  it.  But  if  the  two  types  of  atom  in 
the  lattice  are  very  similar,  they  will  not  distort  it  much,  so  that  it  will 
be  practically  perfect.  For  compositions  intermediate  between  those  in 
which  one  of  the  phases  can  exist,  the  stable  situation  will  be  a  mixture  of 
the  two  phases.  This  occurs,  in  the  solid,  ordinarily  as  a  mixture  of  tiny 
crystals  of  the  two  phases,  commonly  of  microscopic  size,  with  arbitrary 
arrangements  and  sizes.  It  is  obvious  that  the  properties  of  such  a 
mixture  will  depend  a  great  deal  on  the  size  and  orientation  of  the  crystal 
grains;  these  are  things  not  considered  in  the  thermodynamical  theory  at 
all. 

2.  Energy  and  Entropy  of  Phases  of  Variable  Composition. — The 
remarks  we  have  just  made  about  mixtiires  of  crystalline  phases  raise  the 
question,  what  is  a  single  phase  anyway?  We  have  not  so  far  answered 
this  question,  preferring  to  wait  until  we  had  some  examples  to  consider. 
A  single  phase  is  a  mixture  that  is  homogeneous  right  down  to  atomic 
dimensions.  If  it  has  an  arbitrary  composition,  it  is  obvious  that  really 
on  the  atomic  scale  it  cannot  be  homogeneous,  but  if  it  is  a  single  phase 
we  assume  that  there  is  no  tendency  for  the  two  types  of  atom  to  segregate 
into  different  patches,  even  patches  of  only  a  few  atoms  across.  On  the 
other  hand,  a  mixture  of  phases  is  supposed  to  be  one  in  which  the  two 
types  of  atoms  segregate  into  patches  of  microscopic  or  larger  size.  These 
two  types  of  substance  have  quite  different  thermodynamic  behavior, 
both  as  to  internal  energy  and  as  to  entropy.  We  shall  consider  this 
distinction,  particularly  for  a  metallic  solid,  but  in  a  way  which  applies 
equally  well  to  a  liquid  or  other  type  of  solid. 

Suppose  our  substance  is  made  of  constituents  a  and  b.  Let  the 
relative  concentration  of  a  be  ca  =  £;  of  6,  c&  =  1  —  x.  Assume  the 
atoms  arc  arranged  on  a  lattice  in  such  a  way  that  each  atom  has  s  neigh- 
bors  (s  =  8  for  the  body-centered  cubic  structure,  12  for  face-centered 
cubic  and  hexagonal  close  packed,  etc.).  In  a  real  solid  solution,  or 
homogeneous  phase,  there  will  be  a  chance  x  of  finding  an  atom  a  at  any 
lattice  point,  a  chance  1  —  .r  of  finding  an  atom  6.  We  assume  a  really 
random  arrangement,  so  that  the  chance  of  finding  an  atom  at  a  given 
lattice  point  is  independent  of  what  happens  to  be  at  the  neighboring 
points.  This  assumption  will  be  examined  more  closely  in  the  next  chap- 
ter, where  we  take  up  questions  of  order  and  disorder  in  lattices.  Then 
out  of  the  5  neighbors  of  any  atom,  on  the  average  sx  will  be  of  type  a, 
s(l  —  x)  of  type  b.  If  we  consider  all  the  pairs  of  neighbors  in  the  crystal, 
we  shall  have 


276  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVIT 

Nsx*      .       ,  L 

-  ^  -  pairs  of  type  na 

£t 

Ns(\  -  .r)2       .        ,  4         M 
--  £  pairs  of  type  60 

Nsx(l  —  a:)  pairs  of  type  ab.  (2.1) 

Hero  N  is  the  total  number  of  atoms  of  both  sorts.  The  factors  £  in 
Eq.  (2.1)  arise4  because  each  pair  must  be  counted  only  once,  not  twice 
as  we  should  if  \\e  said  that  the  number  of  pairs  of  type  aa  equaled  the 
number  of  atoms  of  type  a  (Nx)  times  the  number  of  neighbors  of  type  a 
which  each  one  had  (s.r).  Now  we  make  a  simple  assumption  regarding 
the  energy  of  the  crystal  at  the  absolute  zero.  We  assume  that  the  total 
energy  can  be  written  as  a  sum  of  terms,  one  for  each  pair  of  nearest  neigh- 
bors. We  shall  be  interested  in  this  energy  only  at  the  normal  distance 
of  separation  of  atoms.  At  this  distance,  we  shall  assume  that  the  energy 
of  a  pair  aa  is  Eaa,  of  a  pair  bb  is  7?w>,  and  of  a  pair  ab,  Eab.  All  these 
quantities  will  be  negative,  if  we  assume  the  zero  state  of  energy  is  the 
state  of  infinite  separation  of  the  atoms  (the  most  convenient  assumption 
for  the  present  purpose)  and  if  all  pairs  of  atoms  attract  each  other. 
Then  the  total  energy  of  the  crystal,  at  the  absolute  zero,  will  be 

~E(Ul  +  N*(}~  ^Ebh  +  Nsx(l  -  x)Eab 

+  2r(l  -  x)(Eab  -  ^-f^)}     (2.2) 

According  to  our  assumptions,  the  energy  of  a  crystal  wholly  of  a  is 
(Ns/2)Eaa,  and  wholly  of  6  is  (Ns/2)Ehb.  Thus  the  first  two  terms  on  the 
right  side  of  Eq.  (2.2)  give  the  sum  of  the  energies  of  a  fraction  x  of 
the  substance  «,  and  a  fraction  (1  —  x)  of  6.  These  two  would  give  the 
whole  energy  in  case  wre  simply  had  a  mixture  of  crystals  of  a  and  b.  But 
the  third  term,  involving  x(l  —  x),  is  an  additional  term  arising  from  the 
mutual  interactions  of  the  two  types  of  atoms.  The  function  x(}  —  x)  is 
always  positive  for  values  of  x  between  0  and  1,  being  a  parabolic  function 
with  maximum  of  £  when  x  =  ?,  and  going  to  zero  at  x  =  0  or  1 .  Thus 
this  last  term  has  a  sign  which  is  the  same  as  that  of  Eab  —  (Eatt  +  Ebb)/2. 
If  Eab  is  more  positive  than  the  mean  of  Eaa  and  EM  (that  is,  if  atoms  a 
and  b  attract  each  other  less  strongly  than  they  attract  atoms  of  their  own 
kind),  then  the  term  is  positive.  In  this  case,  the  solution  will  have 
higher  energy  than  the  mixture  of  crystals,  and  if  the  entropy  term  in 
the  free  energy  does  not  interfere,  the  mixture  of  crystals  will  be  the  more 
stable.  On  the  other  hand,  if  Eab  is  more  negative  than  the  mean  of  Eau 
and  EM,  so  that  atoms  of  the  opposite  sort  attract  more  strongly  than 
either  one  attracts  its  own  kind,  the  term  will  be  negative  and  the  solution 


SBC.  2]  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  277 

will  have  the  lower  energy.  In  order  to  get  the  actual  internal  energy  at 
any  temperature,  of  course  we  must  add  a  specific  heat  term.  We  shall 
adopt  the  crude  hypothesis  that  the  specific  heat  is  independent  of 
composition.  This  will  be  approximately  true  with  systems  made  of  two 
similar  components.  Then  in  general  we  should  have 

rr       Ns\    T1      .    ,.         >.  „      io/i         \i  v         H<in  +  ^MM 
"2"         a  +  ^    ~~  ^       +      ^    "~  x\Aab  ~~          2        ) 

f*T 

+        CpdT.     (2.3) 
Jo 

Next,  lot  us  consider  tho  entropy  of  the  homogeneous  phase  and 
compare  it  with  the  entropy  of  a  mixture  of  two  pure  components.  Tho 
entropy  of  tho  puro  oompononts  will  bo  just  the  part  determined  from  the 


specific  heat,  or   I     -ni  dT.     But  in  tho  solution  there  will  bo  an  additional 
Jo    J 

term,  the  entropy  of  mixing.  This,  as  a,  matter  of  fact,  is  just  the  same* 
term  found  for  gases  in  Kq.  (2.12),  Chap.  VIII:  it  is 

Art  -  -Nk[x  In  x  +  (1  -  x)  hi  (1  -  jr)], 

in  the  notation  of  the  present  chapter.  We  can,  however,  justify  it 
directly  without  appealing  to  the  theory  of  gases,  which  certainly  cannot 
be  expected  to  apply  directly  to  tho  present  case.  Wo  use  a  method  like 
that  used  in  deriving  Eq.  (3.5),  Chap.  XVI.  We  have  a  lattice  with 
N  points,  accomodating  NX  atoms  of  one  sort,  N(l  —  x)  of  another  sort. 
There  are  then 

AM 
W  =  ______  -    '  (9  4 

(Nx)\[N(l  -  ( 


ways  of  arranging  the  atoms  on  the  lattice  points,  almost  all  of  whirl) 
have  approximately  the  same  energy.  Using  Stirling's  formula,  this 
becomes 

w  - 


By  Boltzmann's  relation  S  =  k  In  W,  this  gives  for  the  entropy  of  mixing 
Atf  =  -Nk[x  In  x  +  (1  -  x)  In  (1  -  x)].  (2.6) 

The  other  thermodynamic  functions  are  also  easily  found.  If  we  confine 
ourselves  to  low  pressures,  of  the  order  of  atmospheric  pressure,  we  can 
neglect  the  term  P  V  in  the  Gibbs  free  energy  of  liquid  or  solid.  Then  the 
Helmholtz  and  Gibbs  free  energies  are  approximately  equal  and  are  given 
by 


278 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 


(a)  Eab  -  (Kaa  + 


>  0; 


Ox, 


0.5 
(q)Eqb-E*°*Ebb>0 


x,  I 


05 


1   x 


A  =  (?  =  U  +  NkT[x  In  x  +  (1  -  x)  In  (1  -  x)]  -  T  \    ^  dT,     (2.7) 

Jo    i 

where  U  is  given  in  Eq.  (2.3). 

We  have  now  found  approximate  values  for  the  thermodynamic 
functions  of  our  homogeneous  phase.  The  entropy  is  greater  than  that 
of  the  mixture  by  the  term  (2.6),  and  the  internal  energy  is  either 
greater  or  smaller,  as  we  have  seen.  To  illustrate  our  results,  we  give 
in  Fig.  XVII-1  a  sketch  of  G  as  a  function  of  a*,  for  three  cases: 

(b)  Eab  -  (Eaa  +  E^/2  =  0;  (c)  Eab  - 
(Eon  +  Ebb)/2  <  0.  We  observe  that  in 
each  case  the  free  energy  begins  to 
decrease  sharply  as  x  increases  from 
zero  or  decreases  from  unity.  This  is 
on  account  of  the  logarithmic  function 
in  the  entropy  of  mixing  (2.6).  But  in 
case  (a),  whero  the  atoms  prefer  to 
segregate  rather  than  forming  a  homo- 
geneous phase,  the  free  energy  then  rises 
for  intermediate  concentrations,  while  in 
case  (6),  where  tho  atoms  are  indifferent 
to  their  neighbors,  or  in  (c)  where  they 
definitely  prefer  unlike*  neighbors,  the 
free  energy  falls  for  intermediate  con- 
centrations, moro  strongly  in  case  (c). 
In  each  case  we  have  drawn  a  dotted  line 
connecting  the  points  x  =  0  and  x  =  1. 
This  represents  the  free  energy  of  tho 
mixture  of  crystals  of  tho  pure  compo- 
nents. We  see  that  in  cases  (6)  and  (c) 
the  solution  always  has  a  lower  free 
energy  than  the  mixture  of  crystals,  but 
<  0  in  (a)  there  is  a  range  where  it  does  not. 

FIG.  xvn-i.-GibbB  free  energy  of  However,  we  shall  see  in  the  next  section 

a     binary     system,     as     function     of    that  \VC  must  look  a  little  more  Carefully 
concentration.  into  the  situation  before  bc}ng  surc  what 

forms   the   stable   state   of   the   system. 

3.  The  Condition  for  Equilibrium  between  Phases. — Suppose  we 
have  two  homogeneous  phases,  one  with  composition  XL  and  free  energy 
Gij  the  other  with  composition  z2  and  free  energy  Gr2.  By  mixing  these 
two  phases  in  suitable  proportions,  the  resulting  mixture  can  have  a 
composition  anywhere  between  x\  and  rr2.  And  the  free  energy,  being 
simply  the  suitably  weighted  sum  of  the  free  energies  of  the  two  phases,  is 
given  on  a  G  vs.  x  plot  simply  as  a  straight  line  joining  the  points  Gi,  x\ 


(b)Eab- 


•  =  0 


SKC.  31  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  279 

and  (72,  #2-  That  is,  for  an  intermediate  composition  corresponding  to  a 
mixture,  the  free  energy  has  a  proportional  intermediate  value  between 
the  free  energies  of  the  two  phases  being  mixed.  We  saw  a  special  case 
of  this  in  Fig.  XVII-1,  where  the  dotted  line  represents  the  free  energy 
of  a  mixture  of  the  two  phases  with  x  =  0,  x  =  i  respectively. 

Now  suppose  we  take  the  curve  of  Fig.  XVII-1  (a)  and  ask  whether 
by  mixing  two  phases  represented  by  different  points  on  this  curve,  we 
can  perhaps  get  a  lower  Gibbs  free  energy  than  for  the  homogeneous 
phase.  It  is  obvious  that  we  can  and  that  the  lowest  possible  lino 
connecting  two  points  on  the  curve  is  the  mutual  tangent  to  the  two 
minima  of  the  curve,  shown  by  (?i(?2  in  Fig.  XVII-1  (a).  The  point  G\ 
represents  the  free  energy  of  a  homogeneous  phase  of  composition  Xi  and 
the  point  (?2  of  composition  #2.  Between  these  compositions,  a  mixture 
of  thffije  f,wo  homogeneous  phases  represented  by  the  dotte^  line  will 
have  lower  Gibbs  free  energy  than  the  homogeneous  phase,  and  will 
represent  the  stable  situation.  For  x  loss  than  x\,  or  greater  than  xz,  the 
straight  line  is  meaningless;  for  it  would  represent  a  mixture  with  more 
than  100  per  cent  of  one  phase,  less  than  zero  of  the  other.  Thus  for  x 
less  than  Xi,  or  greater  than  xz,  the  homogeneous  phase  is  the  stable 
one.  In  other  words,  we  have  a  case  of  a  system  that  has  only  two 
restricted  ranges  of  composition  in  which  a  single  phase  is  stable,  while 
between  these  ranges  we  can  only  have  a  mixture  of  phases. 

We  can  now  apply  the  conditions  just  illustrated  to  some  actual 
examples.  Ordinarily  we  have  two  separate  curves  of  G  against  xt  to 
represent  the  two  phases;  Fig.  XVII-1  (a)  was  a  special  case  in  that  a 
single  curve  had  two  minima.  And_in  a  region  where  a  common  tangent 
to  the  curve  lies  lower  than  either  curve  between  the  points  of  tangcncy, 
the  mixture  of  the  two  phases  represented  by  the  points  of  tanpencv  will 
be  the  stable  phase.  First  we  consider  the  equilibrium  between  liquid 
and  solid,  in  a  case  where  the  components  are  soluble  in  all  proportions, 
both  in  solid  and  liquid  phases.  For  instance,  we  can  take  the  case  of 
melting  of  a  copper-nickel  alloy.  To  fix  our  ideas,  let  copper  be  consti- 
tuent a,  nickel  constituent  6,  so  that  x  —  0  corresponds  to  pure  nickel, 
.r  =  1  to  pure  copper.  We  shall  assume  that  in  both  liquid  arid  solid  the 
free  energy  has  the  form  of  Fig.  XVII-1  (6),  in  which  the  bond  between  a 
copper  and  a  nickel  atom  is  just  the  mean  of  that  between  two  coppers 
and  two  nickels.  This  represents  properly  the  case  with  two  such  similar 
atoms.  Such  a  curve  departs  from  a  straight  line  only  by  the  entropy  of 
mixing,  which  is  definitely  known,  the  same  in  liquid  and  solid.  Thus  it 
is  determined  if  we  know  the  free  energy  of  liquid  and  solid  copper  and 
nickel,  as  functions  of  temperature.  From  our  earlier  work  we  know  how 
to  determine  these,  both  experimentally  and  theoretically.  In  particular, 
we  know  that  the  free  energy  for  liquid  nickel  is  above  that  for  solid 


280 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 


nickel  at  temperatures  below  the  melting  point,  1702°  abs.  but  below  at 
iemperatures  above  the  melting  point,  and  that  similar  relations  hold  for 
copper  with  melting  point  1356°  abs.  Wo  further  know  that  the  rate  of 
change  of  the  difference  between  the  free  energy  of  liquid  and  solid  nickel, 
with  temperature,  is  the  difference  of  thoir  entropies,  or  T  times  the  latent 
heat  of  fusion.  Thus  we  have  enough  information  to  draw  at  least 
approximately  a  set  of  curves  like  those  shown  in  Fig.  XVII-2.  Hero 


Ni  Cu     Ni  Cu     Ni  x,    x2      Cu 

T,  <  1,356°  Abs          T2  =  1,356°  Abs  1,356<T3  <  1,702° 


Ni  Cu        Ni  Cu 

T4  =  1,702°  Abs  T5  =>  1,702°  Abs 

KUJ.   XVII-2.-    (ribbs  free  energy  for  solid  and  liquid  as  function  of  concent  ration,  for 
different  temperatures,  Ni-Cu  system. 

\ve  give G for  UK;  solid,  G1  for  tho  liquid  phase,  as  functions  of  composition, 
for  five  temperatures:  Ti  below  1356°,  7\  at  1356°,  773  between  1356  and 
1702°,  T±  at  1702°,  and  7T5  above  1702°.  Below  1356°,  the  free  energy 
for  the  solid  is  everywhere  below  that  for  the  liquid,  so  that  the  former 
is  the  stable  phase  at  any  composition.  At  1356°,  the  liquid  curve 
touches  the  solid  one,  at  100  per  cent  copper,  and  above  this  temperature 
the  curves  cross,  the  solid  curve  lying  below  in  systems  rich  in  nickel, 
the  liquid  curve  below  in  systems  rich  in  copper.  In  this  case,  T*  in  the 
figure,  we  can  draw  a  common  tangent  to  the  curves,  from  G\  to  (?2  at 


SEC.  3] 


PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS 


281 


1700 


concentrations  x\  and  x%.  In  this  range,  then,  for  concentrations  of 
copper  below  x\,  a  solid  solution  is  stable;  above  x$,  a  liquid  solution  is 
stable;  while  between  x\  and  x2  there  is  a  mixture  of  solid  of  composition 
Xi  and  liquid  of  composition  .r2.  These  two  phases,  in  other  words,  are 
in  equilibrium  with  each  other  in  any  proportions.  At  1702°,  the  range 
in  which  tho  liquid  is  stable  has  extended  to  the  whole  range  of  com- 
positions, the  curve  of  G  for  the  liquid  lying  lower  for  all  higher 
temperatures. 

The  stability  of  the  phases  can  be  shown  in  a  diagram  like  Fig.  X  VII-3, 
railed  a  phase  diagram.  In  this,  temperature  is  plotted  as  ordinate, 
composition  as  abscissa,  and  lines  separate  the  regions  in  which  various 
phases  are  stable.  Thus,  at  high  temperature,  the  liquid  is  stable  for  all 
compositions.  Running  from  1702°  to  1356°  are  iwo  curves,  one  called 
the  liquidus  (the  upper  one)  and 
the  other  called  the  solidus.  For 
any  T-x  point  lying  between  the 
liquidus  and  solidus,  the  stable 
state  is  a  mixture  of  liquid  and  solid. 
Moreover,  we  can  read  off  from  the 
diagram  the  compositions  of  the 
liquid  and  solid  in  equilibrium  at 
any  temperature.  The  horizontal 
line  drawn  in  Fig.  XVII-3,  at  tem- 
perature Z'8  (see  Fig.  XVII-2), 
cuts  the  solidus  at  composition  x± 
and  the  liquidus  at  :r2,  agreeing  with 
the  compositions  for  7T3  in  Fig.  XVII-2.  Then  .ri  represents  the 
composition  of  the  solid,  o*2  of  the  liquid,  in  equilibrium  with  eaeh  other  at 
this  temperature.  Finally,  below  the  solidus  the  stable  phase  is  always 
the  solid. 

From  the  phase  diagram  we  can  draw  information  not  only  about 
equilibrium  but  about  the  process  of  solidification  or  melting.  Suppose 
we  have  a  melt  of  composition  or2,  at  a  temperature  above  the  liquidus,  and 
suppose  we  gradually  cool  the  material.  The  composition  of  course, 
will  not  change  until  we  reach  the  liquidus,  and  solid  begins  to  freeze 
out.  But  now  the  solid  in  equilibrium  with  liquid  of  composition  .r2  has 
the  composition  Xi,  much  richer  in  nickel  than  the  liquid.  This  will  be 
frozen  out,  and  as  a  result  the  remaining  liquid  will  bo  deprived  of  nickel 
and  will  become  richer  in  copper.  Its  concentration  will  then  lie  farther 
to  the  right  in  the  diagram,  so  that  it  will  intersect  tho  liquidus  at  a  lower 
temperature.  As  the  temperature  is  decreased,  then,  some  of  this  liquid, 
perhaps  of  composition  z2,  will  have  solid  of  composition  x(  freeze  from 
it,  further  enriching  the  liquid  in  copper.  This  process  continues,  more 


Fl°* 


Ni 

XVII-'J  —  Phiiso    diagram 
system. 


for 


282  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 

and  more  liquid  freezing  out  until  the  temperature  reaches  1356°,  when 
the  last  portion  of  the  liquid  will  freeze  out  as  pure  copper.  There  are 
two  interesting  results  of  this  process.  In  the  first  place,  the  freezing 
point  is  not  definite;  material  freezes  out  through  a  range  of  temperatures, 
all  the  way  from  the  temperature  corresponding  to  the  point  xz  on  the 
graph  to  the  freezing  point  of  pure  copper.  In  the  second  place,  the 
material  which  has  frozen  out  is  by  no  moans  homogeneous.  Ordinarily 
it  will  freeze  out  in  tiny  crystal  grains.  And  we  shall  observe  that  the 
first  grains  freezing  out  are  rich  in  nickel,  while  successive  grains  are 
more  and  more  rich  in  copper,  unt  il  the  last  material  frozen  is  pure  copper. 
The  over-all  composition  of  the  solid  finally  left  is  of  course  the  same 
as  that  of  the  original  liquid,  but  it  is  not  of  homogeneous  composition 
and  henco  is  not  a  stable  material.  We  can  see  this  from  Fig.  XVII-2, 
where  the  curves  of  G  vs.  composition  for  the  solid  are  convex  downward. 
With  such  a  curve,  the  G  at  a  definite  composition  is  necessarily  lower 
than  the  average  G's  of  two  compositions,  one  richer  and  the  other  poorer 
in  copper,  which  would  contain  the  same  total  amount  of  each  element. 
By  an  extension  of  this  argument,  the  inhomogeneous  material  freezing 
out  of  the  melt  must  have  a  higher  free  energy  than  homogeneous  material 
of  the  same  net  composition,  and  on  account  of  its  thermodynamic  insta- 
bility it  will  gradually  change  over  to  the  homogeneous  form.  This 
change  can  be  greatly  accelerated  by  raising  the  temperature,  since  as 
mentioned  in  Sec.  1,  Chap.  XVI,  the  rate  of  such  a  process,  involving  the 
changing  place  of  atoms,  depends  on  a  factor  exp  (  —  e/kT),  increasing 
rapidly  with  temperature.  Accordingly,  material  of  this  kind  is  often 
annealed,  held  at  a  temperature  slightly  below  the  melting  point  for  a 
considerable  time,  to  allow  thermodynamic  equilibrium  to  take  place 
and  at  the  same  time  to  allow  mechanical  strains  to  be  removed. 

The  reverse  process  of  fusion  can  be  discussed  much  as  we  have  con- 
sidered solidification.  Of  course,  if  we  take  the  solid  just  as  it  has  solidi- 
fied, without  annealing,  there  will  be  crystal  grains  in  it  of  many  different 
compositions,  which  will  melt  at  different  temperatures,  the  liquids 
mixing.  But  if  we  start  with  the  equilibrium  solid,  of  a  definite  composi- 
tion, it  will  begin  to  melt  at  a  definite  temperature.  The  liquid  melting 
out  will  have  a  higher  concentration  of  copper  than  the  solid,  however, 
leaving  a  nickel-rich  material  of  higher  melting  point.  The  last  solid  to 
melt  will  be  rich  in  nickel,  of  such  a  composition  as  to  be  in  equilibrium 
with  the  liquid.  It  is  interesting  to  notice  that  the  process  of  melting 
which  we  have  just  described  is  not  the  exact  reverse  for  solidification. 
This  is  natural  when  we  recall  that  the  solid  produced  directly  in  solidi- 
fication is  not  in  thermodynamic  equilibrium,  so  that  when  the  proc- 
ess is  carried  on  with  ordinary,  finite  velocity  it  is  not  a  reversible 
process. 


SEC.  4] 


PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS 


283 


4.  Phase  Equilibrium  between  Mutually  Insoluble  Solids. — In  the 
preceding  section  we  have  considered  phase  equilibrium  between  solid  and 
liquid,  in  the  case  where  the  components  were  soluble  in  each  other  in  any 
proportions,  in  both  liquid  and  solid.  Now  we  shall  consider  the  case 
where  there  is  practically  no  solubility  of  one  solid  in  the  other;  that  is, 
there  are  two  solid  phases,  each  one  stable  in  only  a  very  narrow  range  of 


FIG.  XVII-4. — Gibbs  free  energy  for  solids  and  liquid  a&  function  of  concentration,  for 
different  temperatures,  in  a  system  with  almost  mutually  insoluble  solids. 

concentration  about  x  =  0  or  x  =  1.  The  free  energy  for  the  solid  will 
have  much  the  form  given  in  Fig.  XVII-1  (a).  But  we  shall  assume  that 
the  minima  of  the  curve  are  extremely  sharp  and  we  shall  not  assume  that 
both  minima  belong  to  the  same  curve.  In  a  case  like  this,  it  is  most 
likely  that  the  pure  phase  of  one  component  will  have  different  crystal 
structure  and  other  properties  from  the  pure  phase  of  the  other,  and  there 
will  be  no  sort  of  continuity  between  the  phases,  as  in  Fig.  XVII-1  (a). 
For  the  liquid  we  shall  again  assume  the  form  of  Fig.  XVII-1  (b).  Then 
we  give  in  Fig.  XVII-4  a  series  of  curves  for  G  against  x  at  increasing 
temperatures,  and  in  Fig.  XVII-5  the  corresponding  phase  diagram.  The 


284 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 


method  of  construction  will  be  plain  by  comparison  with  the  methods 
used  in  Figs.  XVII-2  and  XVII-3.  At  low  temperatures  like  7\,  there  is 
a  very  small  range  of  compositions  from  0  to  Xi,  in  which  phase  a  is  stable, 
and  another  small  range,  from  .r2  to  1,  in  which  phase  ft  is  stable.  Here 
we  have  given  the  name  a  to  the  phase  composed  of  pure  a  with  a  little 
b  dissolved  in  it,  and  ft  to  pure  b  with  a  little  a  dissolved  in  it.  If  the 
substanc.es  a  and  6  are  really  mutually  insoluble  in  the  solid,  the  two 
curves  representing  G  vs.  x  for  solids  a  and  ft  will  have  infinitely  sharp 
minima  in  Fig.  XVII-4,  rising  to  groat  heights  for  x  infinitesimally  greater 
than  zero,  or  infinitesimally  less  than  1.  For  the  whole  range  of  composi- 
tions between  x\  and  Xz,  at  these  low  temperatures,  the  stable  form  will 

be  a  mechanical  mixture  of  crystals  of  a 
and  0. 

At  a  somewhat  higher  temperature, 
between  rl\  arid  7T2,  the  G  curve  for  the 
liquid  will  fall  low  enough  to  be  tangent 
to  the  straight  line  representing  the  mix- 
ture of  Xi  and  a>2.  This  is  for  the  com- 
position denoted  by  x^^  in  Fig.  XVII-5 
and  will  be  discussed  later.  At  higher 
temperatures,  as  T2,  there  is  a  range  of 
compositions  from  0*2  to  x3,  in  which  the 
liquid  is  the  stable  phase,  while  for  com- 
a  *  b  x  positions  from  :TI  to  xz  the  stable  form  is 

Fi<}    xvn-5.    Ph!seICoquihbrium  a  mixture  of  liquid  and  phase  a,  and  from 
diagram  for  a  system  with  almost  #3  to  #4  it  is  a  mixture  of  liquid  and  phase 

mutually  insoluble  solids,  as  given  in    „         A       ,,        ,  .  •          ,       rn       .1 

Fig.  XVII-4.  P*     As  the  temperature  rises  to  7  3,  the 

melting  point  of  pure  material  b,   the 

phase  ft  disappears,  and  at  7T&,  the  melting  point  of  pure  a,  the  phase 
a  disappears,  leaving  only  the  liquid  as  the  stable  phase  above  this 
temperature. 

The  process  of  freezing  is  similar  to  the  previous  case  of  Fig.  XVII-3. 
Suppose  the  liquid  has  a  composition  between  x  =  0  and  #eutoetic.  Then 
as  it  is  cooled,  it  will  follow  along  a  line  like  the  dotted  line  in  Fig.  XVII-5, 
which  intersects  the  line  marked  #2  in  the  figure  at  temperature  T3.  At 
this  temperature  it  will  begin  to  freeze,  but  the  material  freezing  out  will 
be  phase  a  with  the  composition  x\  appropriate  to  that  temperature,  very 
rich  in  component  a.  The  liquid  becomes  enriched  in  6,  so  that  it  has  a 
lower  melting  point,  and  we  may  say  that  the  point  representing  the 
concentration  and  temperature  of  the  liquid  on  Fig.  XVII-5  follows  down 
along  the  curve  x*.  When  the  composition  reaches  the  eutectic  composi- 
tion and  the  temperature  is  still  further  reduced,  a  liquid  phase  is  no 
longer  possible,  and  the  remaining  liquid  freezes  at  a  definite  temperature 


SEC.  4]  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  285 

and  composition.  It  is  to  bo  noticed,  however,  that  the  resulting  solid  is 
still  a  mixture  of  phases  a  and  ft.  With  the  usual  methods  of  freezing,  the 
two  phases  freeze  out  as  alternating  layers  of  platelike  crystals.  Such  a 
solid  is  called  a  eutectic  and  is  of  importance  in  metallurgy.  It  is  inter- 
esting to  observe  that  if  the  composition  of  the  original  liquid  is  just  the 
eutectic  composition,  it  will  all  freeze  at  a  single  temperature,  which 
will  be  the  lowest  possible  freezing  point  for  any  mixture  of  a  and  6.  If 
the  original  composition  is  between  xeutoctit.  and  x  =  1,  the  situation  will 
be  similar  to  what  we  have  described,  only  now  the  point  representing 
the  liquid  will  move  down  curve  .Ta  to  tho  eutectic  composition,  and  the 
solid  freezing  out  will  be  phase  ft,  and  tho  liquid  will  become  enriched  in 
component  a  until  it  reaches  the  eulectic  composition,  when  it,  will  all 
freeze  as  the  cutcctie  mixture  of  a  and  ft.  The  temperature  where  this 
freezing  of  the  eutectir  occurs,  we;  notice,  represents  a  triple  point :  phases 
a,  ft,  and  the  liquid  are  all  stable  at  this  temperature  in  any  proportions, 
corresponding  to  the  fact  that  a  single  tangent  can  be  drawn  in  the  (f-x 
diagram  to  the  curves  representing  all  three  phases.  For  every  pressure, 
there  is  a  temperature  at  which  there  is  such  a  triple  point,  in  contrast 
to  the  situation  with  a  one-component  system,  where  triple  points  exist 
only  for  certain  definite  combinations  of  pressure  and  temperature.  The 
difference  arises  because  there  are  more  independent  variables,  the 
composition  as  well  as  pressure  and  temperature. 

Familiar  examples  of  the  situation  wo  have  just  described  are  found  in 
the  solubility  of  substances  in  water  and  other  solvents.  Thus  in  Fig. 
XVII-6  we  give  the  phase  diagram  for  the  familiar  system  NaCl-water. 
This  diagram  is  not  carried  to  a  very  high  concentration  of  salt,  for  then 
the  curve  corresponding  to  x3  would  rise  to  such  high  temperatures  that 
we  should  be  involved  with  the  vaporization  of  the  water,  which  we  have 
not  wished  to  discuss.  In  this  system,  as  we  have  already  mentioned,  the 
solid  phases  are  practically  completely  insoluble  in  each  other,  the  phase 
corresponding  to  a  being  pure  ice,  ft  being  pure  solid  NaCl,  combined  with 
the  water  of  crystallization  at  low  temperature.  Thus  the  curves  Xi  and 
24  of  Fig.  XVII-5  do  not  appear  in  Fig.  XVII-6  at  all,  coinciding  prac- 
tically with  the  lines  x  =  0  and  x  =  1.  We  can  now  find  a  number  of 
interesting  interpretations  of  Fig.  XVII-6.  In  the  first  place,  if  water 
with  a  smaller  percentage  of  salt  than  the  eutectic  mixture  is  cooled 
down,  the  freezing  point  will  be  below  0°C.,  the  freezing  point  of  pure 
water,  showing  that  the  dissolved  material  has  lowered  the  freezing  point. 
We  shall  calculate  the  amount  of  this  lowering  in  the  next  section.  At 
these  compositions,  the  solid  freezing  out  is  pure  ice.  This  is  familiar 
from  the  fact  that  sea  water,  which  has  less  salt  than  the  eutectic  mixture, 
forms  ice  of  pure  water  without  salt.  On  the  other  hand,  if  the  liquid 
has  a  larger  percentage  of  salt  than  the  eutectic  mixture,  the  solid  freezing 


286 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 


out  as  the  temperature  is  towered  is  pure  salt.  Under  these  circum- 
stances we  should  ordinarily  describe  the  situation  differently.  We 
should  say  that  as  the  temperature  was  lowered,  the  solubility  of  the  salt 
in  water  decreased  enough  so  that  salt  precipitated  out  from  solution.  In 
other  words,  the  curve  separating  the  liquid  region  in  Fig.  XVII-6  from 
the  region  where  liquid  and  NaCl  are  in  equilibrium  may  be  interpreted 
as  the  curve  giving  tho  percentage  of  salt  in  a  saturated  solution  or  the 
solubility  as  a  function  of  temperature.  The  rise  to  the  right  shows  that 
the  solubility  increases  rapidly  with  increasing  temperature. 


100 
90 
80 
70 
60 

</)   50 

$   40 

g*  30 

^    20 

10 

0 

-10 
-20 


Liquid 


Ice  +  Liquid" 


NaCl 


NaCl- 
2H20 
-+Liq. 


5  10          15          20         25         30 

Grams  NaCl  per  100  Grams  Solution 

FIG.  XVII-6. ---Equilibrium  between  NaCl  Jiud  water. 

From  Fig.  XVII-6  wo  can  also  understand  the  behavior  of  freezing 
mixtures  of  ice  and  salt.  Suppose  ice  and  salt,  both  at  approximately 
0°C.,  are  mixed  mechanically  in  approximately  the  right  proportions  to 
give  the  eutcctic  mixture.  We  see  from  Fig.  XVII-6  that  a  solid  of  this 
composition  is  not  in  thennodynamic  equilibrium  at  this  temperature;  the 
stable  phase  is  the  liquid,  which  has  a  lower  free  energy  than  the  mixture 
of  solids.  Thus  the  material  will  spontaneously  liquefy,  the  solid  ice  and 
salt  dissolving  each  other  at  their  surfaces  of  contact  and  forming  brine. 
If  the  process  were  conducted  isothermally,  we  should  end  up  with  a 
liquid.  But  in  the  process  a  good  deal  of  h^at  would  have  to  be  absorbed, 
the  latent  heat  of  fusion  of  the  material.  Actually,  in  using  a  freezing 
mixture,  the  process  is  more  nearly  adiabatic  than  isothermal:  heat  can 
flow  into  the  mixture  from  the  system  which  is  to  be  cooled,  but  that 
system  has  a  small  enough  heat  capacity  so  that  its  temperature  is  rapidly 
reduced  in  the  process.  In  order  to  get  the  necessary  latent  heat,  in  other 
words,  the  freezing  mixture  and  external  system  will  all  cool  down  below 
0°C.,  falling  to  lower  and  lower  temperatures  as  more  and  more  of  the 
freezing  mixture  melts.  The  process  can  continue,  if  the  proportion  of  ice 


SBC.  4] 


PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS 


287 


to  salt  is  just  the  eutectic  proportion,  down  to  the  temperature  — 18°,  the 
lowest  temperature  at  which  the  liquid  can  exist. 

The  most  important  examples  of  the  phase  diagrams  we  have  discussed 
are  found  in  metallurgy.     There,  in  alloys  of  two  metals  with  each  other, 


1,000° 


800°  - 


600°  - 


400° 


Cu  MGJ 

FIQ.  XVII-7. — Phase  equilibrium  diagram  for  the  system  Cu-Mg,  in  which  tho  two  metals 
are  insoluble  in  each  other,   forming  mtermetallie  compounds  of  definite  composition. 


1000   ~ 


400  - 

Cu  Zn 

FIG.  XVII-8. — Phase  equilibrium  diagram  for  the  system  Cu-Zn,  in  which  a  number 
of  phases  of  variable  composition  arc  formed,  mixtures  of  the  phases  being  stable  between 
tho  regions  of  stability  of  the  pure  phases.  The  phase  a  is  face  centered  cubic,  as  Cu  is, 
/3  is  body  centered,  7  is  a  complicated  structure,  €  and  17  are  hexagonal.  The  transition 
between  ft  and  0'  is  an  order-disorder  transition,  (3  being  disordered,  and  /8'  ordered,  as 
discussed  in  the  following  chapter. 

we  generally  find  much  more  complicated  cases  than  those  take  up  so 
far,  but  still  cases  which  can  be  handled  by  the  same  principles.  Thus  in 
Fig.  XVII-7  we  show  the  phase  diagram  for  the  system  Cu-Mg,  two 


288 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 


metals  that  are  almost  entirely  insoluble  in  each  other.  In  this  case 
there  are  four  solid  phases,  each  having  its  own  crystal  structure,  and 
each  stable  in  only  an  extremely  narrow  range,  about  the  compositions 
Cu,  MgCu2,  Mg2Cu,  and  Mg.  The  free  energy  of  each  composition  will 
then  have  an  extremely  sharp  minimum,  so  that  the  construction  neces- 
sary to  derive  the  phase  diagram  will  be  similar  to  Fig.  XVII-4,  but  with 
four  sharp  minima  instead  of  two,  so  that  there  are  three  regions,  rather 
than  one,  in  which  a  mixture  of  two  phases  is  the  stable  solid,  and  three 
eutectics.  For  contrast,  we  give  in  Fig.  XVII-8  the  phase  diagram  for 
the  system  Cu-Zn,  or  brass.  In  this  case  there  are  a  number  of  phases, 
again  each  with  its  own  crystal  structure  but  each  with  a  wide  range  of 
possible  compositions.  The  free  energy  curves  of  the  various  phases  in 
this  case  are  then  not  sharp  like  the  case  of  Cu-Mg  but  have  rather  flat 

minima,  more  as  in  Fig.  XVII-2.  We 
11  shall  not  try  to  follow  the  construction 
of  the  phase  diagram  through  in  detail 
but  shall  merely  state  that  it  can  be 
derived  from  hypothetical  free  energy 
curves  according  to  the  type  of  reason- 
ing already  used  in  this  section  and  the 
preceding  one. 

5.  Lowering  of  Melting  Points  of 
Solutions. — We  have  just  seen  that  the 
lowering  of  the  melting  point  of  a 
solvent  by  action  of  the  solute  can 

0  1  x   easily  be  explained  in   terms   of    the 

FKJ.  xvil-9.™ Gibbs  free  energy  aa  phase  diagram,  and  it  is  an  easy  matter 

function  of  ronoontrntion,  for  lowering  to  find  a  numerical  Value  for  this 
of  freezing  point.  i  .  T  T^-  VTTTT  r»  i_ 

lowering.     In  rig.  XVII-9  we  have  a 

diagram  of  G  against  x,  appropriate  to  this  case.  The  solid  solute  corre- 
sponds to  the  point  G8,  with  rr  =  0,  and  the  liquid  is  given  by  the  curve. 
We  wish  to  find  the  value  of  x  at  which  a  straight  line  through  x  =  0, 
G  =  Ga  is  tangent  to  the  liquid  curve.  To  do  this,  we  must  first  find  the 
equation  of  the  liquid  curve.  We  assume  the  liquid  to  correspond  to  case 
(6)  of  Fig.  XVII-1,  the  internal  energy  being  a  linear  function  of  concen- 
tration. Then,  if  GIO  is  the  free  energy  of  the  liquid  for  x  =  0,  G^  for 
x  =  1,  we  have 

Gi  =  Gh  +  x(Gh  -  Gto)  +  NkT[x  In  x  +  (1  -  x)  In  (1  -  x)],     (5.1) 

where  Gi  is  the  free  energy  for  the  liquid.  The  desired  tangent  is  now 
determined  by  the  condition 


(5.2) 


SEC.  51  PHASE  EQUILIBRIUM  IN  BINARY  SYSTEMS  289 

the  geometrical  condition  that  the  tangent  to  the  curve  GI  at  the  point 
Gi  should  pass  through  the  point  G8  when  x  =  0.  Differentiating  Eq. 
(5.1),  this  gives 

Gi.  +  NkT  In  (1  -  ar)  =  G9.  (5.3) 

Wo  can  now  find  the  difference  (7jfc  —  G8  in  terms  of  the  latent  heat  of 
fusion  of  the  solvent.  From  fundamental  principles  we  have 

(d(Gio  —  Gs)\  ,„        ~,        —  Lm 

-  ST  ---  )P  =  ~(Sl  "~  &)  =  ~~f~'  (5'4) 

where  Si  is  the  entropy  of  the  liquid,  Sa  of  the  solid,  and  Lm  the  latent 
heat  of  fusion.  Computed  just  at  the  melting  point,  the  quantity  in 
Eq.  (5.4)  becomes  —Lm/Tm.  Now  we  shall  not  use  the  result  except  for 
temperatures  very  close  to  the  melting  point,  so  that  we  may  assume  that 
(Gio  —  Gs)  can  be  expanded  as  a  linear  function  of  temperature.  Just 
at  the  melting  point,  by  the  fundamental  principle  of  equilibrium,  it  is 
zero.  Thus  we  have 

Gln  -G8  =  ^(Tm  -  T).  (5.5) 

J-  m 

Inserting  in  Eq.  (5.3),  setting  T  =  Tm  approximately,  and  writing 
Nk  =  /£,  this  gives  us 

_  In  (1  -  x)  =          (Tr*  -  T).  (5.6) 


For  dilute  solutions,  to  which  alone  we  shall  apply  our  results,  x  is  very 
small,  and  we  may  write  In  (1  —  x)  —  —  x.     Then  we  have 

—  (T     _   T\ 

\*  m       •*•  /> 


(Tm  ~  T)  -         x.  (5.7) 

JUm 

Equation  (5.7)  gives  the  lowering  of  the  freezing  point,  Tm  —  T7,  by 
solution  of  another  substance  with  relative  concentration  x.  We  note 
the  important  fact  that  the  result  is  independent  of  the  nature  of  the 
solute:  all  its  properties  have  canceled  out  of  the  final  answer.  Thus  the 
lowering  of  the  freezing  point  can  be  used  as  a  direct  method  of  measuring 
x,  the  relative  number  of  molecules  of  solute  in  solution.  This  is  some- 
times a  very  important  thing  to  know.  Suppose  one  knows  the  mass  of  a 
certain  solute  in  solution  but  does  not  know  its  molecular  weight.  By 
measuring  the  depression  of  the  freezing  point,  using  Eq.  (5.7),  we  can 
find  the  number  of  moles  of  it  in  solution.  By  division,  we  can  find  at 
once  the  mass  per  mole,  or  the  molecular  weight.  This  method  is  of 


290  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XVII 

practical  value  in  finding  the  molecular  weights  of  complicated  substances. 
It  is  also  of  importance  in  cases  where  there  is  association  or  dissociation 
of  a  solute  in  solution.  Some  materials  form  clusters  of  two,  three,  or 
more  molecules  in  solution,  each  cluster  traveling  around  as  a  single 
molecule.  Each  cluster  will  count  as  a  single  molecule  in  the  entropy  of 
mixing,  and  consequently  in  the  depression  of  the  freezing  point.  Thus 
really  there  are  fewer  molecules  than  one  would  suppose  from  the  known 
amount  of  material  in  solution  and  the  usual  molecular  weight,  so  that  tho 
depression  of  the  freezing  point  is  smaller  than  we  should  suppose.  On 
the  contrary,  in  some  cases  substances  have  their  molecules  dissociated  in 
solution.  The  well-known  case  of  this  is  ionic  substances  in  water  solu- 
tion, in  which  the  ions,  rather  than  molecules,  form  the  separate  objects 
in  the  solution.  In  these  cases  there  are  more  particles  in  solution  than 
we  should  suppose,  and  the  freezing  point  is  depressed  by  an  abnormally 
large  amount. 

From  Eq.  (5.7)  we  can  find  at  once  the  amount  of  depression  of  the 
freezing  point  of  different  solvents.  Thus  for  water,  Tm  =  273°  abs., 
Lm  =  80  X  18  cal.  per  mole,  giving  Tn  -  T  =  103°  for  x  =  1.  To  get 
useful  figures,  we  calculate  for  what  the  chemists  denote  as  a  normal 
solution,  containing  1  mole  of  solute  in  1000  gm.  of  water,  or  nnhr  of  a 
mole  of  solute  in  1  mole  of  water.  Thus  in  a  normal  solution  we  expect  a 
lowering  of  the*  freezing  point  of  103  X  .018  =  1.86°C.,  provided  the 
solute  is  neither  associated  nor  dissociated. 


CHAPTER  XVIII 
PHASE  CHANGES  OF  THE  SECOND  ORDER 

In  an  ordinary  change  of  phase,  there  is  a  sharp  transition  tempera- 
ture, for  a  given  pressure,  at  which  the  properties  change  discontinuously 
from  one  phase  to  a  second  one.  In  particular,  there  is  a  discontinuous 
change  of  volume  and  a  discontinuous  change  of  entropy,  resulting  in  a 
latent  heat  and  allowing  the  application  of  Clapeyron's  equation  to  tin* 
transition.  In  recent  years,  a  number  of  cases  have  been  recognized 
in  which  transitions  occur  which  in  most  ways  resemble  real  changes  of 
phase,  but  in  which  the  changes  of  volume  and  entropy,  instead  of  being 
discontinuous,  are  merely  very  rapid.  Volume  and  entropy  change 
greatly  within  a  few  degrees'  temperature  range,  with  the  result  thai-  there 
is  an  abnormally  large  specific;  heat  in  this  neighborhood,  but  no  latent 
heat.  Often  the  specific  heat  rises  to  a  peak,  then  discontinuously  falls 
to  a  smaller  value.  To  distinguish  these  transitions  from  ordinary 
changes  of  phase,  it  has  become  customary  to  denote  ordinary  phase 
changes  as  phase  changes  of  the  first  order,  and  these  sudden  but  not  dis- 
continuous transitions  as  phase  changes  of  the  second  order.  Sometimes 
the  discontinuity  of  the  specific  heat  is  regarded  as  the  distinguishing 
feature  of  a  phase  change  of  the  second  order,  but  we  shall  not  limit  our- 
selves to  cases  having  such  discontinuities. 

There  is  one  well-known  phenomenon  which  might  well  be  considered 
to  be  a  phase  change  of  the  second  order,  though  ordinarily  it  is  not. 
This  is  the  change  from  liquid  to  gas,  at  temperatures  and  pressures  above 
the  critical  point.  In  this  case,  as  the  temperature  is  changed  at  con- 
stant pressure,  we  have  a  very  rapid  change  of  volume  from  a  small 
volume  characteristic  of  a  liquidlike  state  to  the  larger  volume  charac- 
teristic of  a  gaslike  state,  yet  there  is  no  discontinuous  change  as  then* 
is  below  the  critical  point.  And  there  is  a  very  rapid  change  of  entropy, 
from  the  small  value  characteristic  of  the  liquid  to  the  large  value  charac- 
teristic of  the  gas,  as  we  can  see  from  Fig.  XI-6,  resulting  in  a  very 
abnormally  high  value  of  CP  at  temperatures  and  pressures  slightly  above 
the  critical  point.  At  the  critical  point,  where  the  curve  of  S  vs.  T 
becomes  vertical,  so  that  (6S/dT)P  is  infinite,  CP  becomes  infinite.  At 
this  temperature  and  below,  we  cannot  use  the  specific  heat  to  find  the 
change  of  entropy,  but  must  use  a  latent  heat  instead,  representing,  so  to 
speak,  the  finite  integral  under  the  infinitely  high,  but  infinitely  narrow, 
peak  in  the  curve  of  T(dS/dT)P  vs.  T. 

291 


292  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XVIII 

Although  the  liquid-gas  transition  above  the  critical  point,  as  we 
have  seen,  has  the  proper  characteristics  for  a  phase  change  of  the  second 
order,  that  name  is  ordinarily  used  only  for  phaso  changes  in  solids.  Now 
it  seems  hardly  possible  that  there  could  be  a  continuous  transition  from 
ono  solid  phase  to  another  one  with  different  crystal  structure.  There 
have  boon  some  suggestions  that  such  things  are  possible;  that,  for 
instance,  ordinary  equilibrium  lines  in  polymorphic  transitions,  as  shown 
in  Fig.  XI-3,  might  terminate  in  critical  points,  above  which  one  could 
pass  continuously  from  one  phase  to  another.  But  no  such  critical 
points  have  been  found  experimentally,  and  there  is  no  experimental 
indication,  as  from  a  decreasing  discontinuity  in  volume  and  entropy 
between  the  two  phases  as  we  go  to  higher  pressure  and  temperature,  that 
such  critical  points  would  be  reached  if  the  available  ranges  of  pressure 
and  temperature  could  be  increased.  Thus  it  seems  that  our  na'ivc  sup- 
position that  two  different  crystal  structures  are  definitely  different,  and 
that  no  continuous  series  of  states  can  be  imagined  between  thorn,  is 
really  correct,  and  that  phase  changes  of  the  second  ordor  arc  impossible 
between  phases  of  different  structure  and  must  be  looked  for  only  in 
changes  within  a  single  crystal  structure. 

Then*  fire  at  least  three  types  of  change  known  which  do  not  involve 
changes  of  crystal  structure  and  which  show  the  properties  of  phase 
changes  of  the  second  order.  The  best  known  ono  is  the  ferromagnetic 
change,  between  the  magnetized  state,  for  instance  of  iron  or  nickel,  at 
low  temperatures,  and  the  unmagnotizod  state  at  high  temperatures. 
There  is  no  change  of  crystal  structure  associated  with  this  transition,  at. 
least  in  pure  metals,  no  discontinuous  change  of  volume,  and  no  latent 
hoat.  The  magnetization  decreases  gradually  to  zero,  instead  of  changing 
discontinuously,  though  then*  is  a  maximum  temperature,  called  the 
Curio  point,  from  P.  Curio,  who  investigated  it,  at  which  it  drops  rather 
suddenly  to  zero.  And  there  is  no  latent  hoat,  the  entropy  increasing 
rather  rapidly  as  wo  approach  the  Curio  point,  but  nowhere  changing 
discontinuously,  so  that  there  is  an  anomalously  largo  specific  heat.  This 
anomaly  in  the  specific  hoat  is  sometimes  concentrated  in  a  small  enough 
temperature  range  so  that  it  almost  seems  liko  a  latent  hoat  to  crude 
observation;  the  metallurgists,  who  aro  accustomed  to  determining  phaso, 
changes  by  cooling  curves,  which  essentially  measure  discontinuities  or 
rapid  changes  in  entropy,  have  sometimes  classified  these  ferromagnetic 
changes  as  real  phaso  changes.  As  a  matter  of  fact,  mathematical  analy- 
sis shows  that  under  some,  circumstances  in  alloys,  it  is  possible  for  the 
ferromagnetic  change  to  bo  associated  with  a  change  of  crystal  structure 
and  a  phase  change  of  the  first  ordor,  one  phase  being  magnetic  up  to  its 
transition  point,  above  which  a  new  nonferromagnetic  phase  is  stable,  but 
this  is  a  complication  not  found  in  pure  metals.  Though  this  ferro- 


efcsc.  1]  PHASE  CHANGES  OF  THE  SECOND  ORDER  293 

magnetic  change  is  the  most  familiar  example  of  phase  changes  of  the 
second  order,  we  shall  not  discuss  it  hero. 

A  second  type  of  phase  change  of  the  second  order  is  found  with 
certain  crystals  like  NH4C1  containing  ions  (NH4*  in  this  rase)  which 
might  be  supposed  capable  of  rotation  at  high  temperature  but  not  at 
low.  The  ammonium  ion,  being  tot rahcd rally  symmetrical,  is  not  far 
from  spherical,  and  we  can  imagine  it  to  rotate  freely  in  the  crystal  if 
it  is  not  packed  too  tightly.  At  low  temperatures,  however,  it  will  fit 
into  the  lattice  best  in  one  particular  orientation  and  will  tend  merely 
to  oscillate  about  this  orientation.  The  rotating  state,  it  is  found,  has 
the  higher  entropy  and  is  preferred  at  high  temperatures.  The  change 
from  one  state  to  the  other  comes  experimentally  in  a  nither  narrow 
temperature  range,  giving  a  specific  heat  anomaly  but  no  latent  heat,  and 
forming  again  a  phase  change  of  the  second  order.  Unfortunately  the 
theory  is  rather  involved  and  we  shall  not  try  to  give  it  here. 

The  third  type  of  phase  change  of  the  second  order  is  fortunately 
easy  to  treat  theoretically,  at  least  to  an  approximation,  and  it,  is  the  one 
which  will  be  discussed  in  the  present  chapter.  This  is  what  is  known 
as  an  order-disorder  transition  in  an  alloy,  and  can  be  better  understood 
in  terms  of  specific  examples,  which  we  shall  mention  in  the  next  section. 

1.  Order -Disorder  Transitions  in  Alloys. — The  best-known  example 
of  order-disorder  transitions  comes  in  the  0  phase  of  brass,  Cu-Zn,  a 
phase  which  is  stable  at  compositions  in  the  neighborhood  of  50  per  cunt 
of  each  component.  The  crystal  structure  of  this  phase  is  body-centered 
cubic,  an  essential  feature  of  the  situation.  In  this  type  of  lattice,  the 
lattice  points  are  definitely  divided  into  two  groups:  half  the  points  an; 
at  the  corners  of  the  cubes  of  a  simple  cubic  lattice,  the  other  half  at 
the  centers  of  the  cubes.  It  is  to  be  noticed  that,  though  they  are  dis- 
tinct, the  centers  and  corners  of  the  cubes  are  interchangeable.  Now  we 
can  see  the  possibility  of  an  ordered  state  of  Cu-Zn  in  the  neighborhood 
of  50  per  cent  composition:  the  copper  atoms  can  be  at  the  corners  of  the 
cubes,  the  zinc  at  the  centers,  or  vice  versa,  giving  an  ordered  structure  in 
which  each  copper  is  surrounded  by  eight  zincs,  each  zinc  by  eight  cop- 
pers; whereas  in  the  disordered  state  which  we  have  previously  considered, 
each  lattice  point  would  be  equally  likely  to  be  occupied  by  cither  a  copper 
or  a  zinc  atom,  so  that  each  copper  on  the  average  would  be  surrounded  by 
four  coppers  and  four  zincs. 

Just  as  the  body-centered  cubic  structure  can  be  considered  as  made 
of  two  interpenetrating  simple  cubic  lattices,  the  face-centered  cubic 
structure  can  be  made  of  four  simple  cubic  lattices.  There  are  some 
interesting  cases  of  ordered  alloys  with  this  crystal  structure  and  ratios  of 
approximately  one  to  three  of  the  two  components.  An  example  is  found 
in  the  copper-gold  system,  where  such  a  phase  is  found  in  the  neighbor- 


294  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XVIII 

hood  of  the  composition  CuaAu.  Evidently  the  ordered  phase  is  that 
in  which  the  gold  atoms  are  all  on  one  of  the  four  simple  cubic  lattices, 
tho  copper  atoms  occupying  the  other  three. 

We  shall  now  investigate  phase  equilibrium  of  the  Cu-Zn  type,  start- 
ing with  the  simple  case  of  equal  numbers  of  copper  and  zinc  atoms,  later 
taking  the  general  case  of  arbitrary  composition.  We  shall  make  the 
same  assumptions  about  internal  energy  that  we  have  made  in  Sec.  2, 
Chap.  XVII,  so  that  the  problem  in  computing  the  internal  energy  is  to 
find  the  number  of  pairs  of  nearest  neighbors  having  the  type  aa,  a6,  and 
66;  a  and  b  being  the  two  types  of  atoms.  We  assume  that  the  only 
neighbors  of  a  given  atom  to  be  considered  are  the  eight  atoms  at  the 
corners  of  a  cube  surrounding  it,  so  that  all  the  neighbors  of  an  atom 
on  one  of  tho  simple  cubic  lattices  lie  on  the  other  simple  cubic 
lattice. 

We  shall  now  introduce  a  parameter  w,  which  we  shall  call  the  degree 
of  order.  We  shall  define  it  so  that  w  =  1  corresponds  to  having  all  the 
atoms  a  on  one  of  the  simple  cubic  lattices  (which  we  may  call  the  lattice 
a),  all  the  atoms  b  on  the  other  (which  we  call  ft),  w  =  0  will  correspond 
to  having  equal  numbers  of  atoms  a  and  6  on  each  lattice;  w  =  —  1  will 
correspond  to  having  all  the  atoms  6  on  lattice  a,  all  the  atoms  a  on  lattice 
ft.  Thus  10  =  ±1  will  correspond  to  perfect  order,  w  =  0  to  complete 
disorder.  Let  us  now  define  w  more  completely,  in  terms  of  the  number 
of  atoms  a  and  6  on  lattices  a  and  ft.  Let  there  be  N  atoms,  N/2  of  each 
sort,  and  N  lattice  points,  N/2  on  each  of  the  simple  cubic  lattices.  Then 
we  assume  that 

XT  1  e          9  1       J.J.'  (1      + 

Number  ot  as  on  lattice  a  =    - 


Number  of  a's  on  lattice  ft  = 


4 

(1  -  w)N 


4 

Number  of  6's  on  lattice  a  = T- 

4 


Number  of  fr's  on  lattice  ft  =  (1  .  (1.1) 

Clearly  the  assumptions  (1.1)  reduce  to  the  proper  values  in  the  cases 
w  =  ±1,  0,  and  furthermore  they  give  w  as  a  linear  function  of  the 
various  numbers  of  atoms. 

To  find  the  energy,  we  must  find  the  number  of  pairs  of  neighbors  of 
types  aa,  a&,  66.  The  number  of  pairs  of  type  aa  equals  the  number  of 
a's  on  lattice  a,  times  8 /(N/2)  times  the  number  of  a's  on  lattice  ft.  This 
is  on  the  assumption  that  the  distribution  of  atoms  on  lattice  ft  surround- 
ing an  atom  a  on  lattice  a  is  the  same  proportionally  that  it  is  in  the  whole 
lattice  ft,  an  assumption  which  is  not  really  justified  but  which  we  make 


SEC.  1]  PHASE  CHANGES  OF  THE  SECOND  ORDER  295 

for  simplification.     Thus  the  number  of  pairs  aa  is  -  —    ,          times 
4(1  —  w),  or  N(l  —  w>2).     Similarly  we  have 

Number  of  pairs  aa  =  Number  of  pairs  66  =  N(l  —  w2) 
Number  of  pairs  ab  =  N(l  +  w)2  +  N(l  —  w)2 

w2).  (1.2) 


To  find  the  internal  energy  at  the  absolute  zero,  we  now  proceed  as  in 
Sec.  2  of  Chap.  XVII,  multiplying  the  number  of  pairs  aa  by  Ean,  etc. 
Then  we  obtain 

Energy  =  f/0  =  N(l  ~  to2)(Ea(t  +  E»)  +  2N(l  +  w*)E*.     (1.3) 
rfhis  can  be  rewritten  in  the  form 


xEa<l    +    (1     ~    X)Eu   +   2X(1     - 

+  $Nx2w2(Eab  -  ~aa^~\     (1.4) 
\  / 

where  x  =  ^  is  the  relative  composition  of  the  components.  Wo  use  this 
form  (1 .4)  because  it  turns  out  to  be  the  correct  one  in  the  general  case 
where  x  7*  %  and  because  it  is  analogous  to  Eq.  (2.2),  Chap.  XVII.  We 
note  that  for  w  =  0,  the  disordered  state,  Eq.  (1.4)  reduces  exactly  to 
Eq.  (2.2),  Chap.  XVII,  as  it  should.  To  find  the  energy  at  any  tempera- 

PT 

ture,  we  assume  as  in  Chap.  XVII  that  we  add  an  amount  I    Cp  dT,  where 

•/o 

Cp  is  the  specific  heat  of  a  completely  disordered  phase.  The  actual 
specific  heat  will  bo  different  from  this  CP,  because  w  in  Eq.  (1 .4)  will 
prove  to  depend  on  temperature,  giving  an  additional  term  in  the  deriva- 
tive of  energy  with  respect  to  temperature.  With  these  assumptions,  we 
then  have 

U  =  Uo  +  fJCpdT,  (1.5) 

vvhero  Uo  is  given  in  Eq.  (1.4). 

Next  we  consider  the  entropy.     We  have ^— -      atoms  a  and 

(1  —  w)N    .         ,        ,   ...  ,  (1  —  w)N    .  i  (1  +  w)Ar 

.^ — L —  atoms  6  on  lattice  a,  and ~ —  atoms  a  and  -  ' 

4  4  4 

atoms  6  on  lattice  ft.     The  number  of  ways  of  arranging  these  is 

w 

,2 


(1.6) 


296 


INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XVIII 


each  of  the  lattices  a  and  0  furnishing  an  equal  factor,  resulting  in  the 
square  in  Eq.  (1.6).    Using  Boltzmann's  relation,  this  leads  to  an  entropy 


S  -  -Nk(- 


+  w .    1  +  w      1  —  w  ,    1  —  w 

___         111  ~ -f"         rt m         =- 


TCP 


dT 


Nkr 


=  Nk  In  2  -  -^[(1  +  w)  In  (1  +  w)  +  (1  -  w)  In  (1  -  w)} 

T<£dT.     (1.7) 


f 

Jo 


T-Tc/2 


When  w  =  0,  the  second  term  of  Eq.  (1.7)  reduces  to  zero,  leaving 
S  =  Nk  In  2,  agreeing  with  the  value  of  Kq.  (2.6),  Chap.  XVII,  when  we 
sot  x  =  -J,  checking  the  correctness  of  Eq.  (1.7)  in  this  special  case. 

When  w  =  ±  1 ,  however,  the  expression 
(1.7)  reduces  to  zero,  showing  that  tho 
ordered  state  has  zero  entropy.  This  is 
as  we  should  expect;  there  is  only  one 
arrangement  of  the  atoms,  all  the  a's  being 
on  ono  lattice,  all  the  b's  on  the  other,  so 
that  then*  is  no  randomness  at  all. 

2.  Equilibrium  in  Transitions  of  the 
Cu-Zn  Type.— Having  found  the  internal 
energy  and  entropy  as  a  function  of  the 
degree  of  order  and  the  temperature,  in 
Eqs.  (1 .4)  and  (1 .7),  we  can  at  once  set  up 
the  free  energy,  and  find  which  value  of 
the   degree  of  order  gives   the   stablest 
phase  at  any  given  temperature.     In  Fig. 
/  XVIII-1  we  plot  the  Gibbs  free  energy  G 
KKJ.  xvin-i.  -Gibbs  free  energy  as  a  function  of  w,  for  various  tempcra- 

:IH  function  of  the  degree  of  order,  tulvs<  Of  course,  since  equal  positive  and 
for  various  temperatures.  .  „  , 

negative  values  of  w  really  correspond  to 

the  same  state,  the  curves  are  symmetrical  about  tho  line  w  =  0.  The, 
curves  are  drawn  on  the  assumption  that  #„?>  —  (Ean  +  Ebb)/2  is  negative. 
This  ease,  in  which  unlike  atoms  attract  each  other  more  than  like  atoms, 
rase  (c)  of  Fig.  XVII-1,  is  the  only  one  in  which  we  may  expect  the 
ordered  state  to  be  more  stable  than  the  disordered  ono.  For  if  like 
atoms  attract  more  than  unlike,  as  in  case  (a),  Fig.  XVII-1,  the  case  in 
which  atoms  tend  to  segregate  into  two  separate  phases,  we  shall  surely 
find  that  the  disordered  state,  iu  which  each  atom  has  on  the  average  four 
neighbors  of  the  same  kind  as  well  as  four  of  the  opposite  kind,  will  be 
more  stable  than  the  ordered  state,  where  all  neighbors  are  of  the  oppo- 
site kind,  even  at  the  absolute  zero. 


SEC.  2] 


PHASE  CHANGES  OF  THE  SECOND  ORDER 


297 


We  see  that  at  low  temperatures  the  minimum  of  tho  G  curve,  giving 
the  stable  phase,  comes  at  values  of  w  different  from  zero,  approaching 
w  =  ±1  as  the  temperature  approaches  zero.  As  the  temperature  rises, 
the  minima  move  inward  toward  w  =  0,  and  at  a  certain  temperature 
(Tc  in  the  figure),  there  is  a  double  minimum,  with  a  very  flat  curve,  at, 
w  =  0.  Above  this  temperature  there  is  a  single  minimum  at  w  =  0.  In 
other  words,  the  degree  of  order  gradually  decreases  from  perfect  order  at 
T  =  0,  to  complete  disorder  at  and  above  a  certain  temporal  uro  7\. 
This  limiting  temperature  corresponds  to  the  Curio  temperature  in  forro- 
magnetism,  and  by  analogy  it  is  of  ton  referred  to  as  tho  (  'urio  temporal  uro 
in  this  case  as  well.  To  got  tho  minimum  of  tho  curve,  tho  natural  thing 
is  to  differentiate  G  with  respect  to  w,  keeping  T  constant.  Then  wo  have1 


o  = 


in  - 


(2.  i 


Equation  (2.1)  is  a  transcendental  equation  for  w  and  cannot  bo  solved 
explicitly.     We  can  easily  solve  it  graphically,  however,  using  the  form 


0  +  «>)  _ 
—         " 


8  Eaa  +  E,,, 

"  "*  -----  2 


l-w 


Wo  plot  In  (1  +  w)/(\  —  w)  as  a  func- 
tion   of    w,    and    on    the    same    graph 
draw  the  straight  line  w(-8/kT)[Kalt  -       2 
(Eatl  +  ATM,)/2j.     The  intersections  give 
the  required  value  of  ir.     As  we  see  from      1 
Fig.  XVIII-2,  at  low  temperatures  the 
straight  line  is  steep  and  there  will  be      0 
tliree  intersections,  one  at  w  =  0  (evi- 
dently corresponding  to  the  maximum 
of    the    curve,    as    wo    see    from    Fig.    "I 
XVIII-1)    and    two    others,    which   wo 
desire,  at  equal   positive  and  negative 
values  of  w.     As   the   temperature  in- 
creases and   the  slope  of   the4   straight 
line  decreases,  those  intersections  move   Flu-  XVlil-a.    Craphiral  solution  of 
toward  w  =  0  and  finally  coalesce  when 

the  slope  of  the  straight  line  equals  that  of  In  (1  +  w)/(\  —  w)  at  the 
origin.     Now 

(1  +w) 


-2  - 


In 


(1  -w) 


starts  out  from  the  origin  like  2w,  with  a  slope  2,  so  that  for  the  Curie 
point  we  must  have 


298  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XVIII 

Egg    +    Ebb 

=     ' 

(2.3) 


In  terms  of  this,  the  straight  line  in  Fig.  XVIII-2  is  2Tcw/T. 

By  the  graphical  method  of  Eq.  (2.2)  and  Fig.  XVIII-2,  the  curve 

of  Fig.  XVIII-3  is  obtained  for  the  stable  value  of  w,  as  a  function  of 

temperature.  This  shows  the  decrease 
of  w  from  1  to  0,  first  very  gradual,  then 
as  the  Curie  point  is  approached  very 


w  ""  x  rapid,  so  that  the  curve  actually  has  a 

vertical  tangent  at  the  Curie  point. 
The  curve  of  Fig.  XVIII-3  cannot  be 
T/Tc  expressed  analytically,  though  it  can  be 

FIG.   XVIII-3.— Degree  of  order  as  approximated  in  the  two  limits  of  T  =  0 

function  of  temperature.  i   71  __    //r 

Having  found  the  variation  of  w  with  temperature,  we  can  find  the 
specific  heat  anomaly,  or  the  excess  of  specific  heat  over  the  value  CP 
characteristic  of  disorder.  This  excess  is  evidently 

dU\   dw 


dwrdT 


E*«  +  E»\fa 
--  g  -  )df 


NkT  .     (1  +w)dw 

2          (1  -  w)dT 

-  -NkTcw^  (2.4) 

using  Eqs.  (1.4),  (1.7),  (2.2),  and  (2.3).  In  Eq.  (2.4),  it  is  understood 
that  dw/dT  is  the  slope  of  the  curve  of  Fig.  XVIII-3  and  that  it  is  to  be 
determined  graphically.  Since  the  slope  is  negative,  the  excess  specific 
heat  is  positive.  We  give  the  resulting  curve  for  specific  heat  in  Fig. 
XVIII-4,  where  we  see  that  it  comes  to  a  sharp  peak  at  the  Curie  point 
and  above  that  point  drops  to  zero. 

From  the  discussion  we  have  given,  it  is  plain  that  the  change  from 
the  ordered  to  the  disordered  state  occupies  the  whole  temperature  range 
from  zero  degrees  to  Te,  though  it  is  largely  localized  at  temperatures 
slightly  below  Tc.  Thus  this  change,  a  gradual  one  occurring  over  a 
large  temperature  range,  is  just  of  the  sort  that  we  wish  to  call  a  phase 


SEC.  2] 


PHASE  CHANGES  OF  THE  SECOND  ORDER 


299 


change  of  the  second  order.  We  can  make  the  situation  clearer  by 
plotting  curves  for  G  as  a  function  of  T.  We  do  this  for  a  number  of 
values  of  w,  ranging  from  zero  to  unity.  In  a  sense,  we  may  consider  that 
we  have  a  mixture  of  an  infinite  number  of  phases,  corresponding  to  the 
continuous  range  of  w,  and  at  each  temperature  that  particular  phase 
(or  particular  w)  will  be  stable  whose  curve  of  G  against  T  lies  lowest. 
The  resulting  curves  are  shown  in  Fig.  XVIII-5.  To  mako  thorn  clearer, 

1.5 


FIG.  XVI 11-4. — Kxc-css  specific*  heat  arising  from  the  ordered  state,  in  units  of  Nk,  as  func- 
tion of  temperature. 


W=l 


FIG.  XVIII-5. — Gibhs  free  energy  as  function  of  temperature,  for  different  degrees 
of  order,  in  the  order-disorder  transition.  The  envelope  of  the  straight  lines  represents 
the  free  energy  of  the  stable  state. 

we  leave  out  the  terms  coming  from  the  specific  heat  Cp  of  the  disordered 
state,  which  are  common  to  curves  for  all  w's,  and  do  not  affect  the 
,  relative  positions  of  the  curves.  When  this  is  done,  the  curves  become 
straight  lines,  since  the  internal  energy  and  entropy  are  then  independent 
of  temperature.  At  the  absolute  zero,  the  lowest  curve  is  the  one  with 
the  lowest  internal  energy  or  the  ordered  state.  The  disordered  states 
have  greater  entropy,  however,  even  at  the  absolute  zero,  so  that  their 
curves  slope  down  more,  and  at  higher  temperatures  their  free  energies 


300 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XVIII 


lie  lower  than  that  of  the  ordered  state.  From  Fig.  XVIII-5  we  see  that 
there  is  an  envelope  to  the  curves,  and  this  envelope  represents  the  actual 
curve  of  G  vs.  T7,  whose  slope  is  the  negative  entropy  and  whose  second 
derivative  gives  the  specific  heat.  The  particular  value  of  w  whose  curve 
is  tangent  to  the  envelope  at  any  temperature  is  the  stable  w  at  that 
temperature,  as  given  by  Fig.  XVIII-3. 

Graphs  like  Fig.  XVIII-5  show  particularly  plainly  the  difference 
between  phase  changes  of  the  first  and  second  order.  We  can  readily 
imagine  that,  by  slightly  altering  the  mathematical  details,  the  curves 
could  be  changed  to  the  form  of  Fig.  XVIII-6,  in  which,  though  we  have  a 


w=I 


FIG.  XV11I-0.  -  Gibbs  free  energy  as  function  of  temperature,  for  different  degrees  of 
order,  in  a  phase  change  of  the  first  order,  in  which  the  ordered  state  is  stable  below  a 
temperature  T0,  the  completely  disordered  state  above  this  temperature. 

continuous  set  of  phases  from  w  =  0  to  w  =  1,  the  envelope  lies  aboye 
rather  than  below  the  axis  of  abscissas.  In  this  case  the  stable  state 
is  that  with  w  =  1  up  to  a  certain  temperature,  w  =  0  from  there  on,  all 
other  values  of  w  corresponding  to  states  that  are  never  stable.  This 
would  then  be  a  phase  change  of  the  first  order,  as  shown  in  Fig.  XVI-2, 
with  a  discontinuity  in  the  slope  of  the  G  vs.  T  curve,  or  the  entropy,  and 
hence  with  a  latent  heat.  When  we  see  the  small  geometrical  difference 
between  these  two  cases,  we  see  that  in  some  cases  the  distinction  between 
phase  changes  of  the  first  and  the  second  order  is  not  very  fundamental. 
In  this  connection,  it  is  interesting  to  note  that  the  rotation  vibration 
transition  in  NH4C1,  which  we  mentioned  in  a  preceding  paragraph,  is 
clearly  a  phase  change  of  the  second  order,  the  change  occurring  through 
a  considerable  range  of  temperature  or  pressure.  However,  there  is  a 
similar  transition  in  NH4Br,  undoubtedly  due  to  the  same  physical  cause, 
which  at  least  at  high  pressure  takes  place  so  suddenly  that  it  certainly 
seems  to  be  a  phase  change  of  the  first  order.  This  is  probably  a  case 


SEC.  3]  PHASE  CHANGES  OF  THE  SECOND  ORDER  301 

where  the  distinction  is  no  more  significant  than  in  Figs.  XVIII-5  and 
XVIII-6.  We  must  not  forget,  however,  that  there  is  one  real  and 
definite  distinction  between  most  phase  changes  of  the  first  order  and  all 
those  of  the  second  order:  in  every  phase  change  of  the  second  order,  we 
must  be  able  to  imagine  a  continuous  range  of  phases  between  the  two 
extreme  ones  under  discussion,  while  in  a  phase  change  of  the  first  order 
this  is  not  necessary  (though,  as  we  have  seen  in  Fig.  XVIII-6,  it  can 
sometimes  happen),  and  in  the  groat  majority  of  cases  it  is  not  possible. 
3.  Transitions  of  the  Cu-Zn  Type  with  Arbitrary  Composition. — It  is 
not  much  harder  to  discuss  the  general  case  of  arbitrary  composition 
than  it  is  the  simple  case  of  50  per  cent  concentration  taken  up  in  the 
two  preceding  sections.  We  assume  that  there  are  NX  atoms  a,  JV(1  —  x) 
6's,  and  we  shall  limit  ourselves  to  the  case  where  x  is  loss  than  \\  the 
same  formulas  do  not  hold  for  x  greater  than  ^,  but  to  get  this  caso  we  can 
merely  interchange  the  names  of  substances  a  and  b.  As  before,  we  lot 
the  degree  of  order  bo  w.  Then  we  assume 

N 
Number  of  atoms  a  on  lattice  a  =  ^  (1  +  w)x 

N 

Number  of  atoms  I)  on  lattice  a  =  -~[l  —  (1+  w)x] 

Z 

N 
Number  of  atoms  a  on  lattice  ft  =  -^(1  —  w)x 

N 

Number  of  atoms  6  on  lattice  ft  =  -0-[1  —  (1  —  w)x].          (3.1) 

& 

To  justify  these  assumptions,  we  note  that  they  lead  to  the  correct  num- 
bers in  the  three  cases  w  =  0,  ±1,  and  that  they  give  the  numbers  as 
linear  functions  of  w,  conditions  which  determine  Eqs.  (3.1)  uniquely. 
Then  for  the  numbers  of  pairs  we  find 

Number  of  pairs  aa:  4/Vz2(l  —  w*) 

Number  of  pairs  bb:  4JV[(1  —  :r)2  —  x'2w~] 

Number  of  pairs  ah:  ZN[x(l  -  x)  +  x*w*l  (3.2) 

and  for  the  internal  energy  we  have 

U  =  4N\  xEaa  +  (1  -  x)Ebb 

+  8Nx*w{Eab  -  E-^E^\  +   (TCP  dT.     (3.3) 
\  ^        /       Jo 

The  steps  in  the  derivation  of  Eqs.  (3.2)  and  (3.3)  have  not  been  given 
above,  but  the  principles  used  in  their  derivation  are  just  like  those  used 
in  Sec.  1.  We  note  that  Eq.  (3.3)  is  the  one  already  written  in  Eqs.  (1.4) 
and  (1.5)  but  previously  justified  only  for  the  case  x  =  £. 


302 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XV1U 


The  derivation  of  the  entropy  is  also  exactly  analogous  to  that  of  Sec.  1 
and  the  result  is 


f 

S  -   --/{(I  +  w)x  In  (1  +  w)x  +  (1  -  w)x  In  (1  -  w)x 
Z 

}-  fl  -  (1  +  w)x]  In  [1  -  (1  +  w)x]  +  [1  -  (1  -  w)x]  In  [1  -  (1  -  w)x\\ 

dT.     (3.4) 


It  is  easy  to  verify  that  in  the  case  x  =  £  this  leads  to  the  value  already 
found  in  Eq.  (1.7).  From  Eqs.  (3.3)  and  (3.4)  we  can  find  tho  froo 
onorgy  and  carry  out  the  same  sort  of  discussion  that  we  havo  above,  but, 
for  any  concentration.  To  find  the  value  of  w  for  tho  stable  state,  at  any 
value  of  x,  we  differentiate  0  with  respect  to  w  and  sot  it  equal  to  zoro. 

Then  we  havo 


Q  = 


x  In 


O  -w)[l  -  (1  +»)*]' 


or 


In  °  + 

ln  ~(i~-  w)  [i 


[1  -  (1  - 


-  (1+ 


whore  the  Tc  used  in  Eq.  (3.6)  is  the  ono 
defined  in  Eq.  (2.3),  holding  for  the  concon- 
x=0J  ^/\    Cation    x  —  i..     Equation    (3.6)    can    bo 

solved  as  in  the  special  case  x  =  £,  plotting 
the  left  side  of  Eq.  (3.6)  against  w  and 
finding  the  intersection  with  the  straight 
line  given  by  the  right  side.  Qualitatively 
we  find  the  same  sort  of  result  as  in  our 
previous  case,  the  dogree  of  order  going 
from  unity  at  absolute  zero  to  zero  at  a 
compositions,  T  =>  Curie  point.  The  Curie  point,  however, 
depends  on  concentration.  We  find  it,  as 
before,  by  lotting  the  slope  of  the  straight  line  representing  the  right  side 
of  Ea.  (3.6)  be  the  same  as  the  slope  of  the  left  side  at  the  origin,  which  is 
2/(l  —  r).  Equating  these,  we  have 


0 
w 


tor   different 
°'8  Ti' 


Tcx  = 


(3.7) 


where  Tcx  is  the  Curie  temperature  for  concentration  x,  Tc  for  concentra- 
tion x  =  4.     From  Eq.  (3.7)  we  see  that  Tc*  is  a  parabolic  function  of 


SEC.  3] 


PHASE  CHANGES  OF  THE  SECOND  ORDER 


303 


z,  having  its  maximum  for  x  =  £,  and  falling  to  zero  at  the  extreme 
concentration  x  =  0.  This  is  Of  metallurgical  interest,  for  on  phase 
diagrams  in  cases  where  there  is  a  transition  of  the  second  order  it  is  quite 
common  to  draw  a  line  of  Curie  temperature  vs.  composition  to  indicate 
the  transition,  though  thore  is  no  real  equilibrium  of  phases  to  be  indicated 
by  it.  In  a  case  like  the  Cu-Zn  transition,  this  curve  should  theoretically 
have  the  form  (3.7).  The  experimental  data  are  hardly  good  enough  to 
soo  whether  this  is  verified  or  not. 


w=08- 


0.1 


02 


03 


0.4 


05 


FIG.  XVIII-8.     Gibbs  free  energy  as  function  of  composition,  for  different  decrees  ot 

order,  T  =  0.8   Te. 

At  a  temperature  below  the  Curie  point  TC1  it  is  plain  from  Eq.  (3.7) 
that  for  concentrations  nearer  £  than  a  certain  critical  concentration  the 
alloys  will  be  below  thoir  Curie  points  and  will  be  in  partly  ordered 
phases,  while  for  x  loss  than  this  critical  concentration  they  will  bo  above* 
their  Curie  points  and  will  bo  in  the  disordered  state.  This  is  indicated 
in  Fig.  XVIII-7,  where  wo  show  G  as  a  function  of  w  for  different  values  of 
x,  at  a  temperature  of  0.8  Tc.  The  critical  concentration  for  this  tem- 
perature is  0.277,  as  can  bo  found  at  once  from  Eq.  (3.7);  it  is  noted  in 
Fig.  XVIII-7  that  the  curves  for  x  =  0.1  and  0.2  definitely  have  thoir 
minima  at  w  =  0,  indicating  complete  disorder,  while  that  for  x  =  0.3 
is  very  flat  at  the  center,  and  those  for  0.4  and  0.5  definitely  have  minima 
for  w  j±  0,  indicating  a  partly  ordered  state.  Finally,  in  Fig.  XVIII-8 
we  show  G  as  a  function  of  x,  for  different  values  of  w,  at  this  same  tern- 


304  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XVIII 

perature  T  =  0.8  Tc.  For  compositions  up  to  0.277,  as  we  have  men- 
tioned, the  curve  for  w  —  0  lies  the  lowest.  At  higher  concentrations,  the 
other  curves  begin  to  cross  it  and  the  stable  state  corresponds  to  the 
envelope  of  these  curves,  the  lowest  w  rising  from  w  =  0  to  a  maximum 
of  about  w  =  0.80  at  x  =  \.  This  envelope  is  of  interest,  for  it  is  the 
curve  of  G  vs.  x  which  should  really  be  used  to  represent  the  stable  stato 
in  such  a  system  and  which  should  bo  usod  in  investigating  the  equilibrium 
between  this  phase  and  other  phases,  in  the  manner  of  Chap.  XVII.  We 
notice  that  this  envelope  is  a  smooth  curve,  convex  downward,  just  as 
the  curve  for  w  —  0  is,  and  in  fact  it  does  not  greatly  differ  from  that  for 
w  =  0.  Thus  our  discussion  of  phase  equilibrium  of  the  preceding 
chapter,  where  we  entirely  neglected  the  order-disorder  transition,  is  not 
seriously  in  error  for  a  phase  in  which  such  a  transition  is  possible.  The 
reason  is  that,  though  there  is  a  considerable  difference  in  energy  and 
entropy  separately  between  the  ordered  and  disordered  states,  these  make 
contributions  of  opposite  sign  in  the  free  energy,  so  that  it  is  only  slightly 
affected  by  the  degree  of  order. 


PART  III 
ATOMS,  MOLECULES,  AND  THE  STRUCTURE  OF  MATTER 


CHAPTER  XIX 
RADIATION  AND  MATTER 

In  the  development  of  quantum  theory,  light,  or  electromagnetic 
radiation  of  visible  wave  lengths,  has  had  a  very  special  place.  It  was 
the  study  of  black-body  radiation  that  first  showed  without  question  the 
inadequacy  of  classical  mechanics,  and  that  led  Planck  to  the  quantum 
theory.  One  of  the  first  triumphs  of  quantum  theory  was  Einstein's 
prediction  of  the  law  of  photoelectric  emission,  a  prediction  which  was 
beautifully  verified  by  experiment.  And  in  the  development  of  the 
theory  of  atomic  and  molecular  structure,  the  most  complicated  and 
involved  test  which  has  yet  been  given  the  quantum  theory,  the  tool  has 
been  almost  entirely  optical,  the  spectrum,  the  light  emitted  and  absorbed 
by  matter.  Some  of  the  most  difficult  logical  concepts  of  the  quantum 
theory  have  come  in  the  field  of  light.  The  difficulty  of  reconciling  prob- 
lems like  interference  of  light,  which  clearly  indicate  that  it  is  an  electro- 
magnetic wave  motion,  with  problems  like  the  photoelectric  effect,  which 
equally  clearly*  indicate  that  it  is  made  of  individual  particles  of  energy, 
or  photons,  is  well  known.  And  these  difficulties,  indicating  that  light 
really  has  a  sort  of  dual  nature,  gave  the  suggestion  that  matter  might 
have  a  dual  nature  too,  and  that  the  particles  with  which  we  were  familiar 
might  also  be  associated  with  waves.  This  was  the  suggestion  which 
led  to  wave  mechanics  and  which  raised  the  quantum  theory  from  a 
rather  arbitrary  set  of  rules  to  a  well-developed  branch  of  mathematical 
physics. 

Throughout  the  development  of  modern  ideas  of  light,  black-body 
radiation  has  played  an  essential  role.  This  is  simply  light  in  thermal 
equilibrium — the  distribution  of  frequencies  and  intensities  of  light 
which  is  in  equilibrium  with  matter  at  a  given  temperature.  Our  study 
in  this  chapter  will  be  of  black-body  radiation,  and  we  shall  handle  it  by 
direct  thermodynamic  methods,  using  the  quantum  theory  much  as  we  did 
in  the  theory  of  specific  heats.  In  the  following  chapter  we  shall  take  up 
the  kinetics  of  radiation,  the  probabilities  of  emission  and  absorption  of 
light  by  matter.  This  will  lead  us  to  a  kinetic  derivation  of  the  laws  of 
black  body  radiation,  and  at  the  same  time  to  a  usable  method  of  handling 
the  kinetics  of  radiation  problems  out  of  equilibrium,  which  we  very 
commonly  meet  in  the  laboratory. 

1.  Black-body  Radiation  and  the  Stefan-Boltzmann  Law. — Light  is 
simply  electromagnetic  radiation,  a  wave  motion  in  space,  in  which  the 

307 


,308  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIX 

electric  and  the  magnetic  fields  oscillate  rapidly  with  time.  It  can  carry 
energy,  just  as  sound  or  any  other  wave  can  carry  energy.  We  are  all 
familiar  with  this;  most  of  the  available  energy  on  the  earth  was  carried 
here  from  the  sun,  by  electromagnetic  radiation.  Like  all  waves,  it  can 
be  analyzed  into  sinusoidal  or  monochromatic  waves,  in  which  the 
electromagnetic  field  oscillates  sinusoidally  with  time,  with  a  definite 
frequency  v\  the  possibility  of  such  an  analysis  is  a  mathematical  one, 
based  on  Fourier's  series,  and  does  not  imply  anything  about  the  physics 
of  radiation.  The  velocity  of  light,  at  least  in  empty  space,  is  inde- 
pendent of  the  frequency  of  oscillation,  and  is  ordinarily  denoted  by  c, 
equal  to  2.998  X  1010  cm.  per  second.  We  can  associate  a  wave  length 
with  each  frequency  of  oscillation,  by  the  equation 

\v  =  c,  (1.1) 

where  A  is  the  wave  length.  The  mathematics  of  the  light  waves  is 
essentially  like  that  of  sound  waves,  given  in  Sec.  2  of  Chap.  XIV,  and 
we  shall  not  repeat  it  here.  In  that  section,  however,  we  found  that 
elastic  waves  were  of  two  sorts,  longitudinal  and  transverse.  Light  on 
the  contrary  is  only  transverse,  with  two  possible  planes  of  polarization,  or 
directions  for  the  electric  or  magnetic  field,  at  right  angles  to  the  direction 
of  propagation.  We  ordinarily  deal,  in  discussions  like  the  present,  with 
fairly  short  wave  lengths  of  light.  The  long  waves,  as  -found  in  radio 
transmission,  are  of  small  significance  thermodynamically  or  in  atomic 
structure.  Among  waves  shorter,  say,  than  a  tenth  of  a  millimeter,  it  is 
customary  to  speak  of  those  longer  than  7000  A  as  infrared  or  heat  waves, 
those  between  7000  and  4000  A  as  light  (since  the  eye  can  see  only  these 
wave  lengths),  those  between  4000  A  and  perhaps  50  A  as  ultraviolet,  and 
those  shorter  than  50  A  but  longer  than  perhaps  0.01  A  as  x-rays.  Waves 
shorter  than  this  are  hardly  met  in  ordinary  thermodynamic  or  atomic 
processes,  though  of  course  they  are  essential  in  nuclear  processes  and 
cosmic  rays.  Although  there  is  this  classification  of  wave  lengths,  it  is 
purely  a  matter  of  convenience,  and  we  shall  not  have  to  bother  with  it. 
For  our  purposes,  we  may  consider  as  light  any  radiation  from  perhaps 
^  mm.  to  TV  A;  it  is  only  in  this  range  that  the  radiations  we  consider  are 
likely  to  have  appreciable  intensity. 

Ordinary  bodies  at  any  temperature  above  the  absolute  zero  auto- 
matically emit  radiation,  and  are  capable  of  absorbing  radiation  falling  on 
them.  Thus  an  enclosure  containing  bodies  at  a  temperature  above  the 
absolute  zero  cannot  be  in  equilibrium  unless  it  contains  radiation  as 
well.  In  fact,  in  equilibrium,  there  must  be  just  enough  radiation  so  that 
each  square  centimeter  of  surface  of  each  body  emits  just  as  much  radia- 
tion as  it  absorbs.  It  seems  clear  that  there  must  be  a  definite  sort  of 
radiation  in  equilibrium  with  bodies  at  a  definite  temperature.  For  we 


SBC.  1]  RADIATION  AND  MATTER  309 

know  that  all  bodies  at  a  given  temperature  are  in  thermal  equilibrium 
with  each  other,  and  if  they  are  all  in  a  container  with  the  same  thermal 
radiation,  this  radiation  must  be  in  equilibrium  with  each  body  separately, 
and  must  hence  be  independent  of  the  particular  type  of  body,  and 
characteristic  only  of  the  temperature,  and  perhaps  the  volume,  of  the 
container.  It  is  an  experimental  fact,  one  of  the  first  laws  of  temperature 
radiation,  that  the  type  of  radiation — its  wave  lengths,  intensities,  and 
so  on — is  independent  of  the  volume,  depending  only  on  the  temperature. 
This  type  of  radiation,  in  thermal  equilibrium,  is  called  black-body 
radiation,  for  a  reason  which  we  shall  understand  in  a  moment. 

The  first  and  most  elementary  law  of  black-body  radiation  is  Kirch- 
hoff's  law,  a  simple  application  of  the  kinetic  method.  To  understand 
it,  we  must  define  some  terms.  First  we  consider  the  emissive  power 
e\  of  a  surface.  We  consider  the  total  number  of  ergs  of  energy  emitted 
in  the  form  of  radiation  per  second  per  square  centimeter  of  a  surface,  in 
radiation  of  wave  length  between  X  and  X  +  d\,  and  by  definition  set  it 
equal  to  e\d\.  Next  we  consider  the  absorptivity.  Suppose  a  certain 
amount  of  radiant  energy  in  the  wave  length  range  d\  falls  on  1  sq.  cm. 
per  second,  and  suppose  a  fraction  a\  is  absorbed,  the  remainder,  or 
(1  —  a\),  being  reflected.  Then  a\  is  called  the  absorptivity,  and 
(1  —  a\)  is  called  the  reflectivity.  Now  consider  the  simple  requirement 
for  thermal  equilibrium.  We  shall  demand  that,  in  each  separate  range 
of  wave  lengths,  as  much  radiation  is  absorbed  by  our  square  centimeter 
in  thermal  equilibrium  as  is  radiated  by  it.  This  assumption  of  balance 
in  each  range  of  wave  lengths  is  a  particular  example  of  the  principle 
of  detailed  balancing  first  introduced  in  Chap.  VI,  Sec.  2.  Now  let 
I\d\  be  the  amount  of  black-body  radiation  falling  on  1  sq.  cm.  per  second 
in  the  wave  length  range  dX.  This  is  a  function  of  the  wave  length  and 
temperature  only,  as  we  have  mentioned  above.  Then  we  have  the 
following  relation,  holding  for  1  sq.  cm.  of  surface: 

Energy  emitted  per  second  =  e\d\ 

=  energy  absorbed  per  second  =  I\a\d\, 

or 

—  =  I\  =  universal  function  of  X  and  T.  (  1.2) 

G&X 

Equation  (1.2)  expresses  Kirchhoff  s  law:  the  ratio  of  the  emissive  power 
to  the  absorptivity  of  all  bodies  at  the  same  wave  length  and  temperature 
is  the  same.  Put  more  simply,  good  radiators  are  good  absorbers,  poor 
radiators  are  poor  absorbers.  There  are  many  familiar  examples  of  this 
law.  One,  which  is  of  particular  importance  in  spectroscopy,  is  the 
following :  if  an  atom  or  other  system  emits  a  particularly  large  amount  of 


310  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIX 

radiation  at  one  wave  length,  as  it  will  do  if  it  has  a  line  spectrum,  it 
must  also  have  a  particularly  large  absorptivity  at  the  same  wave  length, 
so  that  a  continuous  spectrum  of  radiation,  passing  through  the  body,  will 
have  this  wave  length  absorbed  out  and  will  show  a  dark  line  at  that  point. 

A  black  body  is  by  definition  one  which  absorbs  all  the  light  falling 
on  it,  so  that  none  is  reflected.  That  is,  its  absorptivity  a\  is  unity,  for 
all  wave  lengths.  Then  it  follows  from  Eq.  (1 .2)  that  for  a  black  body  the 
emissive  power  ex  is  equal  to  /x,  the  amount  of  black-body  radiation 
falling  on  1  sq.  cm.  per  second  per  unit  range  of  wave  length.  We  can 
understand  the  implications  of  this  statement  better  if  we  consider  what 
is  called  a  hollow  cavity.  This  is  an  enclosure,  with  perfectly  opaque 
walls,  so  that  no  radiation  can  escape  from  it.  It  contains  matter  and 
radiation  in  equilibrium  at  a  given  temperature.  Thus  the  radiation  is 
black-body  radiation  characteristic  of  that  temperature.  Now  suppose 
we  make  a  very  small  opening  in  the  enclosure,  not  big  enough  to  disturb 
the  equilibrium  but  big  enough  to  let  a  little  radiation  out.  We  can 
approximate  the  situation  in  practice  quite  well  by  having  a  well  insulated 
electric  furnace  for  the  cavity,  with  a  small  window  for  the  opening.  All 
the  radiation  falling  on  the  opening  gets  out,  so  that  if  we  look  at  what 
emerges,  it  represents  exactly  the  black-body  radiation  falling  on  the  area 
of  the  opening  per  second.  Such  a  furnace  makes  in  practice  the  moist 
convenient  way  of  getting  black-body  radiation.  But  now  by  Kirchhoff  's 
law  and  the  definition  of  a  black  body,  we  see  that  if  we  have  a  small  piece 
of  black  body,  of  the  shape  of  the  opening  in  our  cavity,  and  if  we  heat  it 
to  the  temperature  of  the  cavity,  it  will  emit  exactly  the  same  sort  of 
radiation  as  the  opening  in  the  cavity.  This  is  the  reason  why  our  radia- 
tion from  the  cavity,  radiation  in  thermal  equilibrium,  is  also  called  black- 
body  radiation.  A  black  body  is  the  only  one  which  has  this  property 
of  emitting  the  same  sort  of  radiation  as  a  cavity.  Any  other  body  will 
emit  an  amount  a\I\  d\  per  square  centimeter  per  second  in  the  range 
d\,  and  since  a\  must  by  definition  be  less  than  or  equal  to  unity,  the 
other  body  will  emit  less  light  of  each  wave  length  than  a  black  body.  A 
body  which  has  very  small  absorptivity  may  emit  hardly  anything.  Thus 
quartz  transmits  practically  all  the  light  that  falls  on  it,  without  absorp- 
tion. When  it  is  heated  to  a  temperature  at  which  a  metal,  for  instance, 
would  be  red  or  white  hot  and  would  emit  a  great  deal  of  radiation,  the 
quartz  emits  hardly  any  radiation  at  all,  in  comparison. 

Now  that  we  understand  the  emissive  power  and  absorptivity  of 
bodies,  we  should  consider  /x,  the  universal  function  of  wave  length  and 
temperature  describing  black-body  radiation.  It  is  a  little  more  con- 
venient not  to  use  this  quantity,  but  a  closely  related  one,  u*.  This 
represents,  not  the  energy  falling  on  1  sq.  cm.  per  second,  but  the  energy 
contained  in  a  cubic  centimeter  of  volume,  or  what  is  called  the  energy 


SEC.  1]  RADIATION  AND  MATTER  31 1 

density.  If  there  is  energy  in  transit  in  a  light  beam,  it  is  obvious  that 
the  energy  must  be  located  somewhere  while  it  is  traveling,  and  that  we 
can  talk  about  the  amount  of  energy,  or  the  number  of  ergs,  per  cubic 
centimeter.  It  is  a  simple  geometrical  matter  to  find  the  relation  between 
the  energy  density  and  the  intensity.  If  we  consider  light  of  a  definite 
direction  of  propagation,  then  the  amount  of  it  which  will  strike  unit 
cross  section  per  second  is  the  amount  contained  in  a  prism  whose  base 
is  1  sq.  cm.  and  whose  slant  height  along  the  direction  of  propagation  is 
the  velocity  of  light  c.  This  amount  is  the  volume  of  the  prism  (c  cos  0, 
if  0  is  the  angle  between  the  direction  of  propagation  and  the  normal  to 
the  surface),  multiplied  by  the  energy  of  the  light  wave  per  unit  volume. 
Thus  we  can  find  very  easily  the  amount  of  light  of  this  definite  direction 
of  propagation  falling  on  1  sq.  cm.  per  second,  if  we  know  the  energy 
density,  and  by  integration  we  can  find  the  amount  of  light  of  all  direc- 
tions falling  on  the  surface.  We  shall  not  do  it,  since  we  shall  not  need 
the  relation.  In  addition  to  this  difference  between  uv  and  7\,  the  former 
refers  to  frequency  rather  than  wave  length,  so  that  uvdv  by  definition  is 
the  energy  per  unit  volume  in  the  frequency  range  from  v  to  v  +  dv. 

In  addition  to  energy,  light  can  carry  momentum.  That  means  that 
if  it  falls  on  a  surface  and  is  absorbed,  it  transfers  momentum  to  the 
surface,  or  exerts  a  force  on  it.  This  force  is  called  radiation  pressure. 
In  ordinary  laboratory  experiments  it  is  so  small  as  to  be  very  difficult 
to  detect,  but  there  are  some  astrophysical  cases  where,  on  account  of  the 
high  density  of  radiation  and  the  smallness  of  other  forces,  the  radiation 
pressure  is  a  very  important  effect.  The  pressure  on  a  reflecting  surface, 
at  which  the  momentum  of  the  radiation  is  reversed  instead  of  just  being 
reduced  to  zero,  is  twice  that  on  an  absorbing  surface.  Now  the  radiation 
pressure  can  be  computed  from  electromagnetic  theory,  and  it  turns  out 
that  in  isotropic  radiation  (radiation  in  which  there  are  beams  of  light 
traveling  in  all  directions,  as  in  black-body  radiation),  the  pressure  against 
a  reflecting  wall  is  given  by  the  simple  relation 

P  =  £  X  energy  density 

\,dv.  (1.3) 


From  Eq.  (1.3)  we  can  easily  piove  a  law  called  the  Stefan-Boltzmann 
law  relating  the  density  of  radiation  to  the  temperature. 

Let  us  regard  our  radiation  as  a  thermodynamic  system,  of  pressure 
P,  volume  7,  and  internal  energy  [7,  given  by 

U  *  V*u,d*.  (1.4) 


Then,  from  Eqs.  (1.3)  and  (1.4),  the  equation  of  state  of  the  radiation  is 

PV  -  W,  (1.5) 


312  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIX 

which  compares  with  PV  «  ft/  for  a  perfect  gas.  There  is  the  further 
fact,  quite  in  contrast  to  a  perfect  gas,  that  the  pressure  depends  only  on 
the  temperature,  being  independent  of  the  volume.  Then  we  have 


W)T  *  3/>'  (L6) 

using  the  fact  that  P  is  independent  of  V.     But  by  a  simple  thormo- 
dynamic  relation  we  know  that  in  general 


-'• 

Combining  Eqs.  (1.6)  and  (1.7),  we  have 

-  (1.8) 

The  pressure  in  Eq.  (1.8)  is  really  a  function  of  the  temperature  only,  so 
that  the  partial  derivative  can  be  written  as  an  ordinary  derivative,  and 
we  can  express  the  relation  as 


(IJ) 


which  can  be  integrated  to  give 


In  P  =  4  In  T  +  const., 

P  =  const.  T\  (1.10) 

or,  using  Kq.  (1.3), 

f*uvdv  =  const.  T4.  (1.11) 

«/o 

Equation  (1.11),  stating  that  the  total  energy  per  unit  volume  is  propor- 
tional to  the  fourth  power  of  the  absolute  temperature,  is  the  Stefan- 
Boltzmann  law.  Since  the  intensity  of  radiation,  the  amount  falling  on  a 
square  centimeter  in  a  second,  is  proportional  to  the  energy  per  unit 
volume,  we  may  also  state  the  law  in  the  form  that  the  total  intensity  of 
black-body  radiation  is  proportional  to  the  fourth  power  of  the  absolute 
temperature.  This  law  is  important  in  the  practical  measurement  of 
high  temperatures  by  the  total  radiation  pyrometer.  This  is  an  instru- 
ment which  focuses  light  from  a  hot  body  onto  a  thermopile,  which 
absorbs  the  radiation  energy  and  measures  it  by  finding  the  rise  of  tem- 
perature it  produces.  The  pyrometer  can  be  calibrated  at  low  tempera- 
tures that  can  be  measured  by  other  means.  Then,  by  Stefan's  law,  at 
higher  temperatures  the  amount  of  radiation  must  go  up  as  the  fourth 


SEC.  2]  RADIATION  AND  MATTER  313 

power  of  the  temperature,  from  which  we  can  deduce  the  temperature 
of  very  hot  bodies,  to  which  no  other  method  of  temperature  measurement 
is  applicable.  Since  Stefan's  law  is  based  on  such  simple  and  fundamental 
assumptions,  there  is  no  reason  to  think  that  it  is  not  perfectly  exact,  so 
that  this  forms  a  valid  method  of  measuring  high  temperatures. 

2.  The  Planck  Radiation  Law.  —  The  Stefan-Boltzmann  law  gives  us  a 
little  information  about  the  function  uvj  but  not  a  great  deal.  We  shall 
next  see  how  the  function  can  be  evaluated  exactly.  There  are  many 
ways  of  doing  this,  but  the  first  way  we  shall  use  is  a  purely  statistical 
one.  We  can  outline  the  method  very  easily.  We  consider  a  hollow 
cavity  with  perfectly  reflecting  walls,  making  it  rectangular  for  con- 
venience. In  such  a  cavity  we  can  have  standing  waves  of  light;  these 
waves,  in  fact,  constitute  the  thermal  radiation.  There  will  be  a  discrete 
set  of  possible  vibrations  or  overtones  of  a  fundamental  vibration,  just 
as  we  had  a  discrete  set  of  sound  vibrations  in  a  rectangular  solid  in  Chap. 
XIV;  only  instead  of  having  a  finite  number  of  overtones,  as  we  did  with 
the  sound  vibrations  on  account  of  the  atomic  nature  of  the  material,  the 
number  of  overtones  here  is  infinite  and  stretches  up  to  infinite  fre- 
quencies. As  with  the  sound  vibrations,  each  overtone  acts,  as  far  as  the 
quantum  theory  is  concerned,  like  a  linear  oscillator.  Its  energy  cannot 
take  on  any  arbitrary  value,  but  only  certain  quantized  values,  multiples 
of  hv,  where  v  is  the  frequency  of  that  particular  overtone.  This  leads 
at  once  to  a  calculation  of  the  mean  energy  of  each  overtone,  just  as  we 
found  in  our  calculation  of  the  specific  heat  of  solids,  and  from  that  we 
can  find  the  mean  energy  per  unit  volume  in  the  frequency  range  dv,  and 
so  can  find  uv. 

First,  let  us  find  the  number  of  overtone  vibrations  in  the  range  dv. 
We  follow  Chap.  XIV  in  detail  and  for  that  reason  can  omit  a  great  deal 
of  calculation.  In  Sec.  2  of  that  chapter,  we  found  the  number  of  over- 
tones in  the  range  dp,  in  a  problem  of  elastic  vibration,  in  which  the 
velocity  of  longitudinal  waves  was  i>/,  that  of  transverse  waves  vt.  From 
Eqs.  (2.20)  and  (2.21)  of  that  chapter,  the  number  of  overtones  of  longi- 
tudinal vibration  in  the  range  di>,  in  a  container  of  volume  F,  was 


dN  =  4™»  di.  (2.1) 

vi 

and  of  transverse  vibrations 

dN  =  87TJ/2  drt-  (2.2) 


There  was  an  upper,  limiting  frequency  for  the  elastic  vibrations,  but 
as  we  have  just  stated  there  is  not  for  optical  vibrations.  In  our  optical 
case,  we  can  take  over  Eq.  (2.2)  without  change.  Light  waves  are  only 


314  INTRODUCTION  TO  CHEMICAL  PHYSICS          [€HAP.  XIX 

transverse,  so  that  the  overtones  of  Eq.  (2.1)  are  not  present  but  those  of 
Eq.  (2.2)  are.     Since  the  velocity  of  light  is  c,  we  have 


dN  -  8™2  d  (2.3) 

as  the  number  of  standing  waves  in  volume  V  and  the  frequency  range  dv. 
Next  we  want  to  know  the  moan  energy  of  each  of  these  standing 
waves  at  the  temperature  T.  Before  the  invention  of  the  quantum 
theory,  it  was  assumed  that  the  oscillators  followed  classical  statistics. 
Then,  being  linear  oscillators,  the  mean  energy  would  have  to  be  kT  at 
temperature  T.  From  this  it  would  follow  at  once  that  the  energy  density 
u,9  which  can  be  found  by  multiplying  dN  in  Eq.  (2.3)  by  the  mean  energy 
of  an  oscillator,  and  dividing  by  V  and  dv,  is 


,0  A. 
uv  =  —  ^—  -  (2.4) 

Kquation  (2.4)  is  the  so-called  Kayleigh-Jcans  law  of  radiation.  It  was 
derived,  essentially  as  we  have  done,  from  classical  theory,  and  it  is  the 
only  possible  radiation  law  that  can  be  found  from  classical  theory.  Yet 
it  is  obviously  absuru,  as  was  realized  as  soon  as  it  was  derived.  For  it 
indicates  that  uv  increases  continually  with  v.  At  any  temperature,  the 
farther  out  we  go  toward  the  ultraviolet,  the  more  intense  is  the  tempera- 
ture radiation,  until  finally  it  becomes  infinitely  strong  as  we  go  out 
through  the  x-rays  to  the  gamma  rays.  This  is  ridiculous;  at  low  tem- 
peratures the  radiation  has  a  maximum  in  the  infrared,  and  has  fallen 
practically  to  zero  intensity  by  the  time  we  go  to  the  visible  part  of  the 
spectrum,  while  even  a  white-hot  body  has  a  good  deal  of  visible  radiation 
(hence  the  "white  heat"),  but  very  little  far  ultraviolet,  and  practically 
no  x-radiation.  There  must  be  some  additional  feature,  missing  in  the 
classical  theory,  which  will  have  as  a  result  that  the  overtones  of  high 
frequency,  the  visible  and  even  more  the  ultraviolet  and  x-ray  frequencies, 
have  much  less  energy  at  low  temperatures  than  the  equipartition  value, 
and  in  fact  at  really  low  temperatures  have  practically  no  energy  at  all. 
But  this  is  just  what  the  quantum  theory  does,  as  we  have  seen  by  many 
examples.  The  examples,  and  the  quantum  theory  itself,  however,  were 
not  available  when  this  problem  first  had  to  be  discussed;  for  it  was  to 
remove  this  difficulty  in  the  theory  of  black-body  radiation  that  Planck 
first  invented  the  quantum  theory. 

In  Chap.  IX,  Sec.  5,  we  found  the  average  energy  of  a  linear  oscillator 
of  frequency  v  in  the  quantum  theory,  and  found  that  it  was 


A  v   .          v  /n  ,. 

Average  qnergy  =  -TT  +  ~rz  --  >  (2.5) 

»* 


BBC.  2] 


RADIATION  AND  MATTER 


315 


as  in  Eq.  (5.9),  Chap.  IX.  The  term  ?hv  was  the  energy  at  the  absolute 
zero  of  temperature;  the  other  term  represented  the  part  of  the  energy 
that  depended  on  temperature.  The  first  term,  sometimes  called  the 
zero-point  energy,  arose  because  we  assumed  that  the  quantum  condition 
was  En  =  (n  +  %)hi>,  instead  of  En  =  nhv.  We  can  now  use  the  expres- 
sion (2.5)  for  the  average  energy  of  an  overtone,  but  we  must  leave  out 
the  zero-point  energy.  For,  since  the  number  of  overtones  is  infinite,  this 
would  lead  to  an  infinite  energy  density,  even  at  a  temperature  of  absolute 
zero.  The  reason  for  doing  this  is  not  very  clear,  oven  in  the  present  state 
of  the  quantum  theory.  We  can  do  it  quite  arbitrarily,  or  we  can  say 
that  the  quantum  condition  should  be  En  =  nhv  (an  assumption,  how- 
ever, for  which  there  is  no  justification  in  wave  mechanics),  or  we  can 


FIG.  XIX-l.- 


-Enorgy  density  from  Planck's  distribution  law,  for  fom  temperatures  in  thw 
ratio  1.2.3.4. 


say  that  the  infinite  zero-point  energy  is  really  there  but,  since  it  is  inde- 
pendent of  temperature,  we  do  not  observe  it.  No  one  of  these  reasons  is 
very  satisfactory.  Unfortunately,  though  it  was  the  branch  of  physics 
in  which  quantum  theory  originated,  radiation  theory  still  has  more 
difficulties  in  it  than  any  other  parts  of  quantum  theory.  We  shall  then 
simply  apologize  for  leaving  out  the  term  ^hv  in  Eq.  (2.5),  and  shall  hope 
that  at  some  future  time  the  theory  will  be  well  enough  understood  so 
that  we  can  justify  it. 

If  we  assume  the  expression  (2.5),  without  the  zero-point  energy,  for 
the  average  energy  of  a  standing  wave,  and  take  Eq.  (2.3)  for  the  number 
of  standing  waves  in  volume  V  and  frequency  range  dvy  we  can  at  once 
derive  uv,  and  we  have 


uv 


1 


C3        *! 


(2.6) 


ekT  - 


316  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIX 

Equation  (2.6)  is  Planck's  radiation  law,  and  as  far  as  available  experi- 
ments show,  it  is  the  exactly  correct  law  of  black-body  radiation.  Curves 
of  uv  as  a  function  of  v,  for  different  temperatures,  are  shown  in  Fig. 
XIX-1.  At  low  frequencies,  even  for  room  temperatures,  the  frequencies 
are  so  low  that  the  energy  of  an  oscillator  has  practically  the  classical 
value,  and  the  Rayleigh-Jeans  law  is  correct.  At  higher  frequencies, 
however,  this  is  not  the  case,  and  the  curves,  instead  of  rising  indefinitely 
toward  high  frequencies,  curve  down  again  and  go  very  sharply  to  negligi- 
ble values.  The  maximum  of  the  curve  shifts  to  higher  frequencies  as 
the  temperature  rises,  checking  the  experimental  fact  that  bodies  look 
red,  then  white,  then  blue,  as  their  temperature  rises.  The  area  under 
the  curve  rises  rapidly  with  temperature.  It  is,  in  fact,  this  area  that 
must  be  proportional  to  the  fourth  power  of  the  temperature,  according 
to  the4  Stefari-Boltzmaiin  law.  We  can  easily  verify  that  Planck's  law  is 
in  accordance  with  that  law,  and  at  the  same  time  find  the  constant 
in  Eq.  (1.11),  by  integrating  uv  from  Eq.  (2.6).  We  have 

I     uv  dv  =    I     —~  -j- dv 

Jo  Jo  ekf  _  i 

=  ^TiS-4  f  YS)  V1    ~d(lf) 

/t  C/  I  (\       \lv  J.     I       \  Iv  JL     I 

«/°      \        /    ekT  _    ]     \        / 

4T4  r  «         ! 

c3~Jo   **  e*  -  1  x 

-,  (2.7) 


where  we  have  used  the  relation 


a  =  (l  +  ~  +  ~  +'••)  =  1.0823 


(2.8) 


3.  Einstein's  Hypothesis  and  the  Interaction  of  Radiation  and 
Matter.  —  To  explain  the  law  of  black-body  radiation,  Planck  had  to 
assume  that  the  energy  of  a  given  standing  wave  of  light  of  frequency  v 
could  be  only  an  integral  multiple  of  the  unit  hv.  Thus  this  carries  with 
it  a  remarkable  result:  the  energy  of  the  light  can  change  only  by  the  quite 
finite  amount  hv  or  a  multiple  of  it.  This  is  quite  contrary  to  what  the 
wave  theory  of  light  indicates.  The  theory  of  emission  and  absorption  of 
energy  has  been  thoroughly  worked  out,  on  the  wave  theory.  An  oscillat- 
ing electric  charge  has  oscillating  electric  and  magnetic  fields,  and  at 
distant  points  these  fields  constitute  the  radiation  field,  or  the  light  wave 


SEC.  3]  RADIATION  AND  MATTER  317 

sent  out  from  the  charge.  The  field  carries  energy  out  at  a  uniform  and 
continuous  rate,  and  the  charge  loses  energy  at  the  same  rate,  as  one  can 
see  from  the  principle  of  conservation  of  energy,  and  gradually  comes  to 
rest.  To  describe  absorption,  we  assume  a  light  wave,  with  its  alternating 
electric  field,  to  act  on  an  electric  charge  which  is  capable  of  oscillation. 
The  field  exerts  forces  on  the  charge,  gradually  setting  it  into  motion  with 
greater  and  greater  amplitude,  so  that  it  gradually  and  continuously 
absorbs  energy.  Both  processes,  emission  and  absorption,  then,  are 
continuous  according  to  the  wave  theory,  and  yet  the  quantum  theory 
assumes  that  the  energy  must  change  by  finite  amounts  hv. 

Einstein,  dearly  understanding  this  conflict  of  theories,  made  an 
assumption  that  seemed  extreme  in  1905  when  he  made  it,  but  which  has 
later  come  to  point  the  whole  direction  of  development  of  quantum 
theory.  He  assumed  that  the  energy  of  a  radiation  field  could  not  be 
considered  continuously  distributed  through  space,  as  the  wave  theory 
indicated,  but  instead  that  it  was  carried  by  particles,  then  called  light 
quanta,  now  more  often  called  photons,  each  of  energy  hv.  If  this  hypo- 
thesis is  assumed,  it  becomes  obvious  that  absorption  or  emission  of  light 
of  frequency  v  must  consist  of  the  absorption  or  emission  of  a  photon,  so 
that  the  energy  of  the  atom  or  other  system  absorbing  or  emitting  it  must 
change  by  the  amount  hv.  Einstein's  hypothesis,  in  other  words,  was  the 
direct  and  straightforward  consequence  of  Planck's  assumption,  and  it 
received  immediate  arid  remarkable  verification  in  the  theory  of  the 
photoelectric  effect. 

Metals  can  emit  electrons  into  empty  space  at  high  temperatures  by 
the  thermionic  effect  used  in  obtaining  electron  emission  from  filaments  in 
vacuum  tubes.  But  metals  also  can  emit  electrons  at  ordinary  room 
temperature,  if  they  are  illuminated  by  the  proper  light;  this  is  called  the 
photoelectric  effect.  Not  much  was  known  about  the  laws  of  photo- 
electric emission  in  1905,  but  Einstein  applied  his  ideas  of  photons  to  the 
problem,  with  remarkable  results  that  proved  to  be  entirely  correct. 
Einstein  assumed  that  light  of  frequency  v,  falling  on  a  metal,  could  act 
only  as  photons  hv  were  absorbed  by  the  metal.  If  a  photon  was 
absorbed,  it  must  transfer  its  whole  energy  to  an  electron.  Then  the 
electron  in  question  would  have  a  sudden  increase  of  hv  in  its  energy. 
Now  it  requires  a  certain  amount  of  work  to  pull  an  electron  out  of  a 
metal;  if  it  did  not,  the  electrons  would  continually  leak  out  into  empty 
space.  The  minimum  amount  of  work,  that  required  to  pull  out  the  most 
easily  detachable  electron,  is  by  definition  the  work  function  <£.  Then  if 
hv  was  greater  than  the  work  function,  the  electron  might  be  able  to 
escape  from  the  metal,  and  the  maximum  possible  kinetic  energy  which 
it  might  have  as  it  emerged  would  be 

=  hv  -  0.  (3.1) 


318  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIX 

If  the  electron  happened  not  to  be  the  most  easily  detachable  one,  it 
would  require  more  work  than  <£  to  pull  it  out,  and  it  would  have  less 
kinetic  energy  than  Eq.  (3.1)  when  it  emerged,  so  that  that  represents  the 
maximum  possible  kinetic  energy. 

Einstein's  hypothesis,  then,  led  to  two  definite  predictions.  In  the 
first  place,  there  should  bo  a  photoelectric  threshold:  frequencies  less  than 
a  certain  limit,  equal  to  <l>/h,  should  be  incapable  of  ejecting  photoelec- 
trons  from  a  metal.  This  prediction  proved  to  be  verified  experimentally, 
and  with  more  and  more  accurate  determinations  of  work  function  it 
continues  to  hold  true.  It  is  interesting  to  see  where  this  threshold  comes 
in  the  spectrum.  For  this  purpose,  it  is  more  convenient  to  find  the 
wave  length  X  =  c/v  corresponding  to  the  frequency  <t>/h.  If  we  express 
<£  in  electron  volls,  as  is  commonly  done,  (see  Eq.  (1.1),  Chap.  IX),  we 
have  the  relation 


__  he  X  300_       _   _  _ 

4~80~X  l(F10<  ~  "  '  (      ' 


All  wave  lengths  shorter  than  the  threshold  of  Eq.  (3.2)  can  eject  photo- 
electrons.  Thus  a  metal  with  a  small  work  function  of  two  volts  (which 
certain  alkali  metals  have)  has  a  threshold  in  the  red  and  will  react 
photoelectrically  to  visible  light,  while  a  metal  with  a  work  function  of 
six  volts  would  have  a  threshold  about  2000  A,  and  would  be  sensitive 
only  in  the  rather  far  ultraviolet.  Most  real  metals  lie  between  these 
limits. 

The  other  prediction  of  Einstein's  hypothesis  was  as  to  the  maximum 
velocity  of  the  photoeleotrons,  given  by  Eq.  (3.1).  This  is  also  verified 
accurately  by  experiment.  There  is  a  remarkable  feature  connected  with 
this:  the  energy  of  the  electrons  depends  on  the  frequency,  but  not  on  the 
intensity,  of  the  light  ejecting  them.  Double  the  intensity,  and  the 
number  of  photoelectrons  is  doubled,  but  not  the  energy  of  each  indi- 
vidual. This  can  be  carried  to  limits  which  at  first  sight  seem  almost 
absurd,  as  the  intensity  of  light  is  reduced.  Thus  let  the  intensity  be  so 
low  that  it  will  require  some  time,  say  half  a  minute,  for  the  total  energy 
falling  on  a  piece  of  metal  to  equal  the  amount  hv.  The  light  is  obviously 
distributed  all  over  the  piece  of  metal,  and  we  should  suppose  that  its 
energy  would  be  continuously  absorbed  all  over  the  surface.  Yet  that 
is  not  at  all  what  happens.  About  once  every  half  minute,  a  single 
electron  will  be  thrown  off  from  one  particular  spot  of  the  metal,  with  an 
energy  which  in  order  of  magnitude  is  equal  to  all  that  has  fallen  on  the 
whole  plate  for  the  last  half  minute.  It  is  quite  impossible,  on  any 
continuous  theory  like  the  wave  theory,  to  understand  how  all  this 
energy  could  have  become  concentrated  in  a  single  electron.  Yet  it 
is;  photoelectric  cells  can  actually  be  operated  as  we  have  just  described. 


SBC.  3]  RADIATION  AND  MATTER  319 

An  example  like  this  is  the  most  direct  sort  of  experimental  evidence  for 
Einstein's  hypothesis,  that  the  energy  in  light  waves,  at  least  when  it  is 
being  emitted  or  absorbed,  acts  as  if  it  were  concentrated  in  photons. 

For  a  long  time  it  was  felt  that  there  was  an  antagonism  between 
wave  theory  and  photons.  Certainly  the  photoelectric  effect  and  similar 
things  are  most  easily  explained  by  the  theory  of  photons.  On  the  other 
hand,  interference,  diffraction,  and  the  whole  of  physical  optics  cannot  be 
explained  on  any  basis  but  the  wave  theory.  How  could  these  theories 
be  simultaneously  true?  We  can  sec  what  happens  experimentally,  in  a 
case  where  we  must  think  about  both  types  of  theories,  by  asking  what 
would  happen  if  the  very  weak  beam  of  light,  falling  on  the  metal  plate 
of  the  last  paragraph,  had  previously  gone  through  a  narrow  slit,  so  that 
there  was  actually  a  diffraction  pattern  of  light  and  dark  fringes  on  the 
plate,  a  pattern  which  can  be  explained  only  by  the  wave1  theory.  We  say 
that  there  is  a  diffraction  pattern;  this  does  not  seem  to  mean  anything 
with  the  very  faint  light,  for  there  is  no  way  to  observe  it.  We  mean 
only  that  if  nothing  is  changed  but  the  intensity  of  the  light,  and  if  that  is 
raised  far  enough  so  that  the  beam  can  be  observed  by  the  eye,  a  diffrac- 
tion pattern  would  be-  seen  on  the  plate.  But  now  even  with  the  weak 
light,  it  really  has  meaning  to  speak  of  the  diffraction  pattern.  Suppose 
we  marked  off  the  light  and  dark  fringes  using  an  intense  light,  and  then 
returned  to  our  weak  light  and  made  a  statistical  study  of  the  points  on 
the  plate  from  which  electrons  were  ejected.  We  should  find  that  the 
electrons  were  all  emitted  from  what  ought  to  be  the  bright  fringes,  none 
from  the  dark  fringes.  The  wave  theory  tells  us  where  photons  will  be 
absorbed,  on  the  average.  This  can  be  seen  even  more  easily  if  we  replace 
the  photoelectric  plate  by  a  photographic  plate.  This  behaves  in  a  very 
similar  way:  in  weak  light,  occasionally  a  process  takes  place  at  one  single 
spot  of  the  plate,  producing  a  blackened  grain  when  the  plate  is  developed, 
and  the  effect  of  increasing  the  intensity  is  simply  to  increase  the  number 
of  developed  grains,  not  to  change  the  blackening  of  an  individual  grain. 
Then  a  weak  diffraction  pattern,  falling  on  a  plate  for  a  long  time,  will 
result  in  many  blackened  grains  in  the  bright  fringes,  none  in  the  dark 
ones,  so  that  the  final  picture  will  be  almost  exactly  the  same  as  if  there 
had  been  a  stronger  light  beam  acting  for  a  shorter  length  of  time. 

Nature,  in  other  words,  does  not  seem  to  be  worried  about  which  is 
correct,  the  wave  theory  or  the  photon  theory  of  light:  it  uses  both,  and 
both  at  the  same  time,  as  we  have  just  seen.  This  is  now  being  accepted 
as  a  fact,  and  the  theories  are  used  more  or  less  in  the  following  way.  In 
any  problem  where  light  is  concerned,  an  electromagnetic  field,  or  light 
wave,  is  set  up,  according  to  classical  types  of  theories.  But  this  field  is 
not  supposed  to  carry  energy,  as  a  classical  field  does.  Instead,  its 
intensity  at  any  point  is  supposed  to  determine  the  probability  that  a 


320  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XIX 

photon  will  be  found  at  that  point.  It  is  assumed  that  there  is  no  way 
at  all  of  predicting  exactly  where  any  particular  photon  will  go;  we  cannot 
say  in  any  way  whatever,  in  weak  light,  which  electron  of  the  metal  will 
be  ejected  next.  But  on  the  average,  the  wave  theory  allows  us  to  pre- 
dict. This  type  of  statistical  theory  is  quite  different  from  any  that  has 
been  used  in  physics  before.  Whenever  a  statistical  element  has  been 
introduced,  as  in  classical  statistical  mechanics,  it  has  been  simply  to 
avoid  the  trouble  of  going  into  complete  detail  about  a  very  complicated 
situation.  But  in  quantum  theory,  as  we  have  already  mentioned  in 
Chap.  Ill,  Sec.  3,  we  consider  that  it  is  impossible  in  principle  to  go  into 
complete  detail,  and  that  the  statistical  theory  is  all  that  there  is.  When 
one  meets  wave  mechanics,  one  finds  that  the  laws  governing  the  motion 
of  ordinary  particles,  electrons,  and  atoms,  are  also  wavelike  laws,  and 
thai  the  intensity  of  the  wave  gives  the  probability  of  finding  the  particle 
in  a  particular  spot,  but  that  no  law  whatever  seems  to  predict  exactly 
where  it  is  going.  This  is  a  curious  state  of  affairs,  according  to  our  usual 
notions,  but  nature  seems  to  be  made  that  way,  and  the  theoiy  of  radia- 
tion has  been  the  first  place  to  find  it  out. 


CHAPTER  XX 
IONIZATION  AND  EXCITATION  OF  ATOMS 

In  the  second  part  of  this  book,  we  have  been  concerned  with  the 
behavior  of  gases,  liquids,  and  solids,  and  we  have  seen  that  this  behavior 
is  determined  largely  by  the  nature  of  the  interatomic  and  intermolecular 
forces.  These  forces  arise  from  the  electrical  structure  of  the  atoms  and 
molecules,  and  in  this  third  part  we  shall  consider  that  structure  in  a  very 
elementary  way,  giving  particular  attention  to  the  atomic  and  molecular 
binding  in  various  types  of  substances.  Most  of  the  information  which 
we  have  about  atoms  comes  from  spectroscopy,  the  interaction  of  atoms 
and  light,  and  we  must  begin  with  a  discussion  of  the  excited  and  ioni- 
zated  states  of  atoms  and  molecules,  and  the  relation  between  energy 
levels  and  the  electrical  properties  of  the  atoms. 

1.  Bohr's  Frequency  Condition. — We  have  seen  in  Chap.  Ill,  Sec.  3, 
that  according  to  the  quantum  theory  an  atom  or  molecule  can  exist  only 
in  certain  definite  stationary  states  with  definite  energy  levels.  The 
sparing  of  these  energy  levels  depends  on  the  type  of  motion  we  are  con- 
sidering. For  molecular  vibrations  they  lie  so  far  apart  that  their 
energy  differences  are  largo  compared  to  kT  at  ordinary  temperatures,  as 
we  saw  in  Chap.  IX.  For  the  rotation  of  molecules  the  levels  are  closer 
together,  so  that  only  at  very  low  temperatures  was  it  incorrect  to  treat 
the  energy  as  being  continuously  variable.  For  molecular  translation,  as 
in  a  gas,  we  saw  in  Chap.  IV,  Sec.  1,  that  the  levels  came  so  close  together 
that  in  all  cases  we  could  treat  them  as  being  continuous.  Atoms  and 
molecules  can  also  have  energy  levels  in  which  their  electrons  are  excited 
to  higher  energies  than  those  found  at  low  temperatures.  Ordinarily 
the  energy  difference  between  the  lowest  electronic  state  (called  the 
normal  state,  or  the  ground  state)  and  the  states  with  electronic  excitation 
is  much  greater  than  the  energy  differences  concerned  in  molecular 
vibration.  These  differences,  in  fact,  are  so  large  that  at  ordinary  tem- 
peratures no  atoms  at  all  are  found  in  excited  electronic  levels,  so  that  we 
do  not  have  to  consider  them  in  thermal  problems.  The  excited  levels 
have  an  important  bearing,  however,  on  the  problem  of  interatomic 
forces,  and  for  that  reason  we  must  take  them  up  here.  Finally,  an 
electron  can  be  given  so  much  energy  that  it  is  entirely  removed  from 
its  parent  atom,  and  the  atom  is  ionized.  Then  the  electron  can  wander 
freely  through  the  volume  containing  the  atom,  like  an  atom  of  a  perfect 

321 


322 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


[CHAP.  XX 


gas,  and  its  energy  levels  are  so  closely  spaced  as  to  be  continuous.  The 
energy  levels  of  an  atom  or  molecule,  in  other  words,  include  a  set  of  dis- 
crete levels,  and  above  those  a  continuum  of  levels  associated  with 
ionization.  Such  a  set  of  energy  levels  for  an  atom  is  shown  schematically 
in  Fig.  XX-1.  The  energy  difference  between  the  normal  state  and  the 
beginning  of  the  continuum  gives  the  work  required  to  ionize  the  atom  or 
molecule,  or  the  ionization  potential.  The  ionization  potentials  are 
ordinarily  of  the  order  of  magnitude  of  a  number  of  volts.  The  lowest 
excited  energy  level  of  an  atom,  called  in  some  cases  the  resonance  level 
(sometimes  the  word  resonance  level  is  used  for  an  excited  level  somewhat 

higher  than  the  lowest  one,  but  one  which  is 
reached  particularly  easily  from  the  lowest 
one  by  the  absorption  of  radiation),  is 
ordinarily  several  volts  above  the  normal 
state,  the  energy  difference  being  called  the 
resonance  potential.  This  verifies  our  state- 
ment that  electrons  will  not  be  appreciably 
excited  at  ordinary  temperatures.  We  can 
see  this  by  finding  the  characteristic  tempera- 
luro  associated  with  an  energy  difference  of 
the  order  of  magnitude  of  a  volt.  If  we  let 
fcO  =  energy  of  one  electron  volt,  we  have 
0  =  (4.80  X  10-10)/(300  X  1.379  X  10~1G)  = 
11,600°  abs.  Thus  ordinary  temperatures 
tire  very  small  compared  with  such  a  charac- 
teristic temperature.  If  we  consider  the 
possibility  of  an  electronic  specific  heat  com- 
ing from  the  excitation  of  electrons  to  excited 
levels,  we  see  that  such  a  specific  heat  will 
be  quite  negligible,  for  ordinary  substances, 
at  temperatures  below  several  thousand 
degrees.  A  few  exceptional  elements,  how- 
ever, such  as  some  of  the  transition  group  metals,  have  excited  energy 
levels  only  a  few  hundredths  of  a  volt  above  the  normal  state,  and  these 
elements  have  appreciable  electronic  specific  heat  at  ordinary 
temperatures. 

There  are  two  principal  mechanisms  by  which  atoms  or  molecules  can 
jump  from  one  energy  level  to  another.  These  are  the  emission  and 
absorption  of  radiation,  and  collisions.  For  the  moment  we  shall  con- 
sider the  first  process.  An  atom  in  a  given  energy  level  can  have  transi- 
tions to  any  higher  energy  level  with  absorption  of  energy,  or  to  any 
lower  level  with  emission,  each  transition  meaning  a  quite  definite  energy 
difference.  By  the  conservation  of  energy,  the  same  quite  definite 


Ionization 
potent/a/ 


Resonance 


potc 


ihal 


FIQ.  XX-1. — Schematic     set 
energy  levels  for  an  atom. 


of 


SEC.  1]  ION1ZATION  AND  EXCITATION  OF  ATOMS  323 

energy  must  be  converted  into  a  photon  if  light  is  being  emitted,  or  must 
have  come  from  a  photon  if  it  is  being  absorbed.  But  by  Einstein's 
hypothesis,  the  frequency  of  a  photon  is  determined  from  its  energy,  by 
the  relation  energy  =  hv.  Thus  a  transition  between  two  atomic  energy 
levels,  with  energies  E\  and  #2,  must  result  in  the  emission  or  absorption 
of  a  photon  of  frequency  v,  where 

#2  -  El  =  hv.  (1.1) 

With  sharp  and  discrete  energy  levels,  then,  we  must  haw  definite  fre- 
quencies emitted  and  absorbed,  or  must  have  a  sharp  lino  spectrum.  Tho 
relation  (1.1),  as  applied  to  the  spectrum,  is  duo  to  Bohr,  and  is  often 
called  Bohr's  frequency  condition;  it  is  really  tho  foundation  of  the 
theory  of  spec! roscopy.  Regarded  as  an  empirical  fact,  it  states  that  the 
frequencies  observed  in  any  spectrum  can  be  written  as  the  differences 
of  a  sot  of  numbers,  called  terms,  which  are  simply  the  energy  levels  of  tho 
system,  divided  by  h.  Since  with  a  given  table  of  terms  wo  can  find  a 
groat  many  more  differences  than  there  are  terms,  this  means  that  a  given 
complicated  spectrum  can  bo  greatly  simplified  if,  instead  of  tabulating 
all  the  spectral  frequencies,  wo  tabulate  only  the  much  smaller  number  of 
term  values.  And  the  importance  of  Bohr's  frequency  condition  goes 
much  further  than  this.  For  by  observing  tho  frequencies  in  the  spec- 
trum and  finding  the  terms,  we  can  get  the  energy  levels  of  tho  atom  or 
molecule  emitting  the  spectrum.  Wo  can  use  these  directly,  with  no 
more  theory,  in  such  things  as  a  calculation  of  tho  specific  heat.  For 
instance,  the  energy  levels  of  molecular  vibration  and  rotation,  used  in 
finding  tho  specific  heat  of  molecules  in  Chap.  IX,  arc  the  results  of  spoc- 
troscopic  observation.  Furthermore,  wo  can  use  tho  observed  energy 
levels  to  verify,  in  a  very  precise  way,  any  theoretical  calculation  which 
we  have  made  on  the  basis  of  the  quantum  conditions.  The  relation 
between  the  sharp  linos  observed  in  spectra,  and  the  energy  levels  of  the 
atoms  or  molecules  making  the  spoctra,  has  boon  the  most  important  fact 
in  the  development  of  our  knowledge  of  the  structure  of  atoms  and 
molecules. 

Bohr's  frequency  condition  has  one  surprising  feature.  The  fre- 
quency of  emitted  light  is  related,  according  to  it,  to  tho  energy  rather 
than  the  frequency  of  the  motion  in  the  atom  that  produces  it.  This  is 
entirely  contrary  to  classical  theory.  A  vibrating  charge,  oscillating 
with  a  given  frequency,  in  classical  electromagnetic  theory,  sends  out  light 
of  the  frequency  with  which  it  vibrates.  According  to  wave  mechanics, 
however,  there  is  really  not  a  contradiction  here.  For  in  wave  mechanics, 
the  particles,  ordinarily  the  electrons,  which  produce  the  light  do  not  move 
according  to  classical  theory,  but  the  frequencies  actually  present  in  their 
average  motions  are  those  given  by  Bohr's  frequency  condition.  Thus 


324  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XX 

the  relation  between  the  motion  of  the  particles,  and  the  light  they  send 
out,  is  more  nearly  in  accord  with  classical  electromagnetic  theory  than 
we  should  suppose  at  first  sight. 

2.  The  Kinetics  of  Absorption  and  Emission  of  Radiation. — With 
Bohr's  picture  of  the  relation  between  energy  levels  and  discrete  spectral 
lines  in  mind,  Einstein  gave  a  kinetic  derivation  of  the  law  of  black-body 
radiation,  which  is  very  instructive  and  which  has  had  a  great  deal  of 
influence.  Einstein  considered  two  particular  stationary  states  of  an 
atom,  say  the  fth  and  jth  (whore  for  definiteness  wo  assume  that  the  ith 
lies  above  tho  ji\i),  and  the  radiation  which  could  be  omitted  and  absorbed 
in  going  between  these  two  states,  radiation  of  frequency  v»-/,  where 

hvt]  =  Ei  -  E3.  (2.1) 

Suppose  the  atom,  or  group  of  atoms  of  the  same  sort,  capable  of  existing 
in  these  stationary  states,  is  in  thermal  equilibrium  with  radiation  at 
temperature  T.  Then  for  equilibrium,  using  the  principle  of  detailed 
balancing,  the  amount  of  energy  of  frequency  vlt  absorbed  by  the  atoms 
per  second  in  making  the  transition  from  state  j  to  state  i  must  equal  the 
amount  of  the  same  frequency  emitted  per  second  in  going  from  state  i 
to  state  j.  Einstein  made  definite  assumptions  as  to  the  probability  of 
making  those  two  transitions.  In  the  first  place,  consider  absorption. 
The  number  of  atoms  absorbing  photons  per  second  must  surely  be  pro- 
portional first  to  the  number  of  atoms  in  the  lower,  ^th  energy  level,  which 
we  shall  call  N,',  and  to  the  intensity  of  radiation  of  the  frequency  vlj, 
which  is  uvij.  Thus  Einstein  assumed  that  the  number  of  atoms  absorb- 
ing photons  per  second  was 

NtB^u,,,,  (2.2) 

where  /?»•,-  is  a  constant  characteristic  of  the  transition.  Next  consider 
emission.  Quite  clearly  an  atom  in  an  excited  state  can  emit  radiation 
and  jump  to  a  lower  stato  without  any  outside  action  at  all.  This  is 
called  spontaneous  emission,  and  the  probability  of  it  was  assumed  by 
Einstein  to  be  a  constant  independent,  of  the  intensity  of  radiation.  Thus 
he  assumed  the  number  of  atoms  falling  spontaneously  from  the  ith  to 
the  jth  levels  per  second  with  emission  of  radiation  was 

NiAi,-,  (2.3) 

/ 

where  AH  is  another  constant.  But  at  the  same  time  there  must  be 
another  process  of  emission,  as  Einstein  showed  by  considering  very  high 
temperatures,  where  u,a  is  very  large.  In  this  limit,  the  term  (2.2)  is 
bound  to  be  very  large  compared  to  the  term  (2.3),  so  that  with  just  these 
two  terms  equilibrium  is  impossible.  Guided  by  certain  arguments 
based  on  classical  theory,  Einstein  assumed  that  this  additional  probabil- 


SEC.  21  10NIZAT10N  AND  EXCITATION  OA'  ATOMS  325 

ity  of  emission,  generally  called  induced  emission,  was 

,,*  (2.4) 


proportional  as  the  absorption  was  to  the  intensity  of  external  radiation. 
For  equilibrium,  then,  we  must  have  equal  numbers  of  photons  emitted 
and  absorbed  per  second.     Thus  we  must  have 

N*(A>,  +  J3tfUM-,)  =  N^u^  (2.5) 

But  at  the  same  time,  if  there  is  equilibrium,  we  know  that  the  number 
of  atoms  in  the  ith  and  jth  states  must  be  determined  by  the  Boltzmann 
factor.  Thus  we  must  have 

_Ei 

Ni  =  const,  e    kT, 

_Ej 

NJ  =  const,  e   kr, 

N  _<!?,-.»,)  _h^ 

jf  =  c       kT       =  e    kT-  (2.6) 

We  can  now  solve  Eqs.  (2.5)  and  (2.6)  to  find  u^.     We  have 


A  „        1  ,~ 

(    i 


The  energy  density  (2.7)  would  bo  the  Planck  distribution  law,  if  we  had 

Atl  _  Sirhv?,  fl>  ft. 

B^  ~  ~*    '  (2'8) 

as  we  see  by  comparison  with  Eq.  (2.6),  Chap.  XIX.  Einstein  assumed 
that  Eq.  (2.8)  was  true,  and  in  that  way  had  a  partial  derivation  of  the 
Planck  law. 

Einstein's  derivation  of  the  black-body  radiation  law  is  particularly 
important,  for  it  gives  us  an  insight  into  the  kinetics  of  radiation  processes. 
Being  a  kinetic  method,  it  can  be  used  even  when  we  do  not  have  thermal 
equilibrium.  Thus  if  we  know  that  radiation  of  a  certain  intensity  is 
falling  on  atoms,  we  can  find  how  many  will  be  raised  to  the  excited  state 
per  second,  in  terms  of  the  coefficient  By.  But  this  means  that  we  can 
find  the  absorptivity  of  matter  made  of  these  atoms,  at  this  particular 
wave  length.  Conversely,  from  measurements  of  absorptivity,  we 
can  deduce  experimental  values  of  B^.  And  from  Eq.  (2.8)  we  can  find 
the  rate  of  emission,  or  the  emissive  power,  if  we  know  the  absorptiv- 


326  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XX 

ity.  Equation  (2.8),  saying  that  these  two  quantities  are  proportional 
to  each  other,  is  really  very  closely  related  to  Kirchhoff's  law,  discussed  in 
Chap.  XIX,  Sec.  1,  and  Einstein's  whole  method  is  closely  related  to  the 
arguments  of  Kirchhoff. 

We  can  put  Einstein's  assumption  of  spontaneous  and  induced  emis- 
sion in  an  interesting  light  if  we  express  Eq.  (2.5),  not  in  terms  of  the 
energy  density  of  radiation,  but  in  terms  of  the  average  number  of 
photons  Nv  in  the  standing  wave  of  frequency  v.  Let  us  see  just  what  we 
mean  by  this.  We  are  assuming  that  the  energy  of  this  standing  wave 
is  quantized,  equal  to  nhv,  as  in  Chap.  XIX,  Sec.  2;  and  by  Einstein's 
hypothesis  we  are  assuming  that  this  means  that  there  are  really  n 
(or  Nv)  photons  associated  with  this  wave.  In  terms  of  this,  we  see  that 
the  cnorgy  density  uv  is  dotorminod  by  the  relation 

Total  energy  in  dv  =  uvV  dv 

=  number  of  waves  in  dv  times  number  of  photons  in 
each  wave  times  energy  in  each  photon 

__ \7  A7    Ij ..  /O  Q\ 

-     -jjj-     vJMrnv,  v.j) 

using  the  result  of  Eq.  (2.3),  Chapter  XIX.     From  this,  we  have 

^  =  TT  (2-10) 

I/  i  V  \f 

Then  we  can  rewrite  Eq.  (2.5),  using  Eq.  (2.8),  as 

Number  of  photons  emitted  per  second 

-  Al3Nt(Nv  +  1) 

=  Number  of  photons  absorbed  per  second 

=  AVN,N,.  (2.11) 

The  interesting  feature  of  Eq.  (2.11)  is  that  the  induced  and  spontaneous 
emission  combine  into  a  factor  as  simple  as  Nv  +  1.  This  is  strongly 
suggestive  of  the  factors  Nj  +  1,  which  we  met  in  the  probability  of 
transition  in  the.  Einstein-Bose  statistics,  Eq.  (4.2),  of  Chap.  VI.  As  a 
matter  of  fact,  the  Einstein-Bose  statistics,  in  a  slightly  modified  form, 
applies  to  photons.  Since  it  does  not  really  contribute  further  to  our 
understanding  of  radiation,  however,  we  shall  not  carry  through  a  discus- 
sion of  this  relation,  but  merely  mention  its  existence. 

3.  The  Kinetics  of  Collision  and  lonization. — In  the  last  section  we 
have  been  considering  the  emission  and  absorption  of  radiation  as  a 
mechanism  for  the  transfer  of  atoms  or  molecules  from  one  energy  level 
to  another.  The  other  important  mechanism  of  transfer  is  that  of 
collisions  with  another  atom,  molecule,  or  more  often  with  an  electron. 
In  such  a  collision,  the  colliding  particles  can  change  their  energy  levels, 


SEC.  3]  IONIZATION  AND  EXCITATION  OF  ATOMS  327 

and  at  the  same  time  change  their  translational  kinetic  energy,  which  is 
not  considered  in  calculating  their  energy  levels,  by  such  an  amount  that 
the  total  energy  is  conserved.  A  collision  in  which  only  the  translational 
kinetic  energy  changes,  without  change  of  the  internal  energy  levels  of 
the  colliding  particles,  is  called  an  elastic  collision;  this  is  the  typo  of 
collision  considered  in  Chap.  VI,  where  we  were  finding  the  effect  of  colli- 
sions on  the  molecular  distribution  function.  The  type  of  collision  that 
results  in  a  change  of  energy  level,  however,  involving  either  excitation  or 
ionization,  is  called  an  inelastic  collision,  and  it  is  in  such  collisions  that 
we  are  particularly  interested  here.  Wo  consider  the  kinetics  of  such 
collisions  in  the  present  section,  coming  later  to  the  treatment  of  thermal 
equilibrium,  in  particular  the  equilibrium  between  ionization  and  recom- 
bination, as  treated  by  thermodynamics  and  kinetic  theory.  For 
generality,  we  begin  by  considering  the  general  nature  of  collisions 
between  atomic  or  electronic  particles. 

The  processes  which  we  consider  are  collisions,  and  most  of  them 
arc  collisions  of  two  particles,  which  separate  again  after  their  encounter. 
The  probabilities  of  such  collisions  are  described  in  terms  of  a  quantity 
called  a  collision  cross  section,  which  we  now  proceed  to  define.  First 
let  us  consider  a  simple  mechanical  collision.  Suppose  we  have  a  small 
target,  of  area  A  (small  compared  to  1  sq.  cm.).  Then  suppose;  we  fire 
many  projectiles  in  its  direction,  but  suppose  they  arc  not  well  aimed,  so 
that  they  are  equally  likely  to  strike  any  point  of  the  square  centimeter 
containing  the  target.  Then  we  ask,  what  is  the  chance  that  any  one 
of  the  projectiles  will  strike  the  target?  Plainly  this  chance  will  be  the 
ratio  of  the  area  of  the  target,  A,  to  the  area  of  the  whole  square  centi- 
meter, which  is  unity.  In  other  words,  A,  which  we  call  the  collision 
cross  section  in  this  particular  case,  is  the  fraction  of  all  projectiles  that 
hit  the  target.  If  instead  of  one  target  we  had  N  in  the  region  traversed 
by  the  beam  of  unit  cross  section,  and  if  even  the  N  targets  filled  only 
a  small  fraction  of  the  square  centimeter,  so  that  there  was  small  chance 
that  one  target  lay  behind  another,  then  the  chance  that  a  particular 
projectile  would  have  a  collision  would  be  NA,  and  to  get  the  collision 
cross  section  of  a  single  target  we  should  have  to  take  the  fraction  having 
collisions,  and  divide  by  N. 

In  a  similar  way,  in  the  atomic  or  molecular  case,  we  allow  a  beam  of 
colliding  particles  to  strike  the  atoms  or  molecules  that  we  wish  to  investi- 
gate. A  certain  number  of  the  particles  in  the  incident  beam  will  pass 
by  without  collision,  while  a  certain  number  will  collide  and  be  deflected. 
We  count  the  fraction  colliding,  divide  this  fraction  by  the  number  of 
particles  with  which  they  could  have  collided,  and  the  result  is  the  colli- 
sion cross  section.  This  can  plainly  be  made  the  basis  of  an  experi- 
mental method  of  measuring  collision  cross  sections.  We  start  a  beam 


328  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XX 

of  known  intensity  through  a  distribution  of  particles  with  which  they 
may  collide,  and  we  measure  the  intensity  of  the  beam  after  it  has  trav- 
ersed different  lengths  of  path.  We  observe  the  intensity  to  fall  off 
exponentially  with  the  distance,  and  from  that  can  deduce  the  cross 
section  in  the  following  manner. 

Lot  the  beam  have  unit  cross  section,  and  let  a:  be  a  coordinate 
measured  along  the  beam.  The  intensity  of  the  beam,  at  point  x,  is 
defined  as  the  number  of  particles  crossing  the  unit  cross  section  at  x 
per  second.  We  shall  call  it  I(x),  and  shall  find  how  it  varies  with  x. 
Consider  the  collisions  in  the  thin  sheet  between  x  and  x  +  dx.  Let  the 
number  of  particles  per  unit  volume  with  which  the  beam  is  colliding  be 
N/V.  Then  in  the  thin  sheet  between  x  and  x  +  dx,  with  a  volume  dx, 
there  will  be  N  dx/V  particles.  Let  each  of  these  have  collision  cross 
section  A.  Then  the  fraction  of  particles  colliding  in  the  sheet  will  by 
definition  be  NA  dx/V.  This  is,  however,  equal  to  the  fractional  decrease 
in  intensity  of  the  beam  in  this  distance.  That  is, 


Integrating,  this  gives 


dl       NA    .  ,01. 

-y  =  -y   dx.  (3.1) 


N  A 
In  7  =  ~x  +  const.,  (3.2) 


or,  if  the  intensity  is  /0  when  x  =  0,  we  have 


(3.3) 


From  Eq.  (3.3)  we  see  the  exponential  decrease  of  intensity  of  which  we 
have  just  spoken,  and  it  is  clear  that  by  measuring  the  rate  of  exponential 
decrease  we  can  find  the  collision  cross  section  experimentally. 

The  intensity  of  a  beam  falls  to  1/e  of  its  initial  value,  from  Eq.  (3.3), 
in  a  distance 

(3.4) 


(ft 


This  distance  is  often  called  the  mean  free  path.  As  we  can  see,  it  is 
inversely  proportional  to  the  number  of  particles  per  unit  volume,  or 
the  density,  and  inversely  proportional  to  the  collision  cross  section. 
The  mean  free  path  is  most  commonly  discussed  for  the  ordinary  elastic 
collisions  of  two  molecules  in  a  gas.  For  such  collisions,  the  collision 
cross  sections  come  out  of  the  order  of  magnitude  of  the  actual  cross 
sectional  areas  of  the  molecules;  that  is,  of  the  order  of  magnitude  of 


SEC.  3]  IONIZATION  AND  EXCITATION  OF  ATOMS  329 

10~~16  cm2.  In  a  gas  at  normal  pressure  and  temperature,  there  aro 
2.70  X  1019  molecules  per  unit  volume.  Thus,  the  mean  free  path  is 
of  the  order  of  1/(2.70  X  10')  =  3.7  X  10™4  cm.  As  a  matter  of  fact, 
most  values  of  A  are  several  times  this  value,  giving  moan  free  paths 
smaller  than  the  figure  above*.  As  the  pressure  is  reduced,  however, 
the  mean  free  paths  become  quite  long.  Thus  at  0°C.,  but  a  pressure  of 
10~5  atm.,  the  mean  free  paths  become  of  the  order  of  magnitude  of  1  cm.; 
with  pressures  several  thousand  times  smaller  than  this,  which  are 
easily  realized  in  a  high  vacuum,  the  mean  free  path  becomes  many 
meters.  In  other  words,  the  probability  of  collision  in  the  dimensions 
of  an  ordinary  vacuum  tube  becomes  negligible,  and  molecules  shoot  from 
one  side  to  the  other  without  hindrance. 

The  collision  cross  section  is  closely  related  to  the  quantities  Afa,  which 
we  introduced  in  discussing  collisions  in  Sec.  I,  Chap.  VI.  We  were 
speaking  there  about  a  particular  sort  of  collision,  one  in  which  one  of  the 
colliding  particles  before  collision  was  in  cell  i  of  the  phase  space,  the 
other  in  cell  j,  while  after  collision  the  first  was  in  cell  k,  the  second  in 
cell  L  The  number  of  such  collisions  per  unit  time  was  assumed  to  be 
AlkJiNlN].  In  the  present  case,  we  are  treating  all  collisions  of  two 
molecules,  one  moving,  the  other  at  rest,  irrespective  of  the  velocities 
after  collision.  That  is,  the  present  case  corresponds  to  the  case  where 
one  of  the  two  cells  i  or  j  corresponds  to  a  molecule  at  rest,  and  where 
we  sum  over  all  cells  k  and  L  Furthermore,  there  we  were  interested 
in  the  number  of  collisions  per  unit  time,  here  in  the  number  per  unit 
distance  of  path.  It  is  clear  that  if  we  knew  the  Aft's,  we  could  compute 
from  them  the  collision  cross  section  of  the  sort  we  are  now  using.  Our 
collision  cross  section  gives  less  specific  information,  however.  Wo 
expect  to  find  a  different  collision  cross  section  for  each  velocity  of 
impinging  particle,  thoiigh  our  restriction  that  the  particle  with  which 
it  is  colliding  is  at  rest  is  not  really  a  restriction  at  all,  for  it  is  an  easy 
problem  in  mechanics  to  find  what  would  happen  if  both  particles 
were  initially  in  motion,  if  we  know  the  more  restricted  case  where  one  is 
initially  at  rest.  But  the  Aft's  give  additional  information  about  the 
velocities  of  the  two  particles  after  collision.  They  assume  that  the 
total  kinetic  energy  after  collision  equals  the  total  kinetic  energy  before 
collision;  that  is,  they  assume  an  elastic  collision.  Then,  as  mentioned 
in  Sec.  1,  Chap.  VI,  there  are  two  quantities  which  can  be  assigned  at  will 
in  describing  the  collision,  which  may  be  taken  to  be  the  direction  of  one 
of  the  particles  after  collision.  To  give  equivalent  information  to  the  Aft 
in  the  language  of  collision  cross  sections,  we  should  give  not  merely  the 
probability  that  a  colliding  particle  of  given  velocity  strike  a  fixed  parti- 
cle, but  also  the  probability  that  after  collision  it  travel  off  in  a  definite 
direction.  This  leads  to  what  is  called  a  collision  cross  section  for 


330  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XX 

/ 

scattering  in  a  given  direction:  the  probability  that  the  colliding  particle 
have  a  collision,  and  after  collision  that  it  travel  in  a  direction  lying  within 
a  unit  solid  angle  around  a  particular  direction  in  space.  This  cross 
section  gives  as  much  information  as  the  set  of  A$s,  though  in  a  different 
form,  so  that  it  requires  a  rather  complicated  mathematical  analysis, 
which  we  shall  not  carry  out,  to  pass  from  one  to  the  other. 

We  are  now  ready  to  consider  the  collision  cross  sections1  for  some  of 
the  processes  concerned  in  excitation  and  ionization.  First  we  consider 
the  collision  of  an  electron  with  a  neutral  atom.  In  the  first  place, 
there  are  two  possible  types  of  collision,  elastic  and  inelastic.  If  the 
energy  of  the  incident  electron  is  less  than  the  resonance  potential  of  the 
atom,  then  an  inelastic  collision  is  not  possible,  for  the  final  kinetic  energy 
of  the  two  particles  cannot  be  less  than  zero.  Thus  below  the  resonance 
potential  all  collisions  arc  elastic.  The  cross  section  for  elastic  collision 
varies  with  the  velocity  of  the  impinging  electron,  sometimes  in  what 
seems  a  very  erratic  manner.  Generally  it  increases  as  the  velocity 
decreases  to  zero,  but  for  some  atoms,  particularly  the  inert  gas  atoms, 
it  goes  through  a  maximum  at  a  velocity  associated  with  an  energy  of  the 
order  of  10  electron  volts,  then  decreases  again  as  the  velocity  is  decreased, 
until  it  appears  to  become  zero  as  the  velocity  goes  to  zero.  This  effect, 
meaning  that  extremely  slow  electrons  have  extremely  long  mean  free 
paths  in  these  particular  gases,  is  called  the  Ramsauer  effect,  from  its 
discoverer.  The  collision  cross  sections  for  elastic  collision  of  electrons 
and  atoms  have  been  investigated  experimentally  for  all  the  convenient 
atoms,  and  many  molecules,  disclosing  a  wido  variety  of  behaviors. 
They  can  also  be  investigated  theoretically  by  the  wave  mechanics, 
involving  methods  which  cannot  be  explained  here,  and  the  theoretical 
predictions  agree  very  satisfactorily  with  the  experiments,  even  to  the 
extent  of  explaining  the  Ramsauer  effect. 

Above  the  resonance  potential,  an  electron  has  the  possibility  of 
colliding  inclastically  with  an  atom,  raising  it  to  an  excited  level,  as  well 
as  of  colliding  elastieally.  The  probability  of  excitation,  or  the  collision 
cross  section  for  inelastic  collision,  starts  up  as  the  voltage  is  raised  above 
the  excitation  potential,  rising  quite  rapidly  for  some  transitions,  more 
slowly  for  others,  then  reaches  a  maximum,  and  finally  begins  to  decrease 
if  the  electron  is  too  fast.  Of  course,  atoms  can  be  raised  not  merely  to 
their  resonance  level,  but  to  any  other  excited  level,  by  an  electron  of 
suitable  energy,  and  each  one  of  these  transitions  has  a  collision  cross 
section  of  the  type  we  have  mentioned,  starting  from  zero  just  at  the 
suitable  excitation  energy.  The  probabilities  of  excitation  to  high 
energy  levels  are  small,  however;  by  far  the  most  important  inelastic 

1  For  further  information  about  collisions,  see  Massey  and  Mott,  "The  Theory  of 
Atomic  Collisions,"  Oxford  University  Press,  1933. 


SBC.  3]  IONIZATION  AND  EXCITATION  OF  ATOMS  331 

types  of  collision,  at  energies  less  than  the  ionization  potential,  are  those 
in  which  the  atom  is  raised  to  one  of  its  lowest  excited  levels. 

As  soon  as  the  energy  of  the  impinging  electron  becomes  greater 
than  the  ionization  potential,  inelastic  collisions  with  ionization  become 
possible.  Here  again  the  collision  cross  section  starts  rather  rapidly 
from  zero  as  the  potential  is  raised  above  the  ionization  potential,  reaches 
a  maximum  at  the  order  of  magnitude  of  two  or  throe  times  the  ionization 
potential,  and  gradually  falls  off  with  increasing  energy.  The  reason 
for  the  falling  off  with  rising  energy  is  an  elementary  one:  a  fast  electron 
spends  less  time  in  an  atom,  and  consequently  has  less  time  to  ionize  it 
and  less  probability  of  producing  the  transition.  A  collision  with  ioniza- 
tion is  of  course  different  from  an  excitation,  in  that  the  ejected  electron 
also  leaves  the  scene  of  the  collision,  so  that  after  the  collision  we  have 
three  particles,  the  ion  and  two  electrons,  instead  of  two  as  in  the  previous 
case.  This  fact  is  used  in  the  experimental  determination  of  resonance 
and  ionization  potentials.  A  beam  of  electrons,  of  carefully  regulated 
voltage,  is  shot  through  a  gas,  and  as  the  voltage  is  adjusted,  it  is  observed 
that  the  mean  free  path  shows  sharp  breaks  as  a  function  of  voltage  at 
certain  points,  decreasing  sharply  as  certain  critical  voltages  are  passed. 
This  does  not  tell  us  whether  the  critical  voltages  are  resonance  or 
ionization  potentials,  but  if  the  presence  of  additional  electrons  is  also 
observed,  an  increase  in  these  additional  electrons  is  noticed  at  an 
ionizatiou  potential  but  not  at  a  resonance  potential. 

Of  course,  each  of  these  types  of  collision  must  have  an  inverse  type, 
and  the  principle  of  microscopic  reversibility,  discussed  in  Chap.  VI, 
shows  that  the  probability,  or  collision  cross  section,  for  the  inverse 
collision  can  be  determined  from  that  of  the  direct  collision.  The  oppo- 
site or  inverse  to  a  collision  with  excitation  is  what  is  called  a  collision 
of  the  second  kind  (the  ordinary  one  being  called  a  collision  of  the  first 
kind).  In  a  collision  of  the  second  kind  an  electron  of  low  energy  collides 
with  an  excited  atom  or  molecule,  the  atom  has  a  transition  to  its  normal 
state  or  some  lower  energy  level  than  the  one  it  is  in,  and  the  electron 
comes  off  with  more  kinetic  energy  than  it  had  originally.  The  inverse 
to  an  ionization  is  a  three-body  collision:  two  electrons  simultaneously 
strike  an  atom,  one  is  bound  to  the  atom,  which  falls  to  its  normal  state 
or  some  excited  state,  while  the  other  electron,  as  in  a  collision  of  tho 
second  kind,  is  ejected  with  more  kinetic  energy  than  the  two  electrons 
together  had  before  the  collision.  Such  a  process  is  called  a  recombina- 
tion; and  it  is  to  be  noticed  that  we  never  have  a  recombination  just  of 
an  atom  and  an  electron,  for  there  would  be  no  body  to  remove  the  extra 
energy. 

In  addition  to  the  types  of  collision  we  have  just  considered,  where 
an  electron  and  an  atom  or  molecule  collide,  one  can  have  collisions  of 


332  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XX 

two  atoms  or  molecules  with  each  other,  with  excitation  or  ionization. 
It  is  perfectly  possible  to  have  two  fast  atoms  collide  with  each  other, 
originally  in  their  normal  states,  and  result  in  the  excitation  or  ionization 
of  one  or  both  of  the  atoms.  The  collision  of  the  second  kind,  inverse  to 
this,  is  that  in  which  an  excited  atom  and  a  normal  one  collide,  the 
excited  one  falls  to  its  normal  state,  and  the  atoms  gain  kinetic  energy. 
Then  one  can  have  an  exchange  of  excitation:  an  excited  and  a  normal 
atom  collide,  and  after  the  collision  the  excited  one  has  fallen  to  its  normal 
state,  the.  normal  one  is  excited,  and  the  discrepancy  in  energy  is  made 
up  in  the  kinetic  energy  of  translation  of  the  atoms.  Or  one  can  have  an 
interchange  of  ionization:  a  neutral  atom  and  a  positive  ion  collide,  and 
after  collision  the  first  one  has  become  a  positive  ion,  the  second  one  is 
neutral.  We  shall  consider  this  same  process  from  the  point  of  view  of 
statistical  mechanics  in  the  next  section.  In  these  cases  of  collisions  of 
atoms,  it  is  very  difficult  to  calculate  the  probabilities  of  the  various 
processes,  or  the  collision  cross  sections,  and  in  most  cases  few  measure- 
ments have  been  made.  In  general,  however,  it  can  be  said  that  the 
probability  of  elastic;  collision,  with  the  collision  of  two  atoms,  is  much 
greater  than  the  probability  of  any  of  the  various  types  of  inelastic 
collision. 

In  Chap.  X,  we  have  taken  up  the  kinetics  of  chemical  processes,  the 
types  of  collisions  between  molecules  which  result  in  chemical  reactions. 
Thorn  is  no  very  fundamental  distinction  between  those  collisions  and 
the  type  we  have  just  considered.  In  ordinary  chemical  reactions,  the 
colliding  molecules  are  under  all  circumstances  in  their  lowest  electronic 
state;  they  arc  not  excited  or  ionized.  The  reason  is  that  excitation  or 
ionization  potentials,  of  molecules  as  of  atoms,  arc  ordinarily  high  enough 
so  that  the  chance  of  excitation  or  ionization  is  negligible  at  the  tempera- 
tures at  which  reactions  ordinarily  take  place,  or  what  amounts  to  the 
same  thing,  the  colliding  molecules  taking  part  in  the  reaction  almost 
never  have  enough  energy  to  excite  or  ionize  each  other.  This  does  not 
mean,  however,  that  excitation  and  ionization  do  not  sometimes  occur, 
particularly  in  reactions  at  high  temperature;  undoubtedly  in  some  cases 
they  do.  It  is  to  be  noted  that  in  the  case  of  colliding  molecules,  unlike 
colliding  atoms,  inelastic  collisions  are  possible  without  electronic  excita- 
tion: the  molecules  ean  lose  some  of  their  translational  kinetic  energy 
in  the  form  of  rotational  or  vibrational  energy.  In  this  sense,  an  ordinary 
chemical  reaction,  as  explained  in  Sec.  3,  Chap.  X,  is  an  extreme  case  of 
an  inelastic  collision  without  excitation.  But  such  inelastic  collisions 
with  excitation  of  rotation  and  vibration  are  the  mechanism  by  which 
oquipartition  is  maintained  between  translational  and  rotational  and 
vibrational  energy,  in  changes  of  temperature  of  a  gas.  Sometimes  they 
do  not  provide  a  mechanism  efficient  enough  to  result  in  equilibrium 


SEC.  4]  IONIZATION  AND  EXCITATION  OF  ATOMS  333 

between  these  modes  of  motion.  For  example,  in  a  sound  wave,  there 
are  rapid  alternations  of  pressure,  produced  by  the  trauslational  motion 
of  the  gas,  and  these  result  in  rapid  alternations  of  the  translational 
kinetic  energy.  If  equilibrium  between  translation  and  rotation  can 
take  place  fast  enough,  there  will  bo  an  alternating  temperature,  related 
to  the  pressure  by  the  adiabatic  relation,  and  at  each  instant  there  will 
be  thermal  equilibrium.  Actually,  this  holds  for  low  frequencies  of 
sound,  but  there  is  evidence  that  at  very  high  frequencies  the  inelastic 
collisions  are  too  slow  to  produce  equilibrium,  and  the  rotation  does  not 
partake  of  the  fluctuations  in  energy. 

Another  interesting  example  is  found  in  some  cases  in  gas  discharges 
in  molecular  gases.  In  an  arc,  there  are  ordinarily  electrons  of  several 
volts'  energy,  since  an  electron  must  be*  accelerated  up  to  the  lowest- 
resonance  potential  of  the  gases  present  before  it  can  have  an  inelastic*, 
collision  and  reduce  its  energy  again  to  a  low  value.  These  electrons 
have  a  kinetic  energy,  then,  which  gas  moleeules  \\ould  acquire  only  at 
temperatures  of  a  good  many  thousand  degrees.  The  electrons  collide* 
elastically  with  atoms,  and  in  these  collisions  the  electrons  tend  to  lose 
energy,  the  atoms  to  gain  it,  for  this  is  just  the  mechanism  by  which 
thermal  equilibrium  and  cquipartition  tend  to  be  brought  about.  If  there 
are  enough  elastic  collisions  before  the  electrons  are  slowed  down  by  an 
inelastic  collision,  the  atoms  or  moleeules  will  tend  to  get  into  thermal 
equilibrium,  as  far  as  their  translation  is  concerned,  corresponding  to 
an  extremely  high  temperature.  That  such  an  equilibrium  is  actually 
set  up  is  observed  by  noticing  that  the  fast;  electrons  in  an  arc  have  a 
distribution  of  velocities  approximating  a  Maxwellian  distribution. 
But  apparently  the  inelastic  collisions  between  molecules,  or  between 
electrons  and  molecules,  are  not  effective  enough  to  give  the  molecules  the 
amount  of  rotational  energy  suitable  to  cquipartition,  in  the  short  length 
of  time  in  which  a  molecule  is  in  the  arc,  before  it  diffuses  to  the  wall  or 
otherwise  can  cool  off.  For  the,  rotational  energy  can  bo  observed  by 
band  spectrum  observation,  and  in  some  cases  it  is  found  that  it  corre- 
sponds to  rather  cool  gas,  though  the  translational  energy  corresponds 
to  a  very  high  temperature. 

4.  The  Equilibrium  of  Atoms  and  Electrons. — From  the  cases  we 
have  taken  up,  we  see  that  the  kinetics  of  collisions  forms  a  complicated 
and  involved  subject,  just  as  the  kinetics  of  chemical  reactions  does. 
Since  this  is  so,  it  is  fortunate  that  in  cases  of  thermal  equilibrium,  we 
can  get  results  by  thermodynamics  which  are  independent  of  the  precise 
mechanism,  and  depend  only  on  ionization  potentials  and  similarly  easily 
measured  quantities.  And  as  we  have  stated,  thermodynamics,  in  the 
form  of  the  principle  of  microscopic  reversibility,  allows  us  to  get  some 
information  about  the  relation  between  the  probability  of  a  direct  process 


334  INTRODUCTION  TO  CHEMICAL  PHYSICS  [CHAP.  XX 

and  of  its  inverse,  though  we  have  not  tried  to  make  any  such  calculations. 
To  see  how  it  is  possible,  we  need  only  notice  that  every  equilibrium  con- 
stant can  be  written  as  the  ratio  of  the  rates  of  two  inverse  reactions,  as 
we  saw  from  our  kinetic  derivation  of  the  mass  action  law  in  Chap.  X, 
so  that  if  we  know  the  equilibrium  constant,  from  energy  considerations, 
and  if  we  have  experimental  or  theoretical  information  about  the  rate 
of  one  of  the  reactions  concerned,  we  can  calculate  the  rate  of  the  inverse 
without  further  hypothesis. 

A  mixture  of  electrons,  ions,  and  atoms  forms  a  system  similar  to  that 
which  we  considered  in  Chap.  X,  dealing  with  chemical  equilibrium  in 
gases.  Equilibrium  is  determined,  as  it  was  there,  by  the  mass  action 
law.  This  law  can  bo  derived  by  balancing  the  rates  of  direct  and  inverse 
collisions,  but  it  can  also  be  derived  from  thermodynamics,  and  the 
equilibrium  constant  can  be  found  from  the  heat  of  reaction  and  the 
chemical  constants  of  the  various  particles  concerned.  The  heats  of 
reaction  can  be  found  from  the  various  ionization  potentials,  quantities 
susceptible  of  independent  measurement,  and  the  chemical  constants 
are  determined  essentially  as  in  Chap.  VIII.  Thus  there  are  no  new 
principles  involved  in  studying  the  equilibrium  of  atoms,  electrons,  and 
ions,  and  we  shall  merely  givo  a  qualitative  discussion  in  this  section, 
the  statements  being  equivalent  to  mathematical  results  which  can  bo 
established  immediately  from  the  methods  of  Chap.  X. 

The  simplest  type  of  problem  is  the  dissociation  of  an  atom  into  a 
positive  ion  and  an  electron.  By  the  methods  of  Chap.  X,  we  find  for 
the  partial  pressures  of  positive  ions,  negative  electrons,  and  neutral 
atoms  the  relation 


where  P(+),  -P(-),  P(n)  are  the  pressures  of  positive  ions,  electrons, 
and  neutral  atoms  respectively.  From  Eq.  (2.6),  Chap.  X,  we  can  find 
the  equilibrium  constant  KP  explicitly.  For  the  reaction  in  which  one 
mole  of  neutral  atoms  disappears,  one  mole  of  positive  ions  and  electrons 

appears,  we  have  v+  =  1,  i/_  =  1,  vn  =  —  1.     Then  the  quantity  —  ^VjUj 

3 

becomes  —  U+  —  f/_  +  Un  =  —I.  P.,  where  I.  P.  stands  for  the  ionization 
potential,  expressed  in  kilogram-calories  per  mole,  or  other  thermal  unit 
in  which  we  also  express  KT.  Then  we  have 

«  -— 
KP(T)  =  e?++*--**r*e   RT  .  (4.2) 

From  Eq.  (4.2),  we  see  that  the  equilibrium  constant  is  zero  at  the 
absolute  zero,  rising  very  slowly  until  the  temperature  becomes  of  the 


SBC.  4]  IONIZATION  AND  EXCITATION  OF  ATOMS  335 

order  of  magnitude  of  the  characteristic  temperature  LP./R,  which  as 
we  have  seen  is  of  the  order  of  10,000°.  Thus  at  ordinary  temperatures, 
from  Eq.  (4.1),  there  will  be  very  little  ionization  in  thermal  equilibrium. 
This  statement  does  not  hold,  however,  at  very  low  pressures.  We  can 
see  this  if  we  write  our  equilibrium  relation  in  terms  of  concentrations, 
following  Eq.  (1.10),  Chap.  X.  Then  we  have 

**=  -  **&).  (4.3) 

CH  * 

From  Eq.  (4.3),  we  see  that  as  the  pressure  is  reduced  at  constant  tem- 
perature, the  dissociation  becomes  greater,  until  finally  at  vanishing 
pressure  the  dissociation  can  become  complete,  even  at  ordinary  tem- 
peratures. This  is  a  result  of  importance  in  astrophysics,  as  has  been 
pointed  out  by  Saha.  In  the  solar  atmosphere,  there  is  spectroscopic 
evidence  of  the  existence  of  rather  highly  ionized  elements,  even  though 
the  temperature  of  the  outer  layers  of  the  atmosphere  is  not  high  enough 
for  us  to  expect  such  ionization,  at  ordinary  pressures.  However,  the 
pressure  in  these  layers  of  the  sun  is  extremely  small,  and  for  that  reason 
the  ionization  is  abnormally  high. 

Another  example  that  can  be  handled  by  ordinary  methods  of  chemical 
equilibrium  is  the  equilibrium  between  an  ion  and  a  neutral  atom  of 
another  substance,  in  which  the  more  electropositive  atom  is  the  one 
forming  the  positive  ion,  in  equilibrium.  Thus,  consider  the  reaction 
Li  +  Ne+  <=±  Li+  +  Ne,  in  which  Li  has  a  much  smaller  ionization  poten- 
tial than  Ne,  or  is  more  electropositive.  The  equilibrium  will  be  given 

by 


v  =  Kp(T),  (4.4) 

CLI  ^NO 

the   pressure   canceling  in  this   case.     And   the  equilibrium   constant 
KP(T)  is  given  by 


-I.P.(Ne)  +       _ 
ef)-i(Li+)-»<Ne)0  RT  ^  (^gj 

Since  the  ionization  potential  of  neon  is  greater  than  that  of  lithium,  the 
equilibrium  constant  reduces  to  zero  at  the  absolute  zero,  showing  that 
at  low  temperatures  the  lithium  is  ionized,  the  neon  unionized.  In 
other  words,  the  element  of  low  ionization  potential,  or  the  electropositive4 
element,  tends  to  lose  electrons  to  the  more  electronegative  element, 
with  high  ionization  potential.  This  tendency  is  complete  at  the  absolute 
zero.  At  higher  temperatures,  however,  as  the  mass  action  law  shows, 
there  will  be  an  equilibrium  with  some  of  each  element  ionized. 


CHAPTER  XXI 
ATOMS  AND  THE  PERIODIC  TABLE 

Interatomic  forces  form  the  basis  of  molecular  structure  and  chemis- 
try, and  we  cannot  understand  them  without  knowing  something  about 
atomic  structure.  We  shall  for  this  reason  give  in  this  chapter  a  very 
brief  discussion  of  the  nuclear  model  of  the  atom,  its  treatment  by  the 
quantum  theory,  and  the  resulting  explanation  of  the  periodic  table. 
There  is  of  course  not  the  slightest  suggestion  of  completeness  in  our 
discussion;  volumes  can  be  written  about  our  present  knowledge  of  atomic 
structure  and  atomic  spectra,  and  the  student  who  wishes  to  understand 
chemical  physics  properly  should  study  atomic  structure  independently. 
Since  however  there  are  many  excellent  treatises  available  on  the  subject, 
we  largely  omit  such  a  discussion  here,  mentioning  only  the  few  points 
that  wo  shall  specifically  use. 

1.  The  Nuclear  Atom. — An  atom  is  an  electrical  structure,  whose 
diameter  is  of  the  order  of  magnitude  of  10~8  cm.,  or  1  angstrom  unit, 
and  whoso  mass  is  of  the  order  of  magnitude  of  10~ 24  gm.  More  precisely, 
an  atom  of  unit  atomic  weight  would  have  a  mass  1.66  X  10~24  gm.,  and 
the  mass  of  any  atom  is  this  unit,  times  its  atomic  weight.  Almost  all 
the*  mass  of  the  atom  is  concentrated  in  a  small  central  body  called  the 
nucleus,  which  determines  the  properties  of  the  atom.  The  diameter  of 
the  nucleus  is  of  the  order  of  magnitude  of  10™13  cm.,  a  quantity  small 
enough  so  that  it  can  be  neglected  in  practically  all  processes  of  a  chemical 
nature.  The  nucleus  carries  a  positive  charge  of  electricity.  This 
charge  is  an  integral  multiple  of  a  unit  charge,  generally  denoted  by  the 
letter  c,  equal  to  4.80  X  10~10  e.s.u.  of  charge.  The  integer  by  which 
we  must  multiply  this  unit  to  get  the  charge  on  the  nucleus  is  called  the 
atomic  number,  and  is  often  denoted  by  Z.  This  atomic  number  proves 
to  be  the  ordinal  number  of  the  corresponding  element  in  the  periodic 
table  of  the  elements,  as  used  by  the  chemists.  Thus  for  the  first  few 
elements  wo  have  hydrogen  H,  Z  =  1 ;  helium  He,  Z  =  2;  lithium  Li, 
Z  =  3;  and  so  on,  up  to  uranium  U,  Z  =  92,  the  heaviest  natural  ele- 
ment. The  electric  charge  of  the  nucleus,  or  the  atomic  number,  is  the 
determining  feature  of  the  atom  chemically,  rather  than  the  atomic 
weight.  In  a  large  number  of  cases,  there  are  several  types  of  nuclei, 
all  with  the  same  atomic  number  but  with  different  atomic  weights. 
Such  nuclei  arc  called  isotopes.  They  prove  to  have  practically  identical 

336 


SEC.  1]  ATOMS  AND  THE  PERIODIC  TABLE  337 

properties,  since  most  properties  depend  on  the  nuclear  charge,  not  its 
mass.  Almost  the  only  property  depending  on  the  mass  is  the  vibra- 
tional  frequency,  as  observed  in  molecular  vibrations,  specific  heat,  etc. 
Thus,  different  isotopes  have  different  characteristic  temperature**  and 
specific  heats,  but  since  the  masses  of  different  isotopes  of  the  same  ele- 
ment do  not  ordinarily  differ  greatly,  these  differences  are  not  very 
important.  Almost  the  only  exception  is  hydrogen,  where  the  heavy 
isotope  has  twice  the  maSvS  of  the  light  isotope,  making  the  properties 
of  heavy  hydrogen,  or  deuterium,  decidedly  different  from  those  of 
ordinary  hydrogen.  The  atomic  weights  of  isotopes  are  almost  exactly 
whole  number  multiples  of  the  unit  1.66  X  10"24  gm.,  but  the  atomic 
weight  measured  chemically  is  the  weighted  mean  of  those  of  its  various 
isotopes,  and  hence  is  not  a  very  fundamental  quantity  theoretically. 
For  our  purposes,  which  are  largely  chemical,  we  need  not  consider  the 
possibility  of  a  change  in  the  properties  of  a  nucleus.  But  many  reac- 
tions are  known,  some  spontaneous  (natural  radioactivity)  and  some 
artificial  (artificial  or  induced  radioactivity),  by  which  nuclei  can  be 
changed,  both  as  to  their  atomic;  weight  and  atomic  number,  and  hence 
converted  from  the  nuclei  of  one  element  to  those  of  another.  We  shall 
assume  that  such  nuclear  transformations  are  not  occurring  in  the  proc- 
esses we  consider. 

In  addition  to  the  nucleus,  the  atom  contains  a  number  of  light,  nega- 
tively charged  particles,  the  electrons.  An  electron  has  a  mass  of 
9.1  X  10~28  gm.,  iVrs-  of  the  mass  of  a  nucleus  of  unit  atomic  weight. 
Its  charge,  of  negative  sign,  has  the  magnitude  of  4.80  X  10~10  e.s.u., 
the  unit  mentioned  above.  There  seem  to  be  no  experiments  which 
give  information  about  its  radius,  though  there  are  some  theoretical 
reasons,  not  very  sound,  for  thinking  it  to  be  of  the  order  of  10~13  cm. ,  If 
the  atom  is  electrically  neutral,  it  must  contain  just  as  many  electrons 
as  the  nucleus  has  unit  charges;  that  is,  the  number  of  electrons  equals 
the  atomic  number.  But  it  is  perfectly  possible  for  the  atom  to  exist 
with  other  numbers  of  electrons  than  this.  If  it  loses  electrons,  becoming 
positively  charged,  it  is  a  positive  ion.  It  can  lose  any  number  of 
electrons  from  one  up  to  its  total  number  Z,  and  we  say  that  it  then  forms 
a  singly  charged,  doubly  charged,  etc.,  positive  ion.  A  positive  ion  is  a 
stable  structure,  like  an  atom,  and  can  exist  indefinitely,  so  long  as  it 
does  not  come  in  contact  with  electrons  or  matter  containing  electrons, 
by  means  of  which  it  can  neutralize  itself  electrically.  On  the  other  hand, 
an  atom  can  sometimes  attach  one  or  more  electrons  to  itself,  becoming 
a  singly  or  multiply  charged  negative  ion.  Such  a  structure  tends  to  be 
inherently  unstable,  for  it  is  negatively  charged  on  the  whole,  repelling 
electrons  and  tending  to  expel  its  own  extra  electrons  and  become  neutral 
again.  It  is  doubtful  if  any  multiply  charged  negative  ions  are  really 


338  INTRODUCTION  TO  CHEMICAL  PHYSICS          [CHAP.  XXI 

stable.  On  the  other  hand,  a  number  of  elements  form  stable,  singly 
charged,  negative  ions.  These  are  the  so-called  electronegative  elements, 
the  halogens  F,  Cl,  Br,  I,  tho  divalent  elements  O  and  S,  arid  perhaps  a 
few  others.  These  elements  have  slightly  lower  energy  in  the  form  of  a 
negative  ion,  as  F",  than  in  the  dissociated  form  of  the  neutral  atom,  as  F, 
and  a  removed  electron.  The  energy  difference  between  these  two  states 
is  called  the  electron  affinity;  as  we  see,  it  is  analogous  to  a  heat  of  reac- 
tion, for  a  reaction  like 

F~<=>F  +  e.  (1.1) 

The  energy  required  to  remove  an  electron  from  a  neutral  atom  is  its 
ionizatiori  potential;  that  required  to  remove  the  second  electron  from  a 
singly  charged  positive  ion  is  the  second  ionization  potential;  and  so  on. 
In  each  case,  the  most  easily  removed  electron  is  supposed  to  be  the  one 
considered,  some  electrons  being  much  more  easily  detachable  than 
others.  Successive  ionization  potentials  get  rapidly  larger  and  larger,  for 
as  the  ion  becomes  more  highly  charged  positively,  an  electron  is  more 
strongly  hold  to  it  by  electrostatic  forces  and  requires  more  work  to 
remove.  Ionization  potentials  and  electron  affinities,  as  we  have  already 
mentioned,  are  commonly  measured  in  electron  volts,  since  electrical 
methods  are  commonly  used  to  measure  them.  For  the  definition  of  the 
electron  volt  and  its  relation  to  thcrmodynamic  units  of  energy,  the  reader 
is  referred  to  Eq.  (1.1),  Chap.  IX,  where  it  is  shown  that  one  electron  volt 
per  atom  is  equivalent  to  23.05  kg.-cal.  per  gram  mole,  so  that  ionization 
potentials  of  several  electron  volts  represent  heats  of  reaction,  for  the 
reaction  in  which  a  neutral  atom  dissociates  into  an  electron  and  a 
positive  ion,  which  are  large,  as  measured  by  thermodynamic  standards, 
as  mentioned  in  the  preceding  chapter. 

2.  Electronic  Energy  Levels  of  an  Atom. — The  electrons  in  atoms  arc 
governed  by  the  quantum  theory  and  consequently  have  various  sta- 
tionary states  and  energy  levels,  which  arc  intimately  related  to  the 
excitation  and  ionization  potentials  and  to  the  structure  of  the  periodic 
table.  We  shall  not  attempt  here  to  give  a  complete  account  of  atomic1 
structure,  in  terms  of  electronic  levels,  but  shall  mention  only  a  few 
important  features  of  the  problem.  A  neutral  atom,  with  atomic  num- 
ber Z,  and  Z  electrons,  each  acted  on  by  the  other  (Z  —  1)  electrons  a*- 
well  as  by  the  nucleus,  forms  a  dynamical  problem  which  is  too  difficult 
to  solve  except  by  approximation,  either  in  classical  mechanics  or  in 
quantum  theory.  The  most  useful  approximation  is  to  replace  the  forcr 
acting  on  an  electron,  depending  as  it  does  on  the  positions  of  all  othei 
electrons  as  well  as  on  the  one  in  question,  by  an  averaged  force,  averaged 
over  all  the  positions  which  the  other  electrons  take  up  during  theii 
motion.  This  on  the  average  is  a  central  force;  that  is,  it  is  an  attraction 


SBC.  2]  ATOMS  AND  THE  PERIODIC  TABLE  339 

pulling  the  electron  toward  the  nucleus,  the  magnitude  depending  only 
on  the  distance  from  the  nucleus.  It  is  a  smaller  attraction  than  that 
of  an  electron  for  a  bare  nucleus,  for  the  other  electrons,  distributed 
about  the  nucleus,  exert  a  repulsion  on  the  average .  Nevertheless,  so 
long  as  it  is  a  central  force,  quantum  theory  can  quite  easily  solve  the 
problem  of  finding  the  energy  levels  and  the  average  positions  of  the 
electrons. 

An  electron  in  a  central  field  has  three  quantum  numbers,  connected 
with  the  three  dimensions  of  space.  One,  called  the  azimuthal  quantum 
number,  is  denoted  by  I,  and  measures  the  angular  momentum  of  the 
electron,  in  units  of  h/2ir.  Just  as  in  the  simple  rotator,  discussed  in 
Section  3,  Chap.  Ill,  the  angular  momentum  must  be  an  integer  times 
h/2ir,  and  here  the  integer  is  I,  taking  on  the  values  0,  1,  2,  ...  For 
each  value  of  I,  we  have  a  series  of  terms  or  energy  levels,  given  by 
integral  values  of  a  second  quantum  number,  called  the  principal  or 
total  quantum  number,  denoted  by  n,  and  by  convention  taking  on  the 
values  I  +  1,  I  +  2,  •  •  •  Following  spectroscopic  notation,  all  the 
levels  of  a  given  I  value  are  grouped  together  to  form  a  series  and  are 
denoted  by  a  letter.  Thus  I  =  0  is  denoted  by  s  (for  the  spectroscopic 
Sharp  series),  I  =  1  by  p  (for  the  Principal  series),  I  —  2  by  d  (for  the 
Diffuse  series),  I  =  3  by/  (for  the  Fundamental  series),  and  I  =  4,  5,  6, 
.  .  .  by  0,  h,  j,  .  .  .  ,  using  further  letters  of  the  alphabet.  A  given 
energy  level  of  the  electron  is  denoted  by  giving  its  value  of  n,  and  then 
the  letter  giving  its  I  value;  as  3p,  a  level  with  n  =  3,  I  =  1.  The  third 
quantum  number  is  connected  with  space  quantization,  as  discussed  in 
Sec.  3,  Chap.  IX,  and  is  denoted  by  mi.  Not  only  the  angular  momentum 
I  is  quantized,  but  also  its  component  along  a  fixed  direction  in  space, 
and  this  is  equal  to  mih/2ir.  The  integer  mi,  then,  can  go  from  the  limits 
of  lh/2ir  (when  the  angular  momentum  points  along  the  direction  in 
question)  to  —lh/2ir  (when  it  is  oppositely  directed),  resulting  in  21  +  1 
different  orientations.  Being  a  problem  with  spherical  symmetry,  the 
energy  does  not  depend  on  the  orientation  of  the  angular  momentum. 
Thus  the  21  +  1  levels  corresponding  to  a  given  n  and  Z,  but  different 
orientations,  all  have  the  same  energy,  so  that  the  problem  is  degenerate, 
and  an  «  level  has  one  sublevel,  a  p  has  three,  a  d  five,  an  /  seven,  etc. 

This  number  of  levels  is  really  doubled,  however,  by  the  electron  spin. 
An  electron  has  an  intrinsic  permanent  magnetism,  and  associated  with 
it  a  permanent  angular  momentum  of  magnitude  %h/2ir.  This  can  be 
oriented  in  either  of  two  opposite  directions,  giving  a  component  of 
±%h/2ir  along  a  fixed  direction.  This,  as  will  be  seen,  is  in  harmony 
with  the  space  quantization  just  described,  for  the  special  case  I  =  £. 
For  each  stationary  state  of  an  electron  neglecting  spin,  we  can  have  the 
two  possible  orientations  of  the  spin,  so  that  actually  an  s  level  has  two 


340  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XXI 

sublevels,  a,  p  has  six,  a  d  ten,  an  /  fourteen.     These  numbers  form  the 
basis  of  the  structure  of  the  periodic  table. 

The  energies  of  these  energy  levels  can  be  given  exactly  only  in  the 
case  of  a  single  electron  rotating  about  a  nucleus  of  charge  Z  units,  in  the 
absence  of  other  electrons  to  shield  it.  In  this  case,  the  energy  is  given 
by  Bohr's  formula 


In  Eq.  (2.1),  m  is  the  mass  of  an  electron  (9.1  X  10~28  gm.),  e  is  its  charge 
(4.80  X  10~10  c.s.u.),  K  is  the  so-called  Rydberg  number,  109,737  cm.-1, 
so  that  Rhc,  where  c  is  the  velocity  of  light  (3.00  X  1010  cm.  per  second), 
is  an  energy,  equal  to  2.17  X  10~"n  erg,  or  13.56  electron  volts,  or  313 
kg.-cal.  per  gram  mole.  The  zero  of  energy  is  the  state  in  which  the 
electron  reaches  an  infinite  distance  from  the  nucleus  with  zero  kinetic 
energy.  In  all  the  stationary  states,  the  energy  is  less  than  this,  or  is 
negative,  so  that  the  electron  can  never  be  entirely  removed  from  the 
atom.  The1  smaller  the  integer  n,  the  lower  the  energy,  so  that  the 
lowest  states  correspond  to  n  =  1,  2,  etc.  At  the  same  time,  the  lower 
the  energy  is,  the  more  closely  bound  to  the  nucleus  the  electron  is,  so 
that  the  orbit,  or  the  region  occupied  by  the  electron,  is  small  for  small 
values  of  n.  The  tightness  of  binding  increases  with  the  nuclear  charge 
Z,  as  we  should  expect,  and  at  the  same  time  the  size  of  the  orbit  decreases. 
We  notice  that  for  an  electron  moving  around  a  nucleus,  the  levels  of 
different  series,  or  different  I  values,  all  have  the  same  energy  provided 
they  have  the  same  principal  quantum  number  n. 

For  a  central  field  like  that  actually  encountered  in  an  atom,  the 
energy  levels  are  quite  different  from  those  given  by  Eq.  (2.1).  They 
are  divided  quite  sharply  into  two  sorts:  low-lying  levels  corresponding 
to  orbits  wholly  within  the  atom,  and  high  levels  corresponding  to  orbits 
partly  or  wholly  outside  the  atom.  For  levels  of  the  first  type,  the 
energy  is  given  approximately  by  a  formula  of  the  typo 

E  =  -Rhc(Z--Z-Z«)\  (2.2) 

In  Eq.  (2.2),  Z0  is  called  a  shielding  constant.  It  measures  the  effect 
of  the  other  electrons  in  reducing  the  nuclear  attraction  for  the  electron 
in  question.  It  is  a  function  of  n  and  /,  increasing  from  practically  zero 
for  the  lowest  n  values  to  a  value  only  slightly  less  than  Z  for  the  outer- 
most orbits  within  the  atom.  For  levels  outside  the  atom,  on  the  other 
hand,  the  energy  is  given  approximately  by 

E  =  -Rhc-  (2.3) 


SEC.  2] 


ATOMS  AND  THE  PERIODIC  TABLE 


341 


Here  6  is  called  a  quantum  defect.  It  depends  strongly  on  Z,  but  is 
approximately  independent  of  n  in  a  single  series,  or  for  a  single  I  value. 
The  value  of  5  decreases  rapidly  with  increasing  I]  thus  the  s  series  may 
have  a  large  quantum  defect,  the  p  series  a  considerably  smaller  one,  and 
the  d  and  higher  series  may  have  very  small  values  of  6,  for  some  particu- 
lar atom.  We  may  illustrate  these  formulas  by  Fig.  XXI-1,  in  which 
the  energies  of  an  electron  in  a  contra!  field  representing  ooppor,  as  a 
function  of  n,  are  shown  on  a  logarithmic  scale.  Tho  sharp  break  between 


FIG  XXI-1. —  ICnergies  of  electrons  in  tho  copper  atom,  in  Rydbcrg  unit**,  IM  a  function 
of  principal  quantum  number  ?i.  Eneigies  are  shown  on.  a  logni  ithmic  scale.  The  energies 
in  the  hydrogen  atom  are  shown  for  comparison. 

the  two  typos  of  energy  levels  is  woll  shown;  Is,  2s,  2p,  3s,  3;;,  3d  belong 
very  definitely  to  the  orbits  lying  within  the  atom,  while  tho  others  are 
outside  and  are  governed  approximately  by  Kq.  (2.3). 

It  has  been  mentioned  that  the  region  occupied  by  tho  electron's 
orbit  increases  in  volume,  as  tho  binding  energy  becomes  less  or  as  the 
quantum  number  n  increases.  For  our  later  use  in  studying  the  sizes  of 
atoms,  it  is  useful  to  know  the  size  of  the  orbit  quantitatively.  These 
sizes  are  not  definitely  determined,  for  the  electron  is  sometimes  found 
at  one  point,  sometimes  at  another,  in  a  given  stationary  state,  and  all  we 
can  give  is  the  distance  from  the  nucleus  at  which  there  is  the  greatest 


342  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XXI 

probability  of  finding  it.  This  is  not  given  by  a  simple  formula,  though 
it  can  be  computed  fairly  accurately  by  wave  mechanics,  but  to  an 
approximation  the  radius  r  of  maximum  charge  density  or  probability  is 
given  by 

fW  =  a0~;  (2.4) 

for  the  case  of  an  electron  moving  about  a  bare  nucleus  of  charge  Z,  in 
an  orbit  of  quantum  number  n,  where  a0  =  h*/4ir2me2  =  0.53  A.  This 
is  the  formula  connected  with  the  type  of  orbit  whose  energy  is  given  by 
Eq.  (2.1).  We  observe  the  increase  of  size  with  increasing  n,  and  the 
decrease  with  increasing  nuclear  charge,  which  we  have  mentioned 
before.  Similarly,  if  the  energy  is  given  by  Eq.  (2.2),  the  radius  is  given 
approximately  by 

n2 
r-max  =  a^7       „  ,  (2.5) 


and  if  the  formula  (2.3)  holds  for  the  energy,  the  radius  is 

rumx  =  «o(n  -  6)2.  (2.6) 

We  may  expect  that  the  radius  of  the  atom,  if  that  expression  has  a 
meaning,  will  be  of  the  order  of  magnitude  of  the  radius  of  the  largest 
orbit  ordinarily  occupied  by  an  electron  in  the  neutral  atom.  In  the 
case  of  copper  this  is  the  4s  orbit,  while  in  the  copper  ion  it  is  the  3d.  In 
the  next  section  we  tabulate  such  quantities  for  the  atoms,  and  in  later 
chapters  we  shall  find  these  radii  of  interest  in  connection  with  the 
dimensions  of  atoms  as  determined  in  other  ways. 

We  can  now  use  the  energy  levels  of  an  electron  in  a  central  field  in 
discussing  the  structure  of  the  atom.  At  the  outset,  we  must  use  a 
fundamental  fact  regarding  electrons:  they  obey  the  Fermi-Dirac  statis- 
tics. That  is,  no  two  electrons  can  occupy  the  same  stationary  state. 
The  principle,  stated  in  this  form,  is  often  called  the  Pauli  exclusion 
principle,  and  it  was  originally  developed  to  provide  an  explanation 
for  the  periodic  table,  which  we  shall  discuss  in  the  next  section.  As  a 
result  of  the  Pauli  exclusion  principle,  there  can  be  only  two  Is  electrons, 
two  2s's,  six  2p's,  etc.  We  can  now  describe  what  is  called  the  con- 
figuration of  an  atom  by  giving  the  number  of  electrons  in  each  quantum 
state.  In  the  usual  notation,  these  numbers  are  written  as  exponents. 
Thus  the  symbol  (ls)2(2s)2(2/;)6(3s)2(3p)6(3rf)104s  would  indicate  a  state 
of  an  atom  with  two  Is  electrons,  two  2s,  etc.,  the  total  number  of  elec- 
trons being  2  +  2  +  6  +  2  +  6  +  10  +  1  =  29,  the  number  appro- 
priate to  the  neutral  copper  atom.  If  all  the  electrons  are  in  the  lowest 
available  energy  level,  as  they  are  in  the  case  above,  the  configuration 


SEC.  2]  ATOMS  AND  THE  PERIODIC  TABLE  343 

corresponds  to  the  normal  or  ground  state  of  the  atom.  If,  on  the  other 
hand,  some  electrons  are  in  higher  levels  than  the  lowest  possible  ones,  the 
configuration  corresponds  to  an  excited  state.  In  the  simplest  case,  only 
one  electron  is  excited;  this  would  correspond  to  a  configuration  like 
(ls)2(2s)2(2p)6(3s)2(3p)H3d)10(5p)  for  copper.  To  save  writing,  the  two 
configurations  indicated  above  would  often  be  abbreviated  simply  as  4s 
and  5p,  the  inner  electrons  being  omitted,  since  they  are  arranged  as  in 
the  normal  state.  It  is  possible  for  more  than  one  electron  to  bo  excited ; 
for  instance,  we  could  have  the  configuration  which  would  ordinarily 
be  written  as  (3d)9(4p)(5s)  (the  Is,  2s,  2p,  3s,  3p  electrons  being  omitted), 
in  which  one  of  the  3d  electrons  is  excited,  say  to  the  4p  level,  and  the  4.x 
is  excited  to  the  5s  level,  or  in  which  the  3d  is  excited  to  the  5s  level,  the 
4s  to  the  4p.  (On  account  of  the  identity  of  electrons,  implied  in  the 
Fermi-Dirac  statistics,  there  is  no  physical  distinction  between  these  two 
ways  of  describing  the  excitation.)  While  more  than  two  electrons  can 
theoretically  be  excited  at  the  same  time,  it  is  very  unusual  for  this  to 
occur.  If  one  or  more  electrons  are  entirely  removed,  so  that  we  have  an 
ion,  the  remaining  electrons  will  have  a  configuration  that  ran  be  indi- 
cated by  the  same  sort  of  symbol  that  would  be  used  for  a  complete 
atom.  For  example,  the  normal  state  of  the  Cu+  ion  has  the  configura- 
tion (ls)2(2«)2(2p)6(3«)2(3p)6(3d)10. 

The  energy  values  which  we  most  often  wish  are  excitation  and 
ionization  potentials,  the  energies  required  to  shift  one  or  more  electrons 
from  one  level  to  another,  or  the  differences  of  energy  between  atoms  or 
ions  in  different  configurations.  We  can  obtain  good  approximations 
to  these  from  our  one-electron  energy  values  of  Eqs.  (2.2)  and  (2.3).  The 
rule  is  simple:  the  energy  required  to  shift  an  electron  from  one  energy 
level  to  another  in  the  atom  is  approximately  equal  to  the  difference  of 
the  corresponding  one-electron  energies.  If  two  electrons  are  shifted, 
we  simply  add  the  energy  differences  for  the  two.  This  rule  is  only 
qualitatively  correct,  but  is  very  useful.  In  particular,  sinee  the  absolute 
values  of  the  quantities  (2.2)  and  (2.3)  represent  the  energies  required  to 
remove  the  corresponding  electrons  from  the  central  field,  the  same 
quantities  in  turn  are  approximate  values  of  the  ionization  potentials  of 
the  atom.  An  atom  can  be  ionized  by  the  removal  of  any  one  of  its 
electrons.  The  ordinary  ionization  potential  is  the  work  required  to 
remove  the  most  loosely  bound  electron;  in  copper,  for  instance,  the 
work  required  to  remove  the  4s  electron  from  the  neutral  atom.  But 
any  other  electron  can  be  removed  instead,  though  it  requires  more 
energy.  If  an  atom  is  bombarded  by  a  fast  electron,  the  most  likely 
type  of  ionization  process  is  that  in  which  an  inner  electron  is  removed, 
as  for  instance  a  Is,  2s,  2p,  etc.  For  such  an  ionization,  in  which  the 
ionization  potential  may  be  many  thousands  of  volts,  the  impinging 


344  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XXI 

electron  must  of  course  have  an  energy  greater  than  the  appropriate 
ionization  potential.  After  an  inner  electron  is  knocked  out,  in  this 
way,  a  transition  is  likely  to  occur  in  which  one  of  the  outer  electrons  falls 
into  the  empty  inner  shell,  the  emitted  energy  coming  off  as  radiation 
of  very  high  frequency.  It  is  in  this  way  that  x-rays  are  produced,  and 
on  account  of  their  part  in  x-ray  emission,  the  inner  energy  levels  are 
known  by  a  notation  derived  from  x-rays.  Thus  the  Is  electrons  are 
known  as  the  K  shell  (since  the  x-rays  emitted  when  an  electron  falls 
into  the  K  shell  are  called  the  K  series  of  x-rays),  and  2s  and  2p  grouped 
together  are  the  L  shell,  the  3s,  3p,  and  3d  together  are  the  M  shell,  and 
so  on. 

In  contrast  to  the  x-ray  ionization,  which  is  not  often  important  in 
chemical  problems,  an  impinging  electron  with  only  a  few  volts'  energy  is 
likely  to  excite  or  ionize  the  outermost  electron.  This  electron  has  an 
energy  given  approximately  by  Kq.  (2.3),  which  thus  gives  roughly  the 
energies  of  the  various  excited  and  ionized  levels  of  the  atom.  As  a 
matter  of  fact,  the  real  situation,  with  all  but  a  few  of  the  simplest  atoms, 
is  very  much  more  complicated  than  would  be  indicated  by  Eq.  (2.3),  on 
account  of  certain  interactions  between  the  outer  electrons  of  the  atom, 
resulting  in  what  are  called  multiplets.  A  given  configuration  of  the 
electrons,  instead  of  corresponding  to  a  single  stationary  state  of  the 
atom,  very  often  corresponds  to  a  large  number  of  energy  levels,  grouped 
more  or  less  closely  about  the  value  given  by  our  elementary  approxima- 
tion of  one-electron  energies.  An  understanding  of  this  multiplet 
structure  is  essential  to  a  real  study  of  molecular  structure,  but  we  shall 
not  follow  the  subject  far  enough  to  need  it.  One  principle  only  will 
be  of  value:  a  closed  shell  of  electrons,  by  which  we  mean  a  shell  contain- 
ing all  the  electrons  it  can  hold,  consistent  with  the  Pauli  exclusion 
principle  [in  other  words,  a  group  like  (Is)2,  (2p)6,  etc.]  contributes  noth- 
ing to  the  multiplet  structure  or  the  complication  of  the  energy  levels. 
Thus  an  atom  all  of  whose  electrons  are  in  closed  shells  (which,  as  we 
shall  see  in  the  next  section,  is  an  inert  gas)  has  no  multiplets,  and  its 
energy  level  is  single.  And  an  atom  consisting  mostly  of  closed  shells, 
but  with  one  or  two  electrons  outside  them,  has  a  multiplet  structure 
characteristic  only  of  the  electrons  outside  closed  shells.  Thus  the  alkali 
metals,  and  copper,  silver,  arid  gold,  all  have  one  electron  outside  closed 
shells  in  their  normal  state  (as  we  have  found  that  copper  has  a  4s  elec- 
tron). As  a  result,  all  these  elements  have  similar  spectra. 

3.  The  Periodic  Table  of  the  Elements. — In  Table  XXI-1  we  list 
the  elements  in  order  of  their  atomic  numbers,  which  are  given  in  addition 
to  their  symbols.  The  atoms  in  the  table  are  arranged  in  rows  and 
columns  in  such  a  way  as  to  exhibit  their  periodic  properties.  The 
diagonal  lines  are  drawn  in  such  a  way  as  to  connect  atoms  of  similar 


SBC.  3] 


ATOMS  AND  THE  PERIODIC  TABLE 


345 


properties.  Table  XXI-2  gives  the  electron  configuration  of  the  normal 
states  of  the  elements.  Table  XXI-3  gives  the  ionization  potentials 
of  the  various  electrons  in  the  lighter  atoms,  in  units  of  Rhc,  the  Rydberg 
energy,  mentioned  in  the  preceding  section.  And  Table  XXI-4  gives 
the  radii  of  the  various  orbits,  as  computed  by  wave  mechanics.  We 
can  now  use  these  tables,  and  other  information,  to  give  a  brief  discussion 
of  the  properties  of  the  elements,  particularly  in  regard  to  their  ability 
to  form  ions,  which  is  fundamental  in  studying  thoir  chemical  behavior. 
In  this  regard,  we  must  remember  that  low  ionization  potentials  cor- 
respond to  easily  removed  electrons,  high  ionization  potentials  to  tightly 
held  electrons. 

* 
TABLE  XXI-1. — THE  PERIODIC  TABLE  OK  THE  ELEMENTS 


Li    3 Na  11 

Be   4 Mgl2 

B     6 Al  13 

C     6 Si    14 

7 P    15 

S    16 
Cl  17 


N 
O     8 
F      9 


NelO- 


-A    18 


In  a  way,  the  most  distinctive  elements  are  the  inert  gases,  He,  Ne, 
A,  Kr,  and  Xe.  As  we  see  from  Table  XXI-2,  all  the  electrons  in  these 
elements  are  in  closed  shells.  They  form  no  chemical  compounds  and 
have  high  ionization  potentials,  showing  very  small  tendency  to  form 
ions.  The  reason  for  their  stability  is  fundamentally  the  fact  that 
electrons  in  closed  shells  are  difficult  to  remove,  as  is  shown  by  an  exami- 
nation of  ionization  potentials  throughout  Table  XXI-3.  That  is,  closed 
shells  form  a  very  stable  structure,  difficult  to  deform  in  such  a  way  as  to 
form  ions  or  molecules.  To  see  why  the  inert  gases  appear  where  they 
do  in  the  periodic  table,  we  may  imagine  that  we  are  building  up  the 
periodic  table,  adding  more  and  more  electrons  to  a  nucleus.  The  first 


346 


INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XXI 


TABLE  XXI-2. — ELECTRON  CONFIGURATIONS  OP  THE  ELEMENTS,  NORMAL  STATES 


K 

L 

M 

N 

0 

P 

7s 

Is 

2s 

2p 

3-9 

3p 

U 

4s 

4p 

4d 

V 

5s 

5p 

U 

6s 

6p 

6<* 

H 

1 

He 

2 

Li 

2 

1 

Be 

2 

2 

B 

2 

2 

1 

C 

2 

2 

2 

N 

2 

2 

3 

i 

0 

2 

2 

4 

, 

P 

2 

2 

5 

Ne 

2 

2 

(i 

«, 

Na 

2 

2 

6 

1 

Mg 

2 

2 

6 

2 

Al 

2 

2 

6 

2 

1 

Si 

2 

2 

6 

2 

2 

P 

2 

2 

6 

2 

3 

S 

2 

2 

6 

2 

4 

Cl 

2 

2 

6 

2 

5 

A 

2 

2 

6 

2 

6 

K 

2 

2 

6 

2 

6 

1 

Ca 

2 

2 

G 

2 

6 

2 

Sc 

2 

2 

6 

2 

6 

I 

2 

Ti 

2 

2 

6 

2 

6 

2 

2 

V 

2 

2 

6 

2 

6 

3 

2 

Cr 

2 

2 

6 

2 

6 

5 

1 

Mn 

2 

2 

6 

2 

6 

5 

2 

Fc 

2 

2 

6 

2 

6 

6 

2 

Co 

2 

2 

6 

2 

6 

7 

2 

Ni 

2 

2 

6 

2 

6 

8 

2 

Cu 

2 

2 

6 

2 

6 

10 

1 

Zn 

2 

2 

6 

2 

6 

10 

2 

Ga 

2 

2 

6 

2 

6 

10 

2 

1 

Gc 

2 

2 

6 

2 

6 

10 

2 

2 

As 

2 

2 

6 

2 

6 

10 

2 

3 

Se 

2 

2 

6 

2 

6 

10 

2 

4 

Br 

2 

2 

6 

2 

6 

10 

2 

5 

Kr 

2 

2 

6 

2 

6 

10 

2 

6 

Rb 

2 

2 

6 

2 

6 

10 

2 

6 

1 

Sr 

2 

2 

6 

2 

6 

10 

2 

6 

2 

Y 

2 

2 

6 

2 

6 

10 

2 

G 

1 

2 

Zr 

2 

2 

6 

2 

6 

10 

2 

6 

2 

2 

Cb 

2 

2 

6 

2 

6 

10 

2 

6 

4 

1 

Mo 

2 

2 

6 

2 

6 

10 

2 

6 

5 

1 

Ma 

2 

2 

6 

2 

6 

10 

2 

6 

6 

1 

Ru 

2 

2 

6 

2 

6 

10 

2 

6 

7 

1 

Rh 

2 

2 

6 

2 

6 

10 

2 

6 

8 

1 

Pel 

2 

2 

6 

2 

6 

10 

2 

6 

10 

SBC.  3] 


ATOMS  AND  THE  PERIODIC  TABLE 


347 


TABLE  XXI-2. — ELECTRON  CONFIGURATIONS  OF  THE  ELEMENTS,  NORMAL  STATES. — 

(Continued) 


K 

L 

M 

N 

0 

p 

Is 

28 

2p 

3s 

3p 

3d 

4s 

AP. 

4d 

4/ 

5s 

IP 

5rf 

6s 

6p 

6d 

7* 

Ag 

2 

2 

6 

2 

6 

10 

2 

6 

10 

T 

Cd 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

In 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

1 

Sn 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

2 

Sb 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

3 

i 

Te 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

4 

I 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

5 

Xe 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

6 

Cs 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

6 

1 

Ba 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

6 

2 

La  - 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

6 

1 

2 

Ce 

2 

2 

6 

2 

6 

10 

2 

6 

10 

1 

2 

6 

1 

2 

Pr 

2 

2 

6 

2 

6 

10 

2 

6 

10 

2 

2 

6 

1 

2 

Nd 

2 

2 

6 

2 

6 

10 

2 

6 

10 

3 

2 

(5 

1 

2 

II 

2 

2 

6 

2 

6 

10 

2 

6 

10 

4 

2 

6 

1 

2 

Sa 

2 

2 

6 

2 

6 

10 

2 

6 

10 

f> 

2 

6 

1 

2 

Er 

2 

2 

6 

2 

6 

10 

2 

6 

10 

6 

2 

6 

1 

2 

Gd 

2 

2 

6 

2 

6 

10 

2 

6 

10 

7 

2 

6 

1 

2 

Tb 

2 

2 

6 

2 

6 

10 

2 

6 

10 

8 

2 

6 

1 

2 

Ds 

2 

2 

6 

2 

6 

10 

2 

6 

10 

9 

2 

0 

1 

2 

Ho 

2 

2 

6 

2 

6 

10 

2 

6 

10 

10 

2 

6 

1 

2 

Er 

2 

2 

6 

2 

6 

10 

2 

6 

10 

11 

2 

6 

1 

2 

Tu 

2 

2 

6 

2 

6 

10 

2 

6 

10 

12 

2 

6 

1 

2 

Yb 

2 

2 

6 

2 

6 

10 

2 

6 

10 

13 

2 

6 

1 

2 

Lu 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

1 

2 

Hf 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

2 

2 

Ta 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

3 

2 

W 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

4 

2 

Re 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

0 

5 

2 

Os 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

0 

2 

Ir 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

7 

2 

Pt 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

9 

1 

Au 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

1 

Hg 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

Tl 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

1 

Pb 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

2 

Bi 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

0 

10 

2 

3 

Po 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

4 

— 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

5 

Rn 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

— 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

1 

Ra 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

2 

Ac 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

1 

2 

Th 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

2 

2 

Pa 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

3 

2 

U 

2 

2 

6 

2 

6 

10 

2 

6 

10 

14 

2 

6 

10 

2 

6 

4 

2 

348  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXI 

TABLE  XXI-3. — IONIZATION  POTENTIALS  OP  THE  LIGHTER  ELEMENTS,  IN  RYDBERGS 


K 

L 

M 

^ 
N 

O 

18 

2s 

2p 

3« 

3p 

3d 

4« 

4p 

4d 

5« 

II 

1.00 

He 

1.81 

Li 

4.80 

0.40 

Be 

(9.3) 

0.69 

B 

(15.2) 

1.29 

0.61 

C 

(22.3) 

1.51 

0.83 

N 

(31.1) 

1.91 

1.07 

O 

(41.5) 

2.10 

1.00 

F 

(53.0) 

2.87 

1.37 

Ne 

(66.1) 

3.56 

1.59 

Na 

(80.9) 

(5.10) 

2.79 

0.38 

Mg 

96.0 

(6.96) 

3  7 

0.56 

Al 

114.8 

(9.05) 

5.3 

0  78 

0.44 

Si 

135.4 

(11.5) 

7.2 

1.10 

0.60 

P 

157.8 

(14.2) 

9.4 

(1  40) 

(0  65) 

S 

181.9 

(17.2) 

11.9 

1.48 

0.76 

Cl 

207.9 

(20.4) 

14.8 

1.81 

0.96 

A 

235.7 

(23.9) 

(18.2) 

2  14 

1.15 

K 

265.6 

(27  8) 

21.5 

(2.6) 

1.2 

0.32 

Ca 

297  4 

(31  9) 

25.5 

(3.1) 

1.9 

0  45 

Sc 

331.2 

(36  2) 

30  0 

(3  6) 

2.7 

0.54 

0  .50 

Ti 

365.8 

(41.0) 

33.6 

(4  2) 

2.6 

0.51 

0.50 

V 

402  7 

(46.0) 

37  9 

(4.8) 

3.0 

0.50 

0.52 

Cr 

441   1 

(51.2) 

42.3 

(5.4) 

3.1 

0  61 

0  50 

Mn 

481  9 

(56.7) 

47  4 

(6.7) 

3.8 

0.68 

0.55 

Fe 

523.9 

62  5 

52  2 

6.9 

4.1 

0.60 

0.58 

Co 

568.1 

(68.5) 

57.7 

7.6 

4.7 

0.63 

0  66 

Ni 

614.1 

74  8 

63  2 

8.2 

5.4 

(0.68) 

0  64 

Cu 

661  6 

81  0 

fi8.9 

8  9 

5.7 

0.77 

0.57 

Zn 

711  7 

88.4 

75.4 

10.1 

6.7 

1.26 

0.69 

Ga 

765.6 

(96.0) 

84.1 

12.4 

8.8 

1.8 

0.87 

0.44 

Ge 

817.6 

(104.0) 

89.3 

13.4 

9.5 

3.2 

1.39 

0.60 

As 

874.0 

112.6 

97.4 

14.9 

10.3 

3.0 

(1-0) 

0.74 

So 

932.0 

(121.9) 

108.4 

16.7 

11.6 

3.9 

(1.7) 

0  70 

Br 

992.6 

(131.5) 

117.8 

19.1 

13.6 

5.4 

(1.9) 

0.87 

Kr 

(1055) 

(141.6) 

(127.2) 

(21.4) 

(15.4) 

(6.8) 

(2.1) 

1.03 

Rb 

1119.4 

152.0 

137.2 

(23.7) 

17.4 

(8.3) 

(2.3) 

1.46 

0.31 

Sr 

1186.0 

162.9 

147.6 

26.2 

19.6 

9.7 

2.5 

(2.1) 

0  42 

Y 

1256.1 

175.8 

159  9 

30.3 

23.3 

13  0 

4.7 

2.9 

0  48 

0  49 

Zr 

1325.7 

186.6 

170.0 

31.8 

24.4 

13.3 

3.8 

2.1 

0  53 

0  51 

Cb 

1398.5 

198.9 

181.7 

34.7 

26.9 

15.2 

4.3 

2  5 

(0.5) 

(0  5) 

Mo 

1473.4 

211.3 

193.7 

37.5 

29.2 

17.1 

5.1 

2.9 

(0.5) 

0.54 

The  ionization  potentials  tabulated  represent  in  each  case  the  least  energy  required  to  remove  the 
electron  in  question  from  the  atom,  in  units  of  the  Rydberg  energy  Rhc  (13.54  electron  volts).  Data  for 
optical  ionization  are  taken  from  Bacher  and  Goudsmit,  "Atomic  Energy  States,'1  McGraw-Hill  Book 
Company,  Inc.,  1932.  Those  for  x-ray  ionization  are  from  Siegbahn,  "  Spektroskopie  der  Rontgen- 
strahlen,"  Springer.  Intermediate  figures  are  interpolated.  Interpolated  or  estimated  values  are 
given  in  parentheses. 


SEC.  3] 


ATOMS  AND  THE  PERIODIC  TABLE 


349 


TABLE  XXI-4. — RADII  OF  ELECTRONIC  ORBITS  IN  THE  LIGHTER  ELEMENTS 

(Angstrom  units) 


K 

L 

M 

N 

Is 

25 

2p 

3s 

3p 

3d 

4s 

4p 

H 

0.53 

Ho 

0  30 

Li 

0.20 

1.50 

Be 

0  143 

1.19 

B 

0.112 

0.88 

0.85 

C 

0.090 

0.67 

0.66 

N 

0.080 

0.56 

0  53 

0 

0.069 

0.48 

0.45 

F 

0  061 

0  41 

0.38 

No 

0.055 

0.37 

0.32 

Na 

0.050 

0.32 

0  28 

1.55 

Mg 

0  046 

0.30 

0.25 

1.32 

Al 

0  042 

0.27 

0.23 

1.16 

1  21 

Si 

0  040 

0  24 

0.21 

0.98 

1.06 

P 

0.037 

0.23 

0.19 

0.88 

0.92 

S 

0  035 

0  21 

0.18 

0.78 

0.82 

Cl 

0  032 

0  20 

0.16 

0.72 

0.75 

A 

0  031 

0.19 

0,155 

0.66 

0.67 

K 

0  029 

0  18 

0  145 

0.60 

0.63 

2.20 

Ca 

0.028 

0  16 

0  133 

0.55 

0.58 

2  03 

Sc 

0.026 

0.16 

0.127 

0  52 

0  54 

0.61 

1.80 

Ti 

0.025 

0.150 

0  122 

0.48 

0.50 

0.55 

1.66 

V 

0  024 

0  143 

0.117 

0.46 

0.47 

0.49 

1  52 

Cr 

0.023 

0.138 

0.112 

0.43 

0.44 

0.45 

1.41 

Mn 

0.022 

0.133 

0  106 

0.40 

0.41 

0.42 

1  31 

Fc 

0.021 

0  127 

0  101 

0  39 

0.39 

0  39 

1.22 

Co 

0.020 

0.122 

0.096 

0.37 

0.37 

0.36 

1.14 

Ni 

0.019 

0.117 

0.090 

0.35 

0.36 

0.34 

1.07 

Cu 

0.019 

0  112 

0.085 

0.34 

0.34 

0.32 

1.03 

Zn 

0.018 

0.106 

0.081 

0.32 

0.32 

0.30 

0.97 

Ga 

0.017 

0.103 

0  078 

0.31 

0.31 

0.28 

0.92 

1.13 

Gc 

0.017 

0  100 

0.076 

0.30 

0.30 

0  27 

0.88 

1.06 

As 

0.016 

0.097 

0.073 

0.29 

0.29 

0.25 

0.84 

1.01 

Se 

0.016 

0.095 

0.071 

0.28 

0,28 

0  24 

0.81 

0.95 

Br 

0.015 

0.092 

0.069 

0  27 

0.27 

0.23 

0.76 

0  90 

Kr 

0.015 

0.090 

0.067 

0.25 

0.25 

0  22 

0.74 

0  86 

The  radii  tabulated  represent  the  distance  from  the  nucleus  at  which  the  radial  charge  density  (the 
charge  contained  in  a  shell  of  unit  thickness)  is  a  maximum.  They  are  computed  from  calculations  of 
Hartree,  in  various  papers  in  "Proceedings  of  the  Royal  Society,"  and  elsewhere.  Since  only  a  few 
atoms  have  been  computed,  most  of  the  values  tabulated  are  interpolated.  The  interpolation  should 
be  fairly  accurate  for  the  inner  electrons  of  an  atom,  but  unfortunately  is  quite  inaccurate  for  the  outer 
electrons,  so  that  these  values  should  not  be  taken  as  exact. 


350  INTRODUCTION  TO  CHEMICAL  PHYSICS         [CHAP.  XXI 

two  electrons  go  into  the  K  shell,  resulting  in  He,  a  stable  structure  with 
just  two  electrons.  The  next  electrons  go  into  the  L  shell,  with  its  sub- 
groups of  2s  and  2p  electrons.  These  electrons  are  grouped  together,  for 
they  are  not  very  different  from  hydrogenlike  electrons  in  their  energy, 
and  as  we  see  from  Eq.  (2.1),  the  energy  of  a  hydrogen  wave  function 
depends  only  on  n,  not  on  I,  so  that  the  2s  and  2p  have  the  same  energy 
in  this  case.  For  the  real  wave  functions,  as  we  see  from  Fig.  XXI-1,  for 
example,  the  energies  of  2s  and  2p  are  not  very  different  from  each  other. 
The  L  shell  can  hold  two  2s  and  six  2p  electrons,  a  total  of  eight,  and  is 
completed  at  neon,  again  a  stable  structure,  with  two  electrons  in  its  K 
shell,  eight  in  its  L  shell.  The  next  electrons  must  go  into  the  still  larger 
M  shell.  Of  its  three  subgroups,  3s,  3p,  and  3d,  the  3s  and  3p,  with 
2  +  6=8  electrons,  have  about  the  same  energy,  while  the  3d  is  defi- 
nitely more  loosely  bound.  Thus  the  3s  and  3p  electrons  are  completed 
with  argon,  with  two  K,  eight  L,  and  eight  M  electrons,  again  a  stable 
structure  and  an  inert  gas.  It  is  in  this  way  that  the  periodicity  with 
period  of  eight,  which  is  such  a  feature  of  the  lighter  elements,  is  brought 
about.  After  argon,  the  order  of  adding  electrons  is  somewhat  peculiar. 
The  next  electrons  added,  in  potassium  and  calcium,  go  into  4s  states, 
which  for  those  elements  have  a  lower  energy  than  tho  3d.  But  with 
scandium,  the  element  beyond  calcium,  the  order  of  levels  changes,  the 
3d  becoming  somewhat  more  tightly  bound.  In  all  the  elements  from 
scandium  to  copper  the  new  electrons  are  being  added  to  the  3d  level, 
the  normal  state  having  either  one  or  two  4s  electrons.  For  all  these 
elements,  the  4s  and  3d  electrons  have  so  nearly  the  same  energy  that  the 
configurations  with  no  4s  electrons,  with  one,  and  with  two,  have  approxi- 
mately the  same  energy,  so  that  there  are  many  energy  levels  near  the 
normal  state.  At  copper,  the  3d  shell  is  filled,  so  that  the  M  shell  con- 
tains its  full  number  of  2  +  6  +  10  =  18  electrons,  and  as  we  have  seen 
from  our  earlier  discussion,  there  is  one  4s  electron.  The  elements 
following  copper  add  more  and  more  4s  and  4p  electrons,  until  the  group 
of  eight  4s  and  4p's  is  filled,  at  krypton.  This  is  again  a  stable  configura- 
tion. After  this,  very  much  (he  same  sort  of  situation  is  repeated  in  the 
atoms  from  rubidium  and  strontium  through  silver,  which  is  similar  to 
copper,  and  then  through  xenon,  which  has  a  complete  M  shell,  and 
complete  4s,  4p,  4d,  5,s,  and  5p  shells.  Following  this,  the  two  electrons 
added  in  caesium  and  barium  go  into  the  6s  shell,  but  then,  instead  of 
the  next  electrons  going  into  the  5d  shell  as  we  might  expect  by  analogy 
with  the  two  preceding  groups  of  the  periodic  table,  they  go  into  the  4/ 
shell,  which  at  that  point  becomes  the  more  tightly  bound  one.  The 
fourteen  elements  in  which  the  4/  is  being  filled  up  are  the  rare  earths,  a 
group  of  extraordinarily  similar  elements  differing  only  in  the  number  of 
4/  electrons,  which  have  such  small  orbits  and  are  so  deeply  buried  inside 


Sfic.  3]  ATOMS  AND  THE  PERIODIC  TABLE  351 

the  atom  that  they  have  almost  no  effect  on  chemical  properties.  After 
finishing  the  rare  earths,  the  5d  shell  is  filled,  in  the  elements  from  hafnium 
to  platinum,  and  the  next  element,  gold,  is  similar  to  copper  and  silver. 
Then  the  6s  and  6p  shells  are  completed,  leading  to  the  heaviest  inert 
gas,  radium  emanation,  and  finally  the  7s  electrons  are  added  in  radium, 
with  presumably  the  Qd  or  5/  in  the  remaining  elements  of  the  periodic 
table. 

Now  that  we  have  surveyed  the  elements,  we  are  in  position  to  under- 
stand why  some  atoms  tend  to  form  positive  ions,  some  negative.  The 
general  rule  is  simple:  atoms  tend  to  gain  or  lose  electrons  enough  so 
that  the  remaining  electrons  will  have  a  stable  structure,  like  one  of  the 
inert  gases,  or  some  other  atom  containing  completed  groups  or  subgroups 
of  electrons.  The  reason  is  plain  from  Table  XXI-3,  at-  least  as  far  as  the 
formation  of  positive  ions  is  concerned:  the  electrons  outside  closed  shells 
have  much  smaller  ionization  potentials  than  those  in  closed  shells  and 
are  removed  by  a  much  smaller  amount  of  energy.  Thus  the  alkali 
metals,  lithium,  sodium,  potassium,  rubidium,  and  caesium,  each  have 
one  easily  removed  electron  outside  an  inert  gas  shell,  and  this  electron 
is  often  lost  in  chemical  processes,  resulting  in  a  positive  ion.  The 
alkaline  earths,  beryllium,  magnesium,  calcium,  strontium,  and  barium, 
similarly  have  two  easily  removable  electrons  and  become  doubly  charged 
positive  ions.  Boron  and  aluminum  lose  three  electrons.  Occasionally 
carbon  and  silicon  lose  four  and  nitrogen  five,  but  these  processes  are 
certainly  very  rare  and  perhaps  never  occur.  The  electrons  become  too 
strongly  bound  as  the  shell  fills  up  for  them  to  be  removed  in  any  ordinary 
chemical  process.  But  oxygen  sometimes  gains  two  electrons  to  form  the 
stable  neon  structure,  and  fluorine  often  gains  one,  forming  doubly  and 
singly  charged  negative  ions  respectively.  Similarly  chlorine,  bromine, 
and  iodine  often  gain  one  electron,  and  possibly  sulphur  occasionally 
gains  two.  In  the  elements  beyond  potassium,  the  situation  is  somewhat 
different.  Potassium  and  calcium  tend  to  lose  one  and  two  electrons 
apiece,  to  simulate  the  argon  structure.  But  the  next  group  of  elements, 
from  scandium  through  nickel,  ordinarily  called  the  iron  group,  tend  to 
lose  only  two  or  three  electrons  apiece,  rather  than  losing  enough  to  form 
a  closed  shell.  Nickel  contains  a  completed  K,  L,  and  M  shell  and  is  a 
rather  stable  structure  itself,  though  not  so  much  so  as  an  inert  gas;  and 
the  next  few  elements  tend  to  lose  electrons  enough  to  have  the  nickel 
structure.  Thus  copper  tends  to  lose  one,  zinc  two,  gallium  three,  and 
germanium  four  electrons,  being  analogous  to  a  certain  extent  to  sodium, 
magnesium,  aluminum,  and  silicon.  Coming  to  the  end  of  this  row, 
selenium  tends  to  gain  two  electrons  like  oxygen  and  sulphur,  and  bromine 
to  gain  one.  Similar  situations  are  met  in  the  remaining  groups  of  the 
periodic  table. 


CHAPTER  XXII 
INTERATOMIC  AND  INTERMOLECULAR  FORCES 

One  of  the  most  fundamental  problems  of  chemical  physics  is  the  study 
of  the  forces  between  atoms  arid  molecules.  We  have  seen  in  many 
preceding  chapters  that  these  forces  are  essential  to  the  explanation  of 
equations  of  state,  specific  heats,  the  equilibrium  of  phases,  chemical 
equilibrium,  and  in  fact  all  the  problems  we  have  taken  up.  The  exact 
evaluation  of  these  forces  from  atomic  theory  is  one  of  the  most  difficult 
branches  of  quantum  theory  and  wave  mechanics.  The  general  prin- 
ciples on  which  the  evaluation  is  based,  however,  are  relatively  simple, 
and  in  this  chapter  we  shall  learn  what  these  general  principles  are,  and 
see  at  least  qualitatively  the  sort  of  results  they  lead  to. 

There  is  one  general  point  of  view  regarding  interatomic  forces  which 
is  worth  keeping  constantly  in  mind.  Our  problem  is  really  one  of  the 
simultaneous  motion  of  the  nuclei  and  electrons  of  the  atomic  or  molecular 
system.  But  the  electrons  are  very  much  lighter  than  the  nuclei  and 
move  very  much  faster.  Thus  it  forms  a  very  good  approximation  to 
assume  first  that  the  nuclei  are  at  rest,  with  the  electrons  moving  around 
them.  We  then  find  the  energy  of  the  whole  system  as  a  function  of  the 
positions  of  the  nuclei.  If  this  energy  changes  when  a  particular  nucleus 
is  moved,  we  conclude  that  there  is  a  force  on  that  nucleus,  such  that  the 
force  times  the  displacement  equals  the  work  done,  or  change  of  energy. 
This  force  can  be  used  in  discussing  the  motion  of  the  nucleus,  studying 
its  translational  or  vibrational  motion,  as  we  have  had  occasion  to  do  in 
previous  chapters.  Our  fundamental  problem,  then,  is  to  find  how  the 
energy  of  a  system  of  atoms  changes  as  the  positions  of  the  nuclei  are 
changed.  In  other  words,  we  must  solve  the  problem  of  the  motion  of 
the  electrons  around  the  nuclei,  assuming  they  are  fixed  in  definite  posi- 
tions. The  forces  between  electrons  arc  essentially  electrostatic;  there 
are  also  magnetic  forces,  but  they  are  ordinarily  small  enough  so  that  they 
can  practically  be  neglected.  Then  the  problem  of  solving  for  the  motion 
of  the  electrons  can  be  separated  into  several  parts.  It  is  a  little  difficult 
to  know  where  to  start  the  discussion,  for  there  is  a  sort  of  circular  type 
of  argument  involved.  Suppose  we  start  by  knowing  how  the  electrons 
move.  Then  we  can  find  their  electrical  charge  distribution,  and  from 
that  we  can  find  the  electrostatic  field  at  any  point  of  space.  But  this 
field  is  what  determines  the  forces  acting  on  the  electrons.  And  those 
forces  must  lead  to  motions  of  the  electrons  which  are  just  the  ones  we 

352 


SEC.  1]          INTERATOMIC  AND  INTERMOLECULAR  FORCES  353 

started  with.  An  electric  field  of  this  type,  leading  to  motions  of  the 
electrons  such  that  the  electrons  themselves,  together  with  the  nuclei, 
can  produce  the  original  field,  is  sometimes  called  a  self-consistent  field. 

As  a  first  attempt  to  solve  the  problem,  let  us  assume  that  each  atom 
is  a  rigid  structure  consisting  of  a  nucleus  and  a  swarm  of  electrons  sur- 
rounding it,  not  affected  by  tho  presence  of  neighboring  atoms.  This 
leads  to  a  problem  in  pure  electrostatics:  the  energy  of  the  whole  system, 
as  a  function  of  the  positions  of  the  nuclei,  is  simply  the  electrostatic 
energy  of  interaction  between  the  charges  of  the  various  atoms.  This 
electrostatic  energy  is  sometimes  called  the  Coulomb  energy,  since  it 
follows  directly  from  Coulomb's  law  stating  that  the  force  between  two 
charges  equals  the  product  of  the  charges  divided  by  the  square  of  the 
distance  between.  This  first  approximation,  however,  is  far  from  ade- 
quate, for  really  tho  electrons  of  each  atom  will  be  displaced  by  the 
electric  fields  of  neighboring  atoms.  We  shall  later,  then,  have  to  study 
this  deformation  of  the  atoms  and  to  find  tho  forces  bo  t  ween  tho  distorted 
atoms. 

1.  The  Electrostatic  Interactions  between  Rigid  Atoms  or  Molecules 
at  Large  Distances.  —  In  this  section,  we  are  to  find  the  forces  between 
two  atoms  or  ions  or  molecules,  assuming  that  each  can  be  represented 
by  a  rigid,  undistortod  distribution  of  charge.  The  discussion  of  these 
electrostatic,  or  Coulomb,  forces  is  conveniently  divided  into  two  parts. 
First,  we  find  the  electric  field  of  the  first  charge  distribution  at  all  points 
of  space;  then,  we  find  the  force  on  the  second  charge  distribution  in  this 
field.  By  fundamental  principles  of  electrostatics,  tho  force  on  tho 
second  distribution  exerted  by  the  first  is  equal  and  opposite  to  the  force 
on  the  first  exerted  by  the  second,  if  we  make  a  corresponding  calculation 
of  the  field  exerted  by  the  second  on  the  first.  Let  us  first  consider,  then, 
the  field  of  a  charge  distribution  consisting  of  a  number  of  charges  e», 
located  at  points  with  coordinates  xlj  y^  zt-.  Rather  than  find  the  field,  it 
is  more  convenient  to  compute  the  potential,  the  sum  of  the  terms  et-/rt- 
for  the  charges,  where  r»  is  the  distance  from  the  charge  to  the  point  x,  y,  z 
where  the  potential  is  being  found.  That  is,  rt-  is  the  length  of  a  vector 
whose  components  are  x  —  xlt  y  —  ?/t,  z  —  zt,  so  that  we  have 


-~*7)2  +72/~^7)2~  +  '(z  -  *J2,  (1.1) 

and  the  potential  is 


(1.2) 


There  is  a  very  important  way  of  expanding  the  potential  (1.2),  in 
case  we  wish  its  value  at  points  far  from  the  center  of  the  charge  distribu- 


354 


INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 


tion.  This  is  the  case  which  we  wish  in  investigating  the  forces  between 
two  atoms  or  molecules  at  a  considerable  distance  from  each  other.  Let 
us  assume,  then,  that  all  the  charges  ei  arc  located  near  a  point  which  we 
may  choose  to  be  the  origin,  so  that  all  the  z/s,  y/s,  and  z/s  are  small,  and 
let  us  assume  that  the  point  #,  y,  z  where  we  are  finding  the  potential  is 
far  off,  so  that  r  =  \/x2  +  y2  +  z2  is  large.  Then  we  can  expand  the 
potential  in  power  series  in  zt-,  y^  and  z,,  regarded  as  small  quantities. 
We  have 


•  •     (1.3) 
=  z,  =  0. 

(1.4) 
(1.5) 


The  derivatives  of  (l/r»)  are  to  be  computed  when  xt  = 
But  from  Eq.  (1.1)  we  have 


When  xt  =  0,  this  becomes 


rt  —  x 


L(i\ 

dx\rj 


1  x 


On  the  other  hand,  we  have 


d_  /A  ^  _  1  £ 
ax\r/  r2  r  ' 


0.6) 


Thus,  comparing  Eqs.  (1.5)  and  (1.6),  we  can  rewrite  Eq.  (1.3)  as 
d 


(1.7) 


From  Eq.  (1.7),  the  potential  of  the  charge  distribution  depends  on  the 
quantities  Set-,  I>elxiy  Seti/t,  SetZi,  and  higher  terms  such  as  Setz?,  etc., 
which  we  have  not  written. 

The  quantity  2e*  is  simply  the  total  charge  of  the  distribution,  and 
the  first  term  of  Eq.  (1.7)  is  the  potential  of  the  total  charge  at  a  distance 
r.  This  term,  then,  is  just  what  we  should  have  if  the  total  charge  were 
concentrated  at  the  origin.  The  next  three  terms  can  be  grouped 


SBC.  1]  INTERATOMIC  AND  INTERMOLECULAR  FORCES  355 


together.  The  quantities  Ze0<9  2^-,  Se^<  form  the  three  components 
of  a  vector,  which  is  known  as  the  dipole  moment  of  the  distribution.  A 
dipole  is  a  pair  of  equal  charges,  say  of  charge  +q  and  —qy  separated  by 
a  distance  d.  For  the  sake  of  argument  let  the  charges  be  located  along 
the  x  axis,  at  d/2  and  —  d/2.  Then  the  three  quantities  above  would  be 


qd>    e*'  ~    fi*  =  °-  That  is>  the  dipol° 

moment  is  equal  in  this  case  to  the  product  of  the  charge  and  the  distance 


--Line  of  force 


FIG.  XXII-1.  —Lines  of  foice  and  equipotentials  of  a  <lip<il«» 

of  separation,  and  it  points  along  the  axis  of  the  dipole,  from  tin*  negative 
to  the  positive  end.  We  now  see  that  as  far  as  the  terms  written  in 
Eq.  (1.7)  are  concerned,  any  two  distributions  with  the  same  net  charge 
and  the  same  dipole  moment  will  have  the  same  potential.  In  the 
particular  case  mentioned  above,  the  potential,  using  Eqa.  (1.6)  and 
(1.7),  is  (qd/r*)(x/r).  Here  x/r  is  simply  the  cosine  of  the  angle  between 
the  radius  r  and  the  x  axis,  a  factor  depending  on  the  direction  but  not 
the  magnitude  of  r.  As  far  as  magnitude  is  concerned,  then,  the  poten- 
tial decreases  as  1/r2,  in  contrast  to  the  potential  of  a  point  charge, 
which  falls  off  as  1/r.  Thus  at  large  distances  the  potential  of  a  dipole  is 
unimportant  compared  to  that  of  a  point  charge.  In  Fig.  XXII-1,  we 


356  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 

show  the  equipotentials  of  this  field  of  a  dipole  and  the  lines  of  force, 
which  are  at  right  angles  to  the  equipotentials,  and  indicate  the  direction 
of  the  force  on  a  point  charge  in  the  field.  The  lines  of  force  are  those 
familiar  from  magnetostatics,  from  the  problem  of  the  magnetic  field  of  a 
bar  magnet,  which  can  be  approximated  by  a  magnetic  dipole. 

In  addition  to  the  terms  of  the  expression  (1.7),  there  are  terms 
involving  higher  powers  of  x^  y^  and  Zi,  and  at  tho  same  time  higher 
derivatives  of  1  /r,  so  that  these  terms  fall  off  more  rapidly  with  increasing 
distance.  Tho  next  terms  after  the  ones  written,  quadratic  in  the  #/s, 
and  with  a  potential  falling  off  as  1/r3,  are  called  quadrupole  terms,  the 
corresponding  moment  being  called  a  quadrupolo  moment.  We  shall 
not  have  occasion  to  use  quadrupole  moments  and  hence  shall  not  develop 
their  theory  here,  though  sometimes  they  are  important. 

Now  that  we  have  found  the  nature  of  the  potential  of  a  charge 
distribution,  we  can  ask  what  sorts  of  eases  we  are  likely  to  find  with  real 
atoms,  molecules,  and  ions.  First  we  consider  a  neutral  atom.  Since 
there  are  as  many  electrons  as  arc  necessary  to  cancel  the  positive  nuclear 
charge,  2et-  is  zero,  and  there  is  no  term  in  the  potential  falling  off  as 
1/r.  The  atom  at  any  instant  will  have  a  dipole  moment,  however;  the 
electrons  move  rapidly  from  place  to  place,  and  it  is  unlikely  that  they 
would  be  so  arranged  at  a  given  instant  that  the  dipole  moment  was 
exactly  zero,  though  it  is  likely  to  be  small,  since  some  electrons  will  be 
on  one  side  of  the  nucleus,  others  on  the  other  side.  On  account  of  the 
motion  of  the  electrons,  this  dipole  moment  will  be  constantly  fluctuating 
in  magnitude  and  direction.  It  is  not  hard  to  show  by  wave  mechanics 
that  its  average  value  must  be  zero.  Now,  for  most  purposes,  we  care 
only  about  the  average  dipole  moment,  for  ordinarily  we  are  interested 
only  in  the  time  average  force  between  atoms  or  molecules,  and  the 
fluctuations  will  average  to  zero.  Thus,  generally,  we  treat  the  dipole 
moment  of  the  atom  as  being  zero.  Only  two  important  cases  come  up  in 
which  the  fluctuating  dipole  moment  is  of  importance.  One  does  not 
concern  interatomic  forces  at  all:  it  is  the  problem  of  radiation.  In 
Chap.  XIX,  Sec.  3,  we  have  mentioned  that  an  oscillating  electric  charge 
in  classical  theory  radiates  energy  in  the  form  of  electromagnetic  waves. 
It  turns  out  that  the  oscillating  dipole  moment  which  we  have  mentioned 
here  is  closely  connected  with  the  radiation  of  light  in  the  quantum 
theory.  The  frequencies  present  in  its  oscillatory  motion  are  those 
emitted,  according  to  Bohr's  frequency  condition,  and  there  is  a  close 
relation  between  the  amplitude  of  any  definite  frequency  in  the  oscillation 
and  the  intensity  of  the  corresponding  frequency  of  radiation.  The 
other  application  of  the  fluctuating  dipole  moment  comes  in  the  calcula- 
tion of  Van  der  Waals  forces,  which  we  shall  consider  later.  It  appears 
that  the  fluctuating  external  field  resulting  from  the  fluctuating  dipole 


SBC.  1]          INTERATOMIC  AND  INTERMOLECULAR  FORCES  357 

moment  can  produce  displacements  of  charge  in  neighboring  atoms,  in 
phase  with  the  fluctuations.  The  force  exerted  by  the  fluctuating  field 
on  this  displaced  charge  does  not  average  to  zero,  on  account  of  the1 
phase  relations,  but  instead  results  in  a  net  attraction  between  the 
molecules,  which  as  we  shall  see  is  the  Van  der  Waals  attraction.  It  is 
rather  natural  from  what  we  have  said  that  it  is  possible  in  wave  mechan- 
ics to  give  a  formula  for  the  Van  der  Waals  force  between  two  atoms 
which  depends  on  the  probabilities  of  the  various  optical  transitions 
which  the  atoms  can  make,  though  we  shall  not  be  able  to  state  this 
formula  since  it  involves  too  much  application  of  quantum  theory. 

As  far  as  the  time  average  is  concerned,  we  have  seen  that  an  atom 
has  no  field  coming  from  its  net  charge  or  from  its  dipolc  moment.  As  a 
matter  of  fact,  in  most  important  cases,  an  atom  has  no  not  field  at  all  at 
external  points.  The  reason  is  that  atoms,  at  least  in  the  special  ease 
where  all  their  electrons  are  in  closed  shells,  as  in  inert  gas  atoms,  arc 
spherically  symmetrical  in  their  average  charge  distributions.  This  can 
be  proved  from  wave  mechanics  and  is  a  property  of  closed  shells.  But 
it  is  a  familiar  theorem  of  electrostatics  that  a  spherically  symmetrical 
charge  distribution  has  a  field  just  equal  to  that  which  it  would  have  if 
all  its  charge  were  placed  at  the  center.  Thus  a  neutral  atom  has  no 
external  field.  The  reason  is  seen  easily  from  Eq.  (1.7).  Each  term  of 
this  expression  after  the  first  one  depends  on  the  angle  between  the  radius 
vector  and  the  axes.  This  is  plain  for  the  terms  written,  where  we  have 
seen  that  they  vary  as  the  cosines  of  the  angles  between  the  radius  and 
the  x,  y,  and  z  axes  respectively,  but  it  proves  to  be  true  also  for  the 
remaining  terms.  But  a  spherically  symmetrical  charge  distribution 
must  obviously  have  a  spherically  symmetrical  potential,  so  that  all 
these  terms  depending  on  angles  must  be  zero.  In  other  words,  a  spher- 
ically symmetrical  distribution,  like  an  atom,  not  only  has  no  average 
dipole  moment,  but  has  no  average  quadrupole  moment  or  moment  of 
any  higher  order. 

Next  after  a  neutral  atom,  we  may  consider  a  positive  or  negative 
ion  of  a  single  atom,  such  as  Na+,  Ba++,  or  Cl~.  As  we  have  seen  in  the 
preceding  chapter,  such  an  ion  always  has  the  configuration  of  an  inert 
gas,  and  hence  is  always  spherically  symmetrical  on  the  average.  Thufc 
an  ion  has  no  dipole  or  higher  moments,  and  its  potential  and  field  are 
just  as  if  its  whole  charge  were  concentrated  at  the  nucleus.  As  a  next 
more  complicated  example,  we  take  a  molecule,  charged  or  uncharged, 
formed  from  two  or  more  atoms  or  ions.  If  the  molecule  is  charged, 
forming  an  ion  like  NH4+,  OH~,  NO3~,  S04  ,  etc.,  then  in  the  first  place 
it  has  a  term  in  the  potential  varying  as  1/r,  determined  by  the  total 
charge  on  the  ion.  In  addition  to  this,  the  ion  or  molecule  may  have  a 
dipole  moment.  When  we  come  to  discussing  specific  ions  and  molecules, 


358  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 

in  later  chapters,  we  shall  see  which  ones  have  dipole  moments,  which 
do  not;  in  general,  for  there  to  be  a  dipole  moment  different  from  zero,  the 
ion  or  molecule  must  be  unsymmetrical  in  some  way,  with  positive  charge 
localized  on  one  side,  negative  on  the  other.  The  ions  NHU*,  NOa ",  and 
864 — ,  as  we  shall  see,  prove  to  be  very  symmetrical,  and  have  no  dipole 
moment,  while  OH~  has  a  dipole  moment,  the  negative  charge  being  at 
the  oxygen  end,  the  positive  at  the  hydrogen  end.  Similarly  there  are 
some  unsymmetrical  neutral  molecules  which  have  dipole  moments. 
An  example  is  HC1,  in  which  the  H  end  tends  to  be  positive,  the  Cl 
negative.  The  dipole  moments  have  been  measured  in  many  of  these 
cases  and  are  generally  found  to  be  much  less  than  one  would  suppose 
from  a  crude  ionic  picture.  One  might  at  first  think,  for  instance,  that 
HC1  was  made  of  a  H+  and  a  Cl"  ion,  joined  together  without  distortion, 
so  that  each  was  spherically  symmetrical.  Then  the  resulting  charge 
distribution  would  have  a  field  at  external  points  like  a  unit  positive 
charge  at  the  position  of  the  hydrogen  nucleus,  and  a  unit  negative  charge 
at  the  chlorine  nucleus,  and  the  dipole  moment  would  equal  the  product 
of  the  electronic  charge  and  the  internuclear  distance.  The  measured 
dipole  moment  is  only  a  small  fraction  of  this,  showing  that  there  has 
been  a  large  distortion  of  the  electronic  distribution  in  the  process  of 
forming  the  molecule.  This  is  the  sort  of  distortion  that  we  must  take 
up  in  a  later  section. 

We  see,  then,  that  at  a  considerable  distance  a  single  atom  has  no 
electric  field,  an  ion  consisting  of  a  single  charged  atom  has  a  field  like  a 
point  charge  concentrated  at  its  center,  and  a  molecule  or  ion  consisting 
of  several  atoms  or  ions  may  have  in  addition  a  dipole  moment,  with  its 
accompanying  field,  as  well  as  having  the  field  of  its  net  charge,  if  it  is  an 
ion.  In  addition,  the  molecule  or  molecular  ion  may  have  quadrupole 
and  higher  moments.  The  effect  of  these  is  usually  small  compared  to 
the  others,  but  in  the  case  of  an  uncharged  molecule  with  no  dipole 
moment,  the  quadrupole  term  would  be  the  first  important  one  in  the 
expansion  of  the  field.  Having  found  the  nature  of  the  field  of  an  atom 
or  ion,  our  next  problem  is  to  find  the  forces  exerted  by  this  field  on 
another  atom  or  ion,  always  assuming  both  to  be  rigid  charge  distribu- 
tions. Fundamentally,  the  problem  is  very  simple:  the  force  exerted  by 
the  field  of  one  atom  or  ion  on  each  element  of  charge  of  the  second  atom 
or  ion  is  simply  the  product  of  the  field  intensity  and  the  charge,  by 
definition,  and  we  need  merely  treat  the  problem  as  one  in  statics,  adding 
the  forces  vectorially  to  find  the  total  force  on  the  atom  or  ion,  and 
adding  their  moments  about  the  center  of  gravity  to  get  the  resultant 
moment  or  torque. 

Thus,  let  the  potential  of  the  electrostatic  field  be  <£,  and  let  the 
field  strength  have  components  Ex,  Ev,  E»,  where  by  well-known  methods 


SBC.  1]          INTERATOMIC  AND  INTERMOLECULAR  FORCES  359 

of  electrostatics  the  field  is  the  negative  of  the  derivative  of  <£  with  respect 
to  displacements  along  the  axes,  so  that  the  product  of  the  force  and  the 
displacement  gives  the  work  done,  or  the  negative  of  the  change  of 
potential.  That  is, 


The  components  Ex,  Ev,  and  Ez  will  be  functions  of  position.  Now 
assume  that  the  ion  or  molecule  on  which  the  force  acts  has  charges  e% 
at  positions  #t,  y^  zt-,  whore  the  origin  is  chosen  to  be  at  the  center  of 
gravity  of  the  ion  or  molecule.  Then,  for  example,  the  x  component  of 
total  force  on  the  ion  or  molecule  is  the  sum  of  the  x  components  of  force 
on  all  its  charges,  and  if  we  write  Ex  at  an  arbitrary  position  by  the  Taylor 
expansion 

Ex(xyz)  =  EM  +  ^  +  d-^y  +  *jfz  +  •  •  -  ,  (1.9) 

whoro  EX(Q)  and  the  derivatives  arc  all  to  be  computed  at  the  origin,  we 
have  the  following  expression  for  the  total  x  component  of  force  on  the  ion 
or  molecule  : 

.+  •  •  •    (UO) 


The  first  term  in  Eq.  (1.10)  represents  the  field  at  the  center  of  gravity, 
times  the  total  charge.  This  term  of  course  is  zero  if  the  molecule  is 
uncharged.  The  next  three  terms  depend  on  the  dipole  moment  arid  the 
rate  of  change  of  field  strength  with  position.  Their  interpretation  is 
very  simple.  If  the  field  strength  is  independent  of  position,  the  electro- 
static forces  on  the  two  poles  of  a  dipole  will  be  equal  and  opposite  and 
will  give  no  net  force  on  the  dipole  as  a  whole.  But  if  the  field  is  stronger 
at  one  end  than  at  the  other,  one  charge  will  be  pulled  more  strongly  in 
one  direction  than  the  other  one  is  in  the  other  direction,  and  there  will 
be  a  net  pull  on  the  dipole  as  a  whole.  This  pull  depends  on  the  orienta- 
tion of  the  dipole  with  respect  to  the  external  field;  if  the  dipole  is  reversed 
in  direction,  so  that  each  component  of  its  dipole  moment  changes  sign, 
the  dipole  terms  in  the  force  expression  (1.10)  change  sign,  showing 
that  the  force  is  reversed. 

A  dipole  is  acted  on  not  only  by  a  force,  but  also  by  a  torque,  in  an 
external  field,  and  this  torque  is  proportional  to  the  field  strength  rather 
than  to  its  rate  of  change  with  position.  The  x  component  of  this  torque, 
regarded  as  a  vector,  is  seen  to  be 

Mx  =  (Se#<)^  -  (2«t*)«*  (1-11) 


360  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 

showing  that  the  torque  is  proportional  to  the  dipole  moment,  the  external 
field,  and  the  sine  of  the  angle  between  them.  To  see  this  in  an  elemen- 
tary way,  we  show  in  Fig.  XXII-2  a  simple  dipole  consisting  of  charges 
±q  at  a  distance  of  separation  d,  the  line  of  centers  making  an  angle 
0  with  the  external  field.  Then  we  see  that  the  field  exerts  a  force  of 
magnitude  qE  on  each  charge,  with  a  lever  arm  d/2  sin  0,  so  that  the 
torque  exerted  on  each  charge  is  q(d/2)E  sin  0,  and  the  total  torque  is 
qdE  sin  0,  where  qd  is  the  dipole  moment.  The  potential  energy  asso- 
ciated with  this  torque  is 

Potential  energy  =  —  qdE  cos  0,  (1.12) 

having  a  minimum  when  the  dipole  points  along  the  direction  of  the 
electric  field.  That  is,  the  field  tends  to  swing  the  dipole  around  so  that 
it  is  parallel  to  the  field. 


External  field 


Fio.  XXII-2. — Illustrating  the  torque  on  a  dipole  in  an  external  force  field. 

We  are  now  in  position  to  understand  the  forces  between  rigid  ions  or 
molecules  at  a  distance  from  each  other.  With  two  ions,  of  course  the 
largest  term  in  the  force  is  the  Coulomb  attraction  or  repulsion  between 
the  net  changes  of  the  ions — an  attraction  if  the  ions  have  unlike  charges, 
repulsion  if  they  have  like  charges.  If  the  molecules  are  uncharged, 
however,  the  largest  term  in  the  interaction  comes  from  dipole-dipole 
interaction.  Each  dipole  is  acted  on  by  a  torque  in  the  field  of  the  other, 
and  if  we  look  into  the  situation,  we  see  that  these  torques  are  in  such 
directions  as  to  tend  to  place  the  dipoles  parallel  to  each  other,  the  posi- 
tive end  of  one  being  closest  to  the  negative  end  of  the  other.  Also,  the 
dipoles  exert  a  net  force  on  each  other,  an  attraction  or  repulsion  depend- 
ing on  orientation.  If  the  orientation  is  that  of  minimum  potential 
energy,  with  the  positive  end  of  one  dipole  opposite  the  negative  end  of 
the  other,  the  net  force  will  be  an  attraction,  for  the  attraction  between 
the  close  unlike  charges  will  more  than  balance  the  repulsion  of  the  more 
distant  like  charges.  We  may  anticipate  by  mentioning  the  sort  of  appli- 
cation we  shall  make  later  to  the  force  between  two  dipole  molecules  in  a 
gao.  In  this  case,  both  dipoles  will  be  rotating.  If  they  rotated  uni- 
formly, they  would  be  pointing  in  one  direction  just  as  often  as  in  the 


SEC.  2]  INTERATOMIC  AND  1NTERMOLECULAR  FORCES  361 

opposite  direction,  so  that  the  net  force  between  them  would  cancel,  since 
as  we  have  seen  this  net  force  changes  sign  when  the  dipole  reverses  its 
direction.  But  they  will  really  not  rotate  uniformly,  for  there  are 
torques  acting  on  them,  tending  to  keep  them  in  a  parallel  position. 
These  torques  will  result  in  a  potential  energy  term,  of  the  nature  of 
Eq.  (1.12),  between  them,  and  if  we  insert  this  term  into  the  Maxwell- 
Boltzmann  distribution  law,  we  shall  find  that  the  dipoles  will  be  oriented 
in  the  position  of  minimum  potential  energy  more  often  than  in  other 
positions.  Thus,  on  the  average,  the  attractions  between  the  dipoles  will 
outweigh  the  repulsions,  and  the  net  effect  of  dipole-dipole  interaction  is 
an  intermolecular  attraction. 

As  two  molecules  or  ions  get  closer  and  closer  together,  higher  terms 
in  the  expansion  of  the  potential  and  the  force  become  important,  and  we 
must  consider  quadrupoles  and  higher  multipoles.  The  whole  expansion 
in  inverse  powers  of  r,  and  direct  powers  of  the  xl's,  becomes  badly 
convergent  when  the  molocules  approach  to  within  a  distance  com- 
parable to  their  own  dimensions.  When  the  charge  distributions  of  two 
atoms  or  molecules  really  begin  to  overlap  each  other,  the  situation 
becomes  entirely  different  and  must  be  handled  by  different  methods. 
We  shall  take  up  in  the  next  section  the  electrostatic  or  Coulomb  inter- 
action of  two  rigid  charge  distributions  representing  atoms,  when  they 
approach  so  closely  as  to  overlap. 

2.  The  Electrostatic  or  Coulomb  Interactions  between  Overlapping 
Rigid  Atoms. — We  have  seen  in  the  preceding  section  that  two  neutral 
spherically  symmetrical  atoms  exert  no  forces  on  each  other,  so  long  as 
they  do  not  overlap  and  so  long  as  we  can  treat  their  charge  distributions 
as  being  rigid,  so  that  they  do  not  distort  each  other.  Once  they  overlap, 
however,  this  conclusion  no  longer  holds.  A  rigid  neutral  atom  consists 
of  a  positive  nucleus  surrounded  by  a  spherical  negative  distribution  of 
charge,  just  great  enough  to  balance  the  charge  on  the  nucleus.  Such  a 
distribution  exerts  no  electrostatic  force  at  outside  points.  At  points 
within  the  charge  distribution,  however,  it  does  exert  a  force,  determined 
by  a  well-known  rule  of  electrostatics :  the  electrostatic  field  at  any  point 
in  a  spherical  distribution  of  charge  is  found  by  constructing  a  sphere, 
with  center  at  the  center  of  symmetry,  passing  through  the  point  where 
the  field  is  to  be  found.  The  charge  within  the  sphere  is  imagined  to  be 
concentrated  at  the  center,  that  outside  the  sphere  is  neglected.  Then 
the  electric  field  is  that  computed  by  the  inverse  square  law  from  the 
charge  concentrated  at  the  center  of  the  sphere,  disregarding  the  outside 
charge.  At  a  point  outside  the  atom,  this  reduces  to  the  same  result 
already  quoted:  the  net  charge  within  the  sphere  is  zero,  so  that  there  is  no 
field.  But  as  we  get  closer  to  the  nucleus,  we  penetrate  into  the  negative 
charge  distribution,  so  that  some  of  the  negative  charge  lies  outside 


362  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 

our  sphere  and  is  to  be  disregarded.  The  charge  within  the  sphere,  which 
we  are  to  imagine  concentrated  at  the  center,  then  has  a  net  positive  value, 
becoming  equal  to  the  charge  on  the  nucleus  as  the  sphere  grows  smaller 
and  smaller.  Thus  the  electric  field  approaches  that  of  the  positive 
nuclear  charge,  in  the  limit  of  small  distances.  It  is  correct  to  consider 
that  the  electrons  shield  the  nucleus  at  external  points,  counteracting 
its  field,  but  this  shielding  effect  decreases  as  we  penetrate  the  electron 
shells.  At  a  given  distance  from  the  nucleus,  the  field  is  like  that  of  a 
charge  of  (Z  —  Z0)  units  concentrated  at  the  center,  where  Z  is  the 
nuclear  charge,  Z0  a  shielding  constant  representing  the  amount  of  elec- 
tronic charge  within  the  sphere,  a  quantity  which  decreases  from  Z  to 
zero  as  we  go  from  great  distances  in  to  the  nucleus.  This  shielding 
constant  Z0  is  essentially  the  same  as  that  introduced  in  Eq.  (2.2),  Chap. 


(a)  (b)  (c) 

FIG.  XXII-3. — Schematic  representation  of  the  overlapping  of  two  atoms.  The 
points  represent  the  nuclei,  the  circles  the  regions  occupied  most  densely  by  negative 
electronic  charge  distributions. 

XXI,  where  we  were  considering  the  effect  of  electronic  shielding  on  the 
motion  of  one  of  the  electrons  of  the  atom. 

It  is  now  easy,  at  least  in  principle,  to  find  the  interatomic  forces 
between  two  rigid  atoms  whose  charge  distributions  penetrate  each  other. 
We  simply  find  the  force  on  each  clement  of  the  charge  of  one  atom, 
exerted  on  it  by  the  field  of  the  other.  It  is  a  difficult  problem  of  integra- 
tion actually  to  compute  this  force,  but  the  results  are  qualitatively 
simple.  Suppose  the  distributions  have  only  penetrated  slightly,  as 
shown  in  (a),  Fig.  XXII-3.  Then  some  negative  charge  of  each  atom  is 
within  the  distribution  of  the  other,  and  hence  is  attracted  by  part  of  the 
nuclear  charge.  Thus  the  first  effect  of  overlapping  is  an  attraction 
between  the  atoms.  This  effect  begins  to  be  counteracted  in  the  case 
(6)  in  the  figure,  however,  when  the  nucleus  of  one  atom  begins  to  pene- 
trate the  charge  distribution  of  the  other.  For  the  nucleus  will  be 
repelled,  not  attracted,  by  the  other  nucleus.  Finally,  in  case  (c),  where 
the  atoms  practically  coincide,  there  will  be  great  repulsion.  For  the 
nuclei  will  repel  very  strongly,  being  very  close  together,  and  exposed  to 
all  of  each  others'  field,  while  the  electronic  distribution  of  each  atom  is 
still  at  a  considerable  average  distance  from  the  nucleus  of  the  other,  and 
hence  is  not  very  strongly  attracted.  Furthermore,  part  of  the  electronic 


SEC.  3] 


INTERATOMIC  AND  INTERMOLECULAR  FORCES 


363 


distribution  of  each  atom  is  on  one  side  of  the  nucleus  of  the  other,  part 
on  the  other  side,  so  that  the  forces  on  it  almost  cancel,  and  exactly  cancel 
when  the  two  atoms  exactly  coincide.  The  net  effect  of  the  Coulomb 
forces,  then,  is  a  potential  energy  curve  similar  to  Fig.  XXII-4,  with  a 
minimum,  corresponding  to  a  position  of  equilibrium,  and  an  infinitely 
high  potential  energy  as  the  nuclei  are  brought  into  contact. 

It  might  be  thought  at  first  sight  that  the  curve  of  Fig.  XXII-4,  which 
surely  has  close  resemblance  to  the  Morse  curve  of  Fig.  IX-1,  would  give 
an  adequate  explanation  of  the  interatomic  forces  that  hold  atoms 
together  into  molecules.  On  closer  examination,  however,  this  proves 
not  to  be  the  case.  The  attractions  of  Fig.  XXII-4  are  not  nearly  strong 
enough  to  account  for  molecular  binding,  and  the  distances  of  separation 


Internuclecir  distance 


Fio.  XXII-4. — Schematic  representation  of  the  electrostatic  01  Coulomb  energy  of  inter- 
action of  two  overlapping  rigid  atoms,  as  shown  in  Fig.  XXII-3. 

are  not  what  they  should  be.  The  reason  is  that  our  assumption  of  rigid 
atoms  breaks  down  completely  when  the  electronic  distributions  begin  to 
overlap.  The  charge  distribution  becomes  greatly  distorted,  and  this 
must  be  taken  into  account  in  calculating  the  energy  and  forces.  We 
shall  now  pass  on  to  a  discussion  of  this  distortion,  first  taking  up  the 
effect  of  polarization,  the  type  of  distortion  met  at  large  distances  of 
separation,  then  the  effect  that  is  usually  called  exchange  interaction, 
which  is  important  when  atoms  overlap. 

3.  Polarization  and  Interatomic  Forces. — An  atom  or  molecule 
in  a  uniform  external  electric  field  is  polarized;  that  is,  it  acquires  an 
induced  dipole  moment,  parallel  and  proportional  to  the  field.  This  is 
the  phenomenon  so  well  known  from  electrostatics,  when  a  charge  brought 
near  a  conductor  induces  a  charge  of  opposite  sign  on  the  near-by  parts  of 
the  conductor.  It  is  not  so  marked  with  an  insulator  as  with  a  conductor, 
but  it  always  occurs.  It  is  illustrated  in  Fig.  XXII-5,  where  we  show 
simply  a  sphere  of  matter  in  an  external  field,  with  induced  positive 
charge  on  the  right  hand  part  of  it  and  negative  charge  on  the  left.  We 
see  that  the  induced  charge  is  similar  to  a  dipole.  The  induced  dipole 
moment,  as  we  have  stated,  is  proportional  to  the  external  field,  and  the 


364  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 

constant  of  proportionality  is  called  the  polarizability  and  denoted  by 
a,  so  that  the  induced  dipole  moment  is  a  times  the  external  field.  The 
polarization  can  be  brought  about  in  either  of  two  ways.  In  the  first 
place,  the  electrons  can  be  displaced  in  the  direction  opposite  to  the  field, 
so  that  the  electronic  distribution  is  distorted  or  deformed.  This  is 
the  only  mechanism  for  polarization  with  atoms  or  symmetrical  molecules. 
With  dipole  molecules,  however,  an  additional  form  of  polarization  is 
possible;  on  account  of  the  Maxwell-Boltzmann  distribution,  the  dipoles 
can  be  oriented  in  such  a  way  as  to  have  a  net  dipole  moment  along  the 
direction  of  the  field.  We  can  easily  compute  the  net  dipole  moment  on 
account  of  this  effect. 

Let  the  permanent  dipole  moment  of  a  molecule  be  /*.  Then,  as  wo 
saw  in  Eq.  (1.12),  its  potential  energy  in  a  field  E  is  —  pE  cos  0.  Its 
component  of  dipole  moment  along  the  direction  of  the  field  is  p  cos  9.  If 

all  orientations  were  equally  likely, 
the  average  component  along  the 
field  would  be  zero.  But  on  ac- 
count  of  the  Maxwoll-Boltzmann 
distribution,  the  probability  of 
finding  the  axis  of  the  dipole  in 
unit  solid  angle  about  a  given  orien- 

Fio.  XXII-5. — Induced    polarization    of    a    ...         .  , .         ,    ,         - — r»r- 

sphere  in  an  external  field.  tatlOn    IS    proportional    to    6      kT     . 

Thus  to  find  the  mean  moment,  we 

multiply  /x  cos  0  by  the  Boltzmann  factor  above  and  integrate  over  solid 
angles.  The  solid  angle  contained  between  6  and  8  +  dd,  or  the  fraction 
of  the  surface  of  unit  sphere  between  these  angles,  is  2?r  sin  0  dB.  Thus 
we  have 


fV  cos  de    kT    2?r  sin  0  dd 
Mean  dipole  moment  =  — ^  cos  0 (3.1) 

C*i~Mr~2v  sin  0  dS 
Jo 

The  integrals  in  Eq.  (3.1)  can  be  evaluated  at  once  by  substituting 
cos  0  =  x,  —  sin  6  dd  =  dx,  and  introducing  the  abbreviation  nE/kT  =  y, 
from  which  at  once  we  have 

/i 
_1XeXV  dX    ^     ey     _J_    e-y     ^      I 

_  _     '..I  ~     ^     -     ^V  »' 

J-l 

The  function  (3.2)  is  shown  as  a  function  of  y,  which  is  proportional  to 
the  external  field,  in  Fig.  XXII-6.  We  see  from  the  figure  that  at  low 
fields  the  mean  dipole  moment  is  proportional  to  the  field,  but  at  high 


SEC.  3] 


INTERATOMIC  AND  INTERMOLECULAR  FORCES 


365 


fields  there  is  a  saturation,  all  the  dipoles  being  parallel  to  the  external 
field.  It  is  only  at  low  fields,  where  there  is  proportionality,  that  we  can 
speak  of  a  polarizability.  To  get  the  value  of  the  polarizability,  we 
should  find  the  initial  slope  of  the  function  (3.2),  We  easily  find,  by 
expanding  in  power  series  in  yy  that  for  small  y  the  function  (3.2)  can  be 
approximated  by  the  straight  line  y/3.  Thus,  remembering  the  definition 
of  y,  we  have  the  dipolc  moment  at  low  fields  equal  to  »2E/3kT  and  the 
polarizability  equal  to  n*/3kT,  decreasing  with  increasing  temperature  as 
we  should  expect. 


__ 

FIG.  XX1I-6. — Function  —  - — _?/  —      as  function  of  y,  giving  mean  dipole  moment 

arising  from  the  rotation  of  dipole  molecules,  as  a  function  of  ?/  =  nE/kT,  where  /z  is  the 
dipole  moment,  K  the  field  strength,  according  to  Eq.  (3.2). 

If  the  part  of  the  polarizability  resulting  from  the  electronic  distortion 
is  ao,  we  then  have 


=  ao      3kf 


(3.3) 


as  the  total  polarizability  of  a  molecule.  This  quantity  can  be  found 
experimentally,  on  account  of  its  connection  with  the  dielectric  constant. 
The  molecular  theory  of  dielectrics  shows  that  a  substance  having  N 
molecules  in  a  volume  V,  each  having  a  polarizability  a,  will  have  a  dielec- 
tric constant  equal  to1 


N 


(3.4) 


If  we  measure  the  dielectric  constant  as  a  function  of  temperature,  then, 
it  should  be  a  linear  function  of  1/T,  and  from  the  constants  of  the  curve 
we  can  find  both  the  electronic  polarizability  ao  and  the  dipole  moment 

1  See  P.  Debye,  "Polare  Molekeln,"  Hirzel,  1929,  for  further  discussion  of  dielec- 
tric constants. 


366  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXH 

M.     It  is  in  this  way  that  the  dipole  moments  of  a  great  many  molecules 
have  been  determined. 

We  now  understand  the  polarization  of  a  molecule  in  an  external  field. 
Next,  we  must  ask  about  intermolccular  forces  resulting  from  this  polari- 
zation. If  a  molecule  has  an  induced  dipole  moment  equal  to  aE  in  the 
direction  of  the  field,  then  we  see  from  Eq.  (1.10)  that  it  is  acted  on  by  a 
force  equal  to  aE(dE/dx)j  if  the  x  axis  is  chosen  in  the  direction  of  the 
external  field  and  of  the  dipole.  This  can  be  written 


showing  that  the  force  pulls  the  molecule  in  the  direction  in  which  the 
magnitude  of  the  field  increases  most  rapidly.  In  the  type  of  problem  wo 
are  considering,  this  means  a  force  of  attraction  toward  the  other  mole- 
cule. The  attraction  will  depend  on  the  nature  of  the  field  of  the  other 
molecule.  Thus  suppose  we  are  considering  the  force  between  a  polar- 
izable  molecule  and  an  ion.  The  field  of  the  ion  varies  as  1/r2,  so  that  E~ 
varies  as  1/r4,  its  derivative  with  respect  to  x  (which  in  this  case  is  r) 
is  proportional  to  1/r5,  so  that  the  force  varies  inversely  as  the  fifth  power 
of  the  distance,  and  the  potential  energy  inversely  as  the  fourth  power. 
The  commoner  ease,  however,  is  that  in  which  a  dipole  molecule  produces 
a  field  that  polarizes  its  neighbors,  resulting  in  an  attraction.  A  molecule 
of  dipole  moment  //  produces  a  field  proportional  to  ju'/r3  at  another 
molecule.  This  field  results,  according  to  Eq.  (3.5),  in  a  force  on  the 
second  molecule  proportional  to  3a/z'2/r7.  The  attractive  energy  will 
then  be  proportional  to  1/r6.  This  depends  on  the  angle  between  the 
dipole  moment  of  the  first  molecule  and  the  line  of  centers  of  the  two,  and 
calculation  shows  that  the  average  over  all  directions  is  given  by1 

Energy  =-|^-  (3.6) 

Equation  (3.6)  is  essentially  the  formula  for  Van  der  Waals  attrac- 
tions between  molecules.  There  are  two  distinct  cases:  the  attractions 
between  molecules  with  or  without  permanent  dipole  moments.  First 
let  us  consider  molecules  without  permanent  moments.  Even  in  this 
case,  we  have  seen  in  Sec.  1  that  the  molecule  will  have  a  fluctuating 
dipole  moment,  which  will  average  to  zero.  Nevertheless  it  can  polarize 
another  molecule  instantaneously,  producing  an  attraction,  and  the  net 
result,  averaged  over  the  fluctuations,  will  be  an  attraction  given  by  Eq. 
(3.6),  where  a  is  the  electronic  polarizability  of  a  molecule,  and  /*'2  the 

1See  for  instance  Slater  and  Frank,  "  Introduction  to  Theoretical  Physics,"  Sec. 
301,  McGraw-Hill  Book  Company,  Inc.,  and  Pauling  and  Wilson,  "Introduction  to 
Quantum  Mechanics/'  Sec.  47,  McGraw-Hill  Book  Company,  Inc.,  1935. 


SEC.  4]          INTERATOMIC  AND  INTERMOLECULAR  FORCES  367 

mean  square  dipole  moment.  It  is  significant  that  it  is  the  mean  square 
moment  that  is  concerned  in  the  attraction,  and  this  mean  square  is 
different  from  zero  even  when  the  mean  moment  vanishes.  This  attrac- 
tion is  the  typical  Van  der  Waals  attraction,  a  force  whose  potential  is 
inversely  proportional  to  the  sixth  power  of  the  interatomic  distance  and 
is  independent  of  temperature.  On  the  other  hand,  if  we  are  considering 
the  forces  between  two  dipole  molecules,  there  will  be  two  changes.  First, 
the  moan  square  dipole  moment  /*'2  of  the  first  molecule  will  now  include 
two  terms :  the  one  coming  from  electronic  fluctuations,  which  we  have 
already  considered,  and  the  one  coming  from  the  fixed  average  dipole 
moment.  Thus,  in  the  first  place,  the  external  field  will  be  greater  than 
before.  Then,  in  the  second  place,  the  polarizability  will  be  given  by 
Eq.  (3.3),  including  both  the  electronic  polarizability,  and  that  coming 
from  orientation  of  the  fixed  dipoles.  In  many  cases  the  second  term 
proves  to  bo  several  timos  as  large  as  the  first,  but  it  decreases  with 
increasing  temporaturo.  Thus  we  may  expect  the  Van  der  Waals  attrac- 
tion between  molecules  with  permanent  dipoles  to  be  several  times  as 
large  as  that  between  similar  molecules  without  the  permanent  dipoles, 
and  furthermore  we  may  expect  the  Van  der  Waals  force  to  decrease  with 
temperature  in  the  dipole  case.  Both  these  predictions  prove  to  be 
borne  out  by  experiment,  as  we  shall  see  in  a  later  chapter  where  we 
take  up  Van  der  Waals  forces  numerically  for  a  variety  of  molecules. 
4.  Exchange  Interactions  between  Atoms  and  Molecules. — In  the 
preceding  section,  we  have  found  how  an  atom  or  molecule  is  distorted 
in  a  uniform  electric  field,  and  have  used  this  to  discuss  its  distortion  in 
the  field  of  another  atom  or  molecule.  This  is  clearly  an  approximation, 
for  the  field  of  a  molecule  is  not  uniform,  though  it  approaches  uniformity 
at  great  distances.  When  two  atoms  or  molecules  approach  closely,  this 
type  of  approximation,  using  merely  an  induced  dipole,  becomes  very 
inaccurate.  We  must  consider  in  this  section  how  the  charge  distribu- 
tions of  two  atoms  or  molecules  are  really  distorted  when  they  come  so 
close  together  that  they  touch  or  overlap.  We  shall  find  that  there 
are  two  very  different  types  of  behavior  possible:  there  may  be  forces  of 
attraction  between  the  atoms  or  molecules,  tending  to  bind  them  together, 
or  there  may  be  forces  of  repulsion.  The  first  case  is  that  of  valence 
binding,  the  second  the  case  of  the  type  of  repulsion  considered  in  Van  der 
Waals  constant  6.  We  shall  first  take  up  the  simplest  case,  the  interac- 
tion of  two  atoms,  then  shall  pass  on  to  molecular  interactions.  The 
problems  we  are  now  meeting  are  among  the  most  complicated  ones  of  the 
quantum  theory,  and  we  shall  make  no  attempt  at  all  to  treat  the  ana- 
lytical background  of  the  theory.  When  one  studies  that  background, 
one  finds  that  there  are  two  different  approximate  methods  of  calcu- 
lation used  in  wave  mechanics,  sometimes  called  the  Heitler-London 


368 


INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 


method  and  the  method  of  molecular  orbitals  respectively.  These  differ, 
not  in  their  fundamentals,  but  in  the  precise  nature  of  the  analytical  steps 
used.  For  that  reason,  we  shall  not  discuss  them  or  their  differences. 
We  shall  rather  try  to  focus  our  attention  on  the  fundamental  physical 
processes  behind  the  intermolecular  actions  and  shall  find  that  we  can 
understand  them  in  terms  of  fundamental  principles,  without  reference  to 
exact  methods  of  calculation. 

Let  us,  then,  bogin  with  the  simplest  possible  problem,  the  interaction 
of  two  atoms.  Unless  they  overlap,  the  only  force  between  them  will  be 
the  Van  der  Waals  attraction,  coming  from  the  polarization  of  each 
atom  by  the  fluctuating  dipole  moment  of  the  other.  This  type  of  inter- 
action persists  even  when  the  valence  electrons  of  the  atoms  do  overlap. 


Po  ten  fior/  energy 


(a)  (b) 

FIG.  XXII-7. — Potential  energy  of  an  electron  111   (a)   the  central  field  representing  an 
atom;  (b)  the  field  representing  two  overlapping  atoms. 

It  is  simple  to  describe  in  words.  When  the  valence  electron  of  one  atom 
is  at  a  given  point,  the  valence  electron  of  the  other  atom,  which  of  course 
is  repelled  by  it,  tends  to  stay  away  from  it.  Thus  the  electrons  do  not 
approach  each  other  so  closely  as  if  the  repulsion  were  absent,  and  as  a 
result  the  interaction  energy  between  them  is  lower  than  if  we  neglected 
this  type  of  interaction.  This  effect,  tending  to  keep  the  electrons  in  the 
pair  of  atoms,  or  the  molecule,  away  from  each  other,  is  sometimes  called 
a  correlation  effect,  since  it  depends  on  a  correlation  between  the  motions 
of  the  two  electrons.  It  results  in  a  lowering  of  energy  or  an  increase  of 
the  strength  of  the  binding  between  the  atoms.  As  we  see,  it  is  the  direct 
extrapolation  of  the  Van  der  Waals  attraction  to  the  case  of  close  approach 
of  the  atoms.  But  as  we  shall  soon  see,  it  is  by  no  means  the  principal 
part  of  the  interatomic  force  but  rather  forms  a  fairly  small  correction 
term. 

In  discussing  the  Coulomb  interactions  between  overlapping  atoms, 
in  Sec.  2,  we  saw  that  as  the  electronic  charge  of  one  atom  begins  to 
penetrate  into  the  electron  shells  of  the  other,  it  becomes  attracted  to 
the  nucleus  of  the  other  atom.  That  is,  it  is  in  a  region  of  lower  potential 
energy  than  it  otherwise  would  be.  This  is  illustrated  in  Fig.  XXII-7. 
There,  in  part  (a),  we  show  the  potential  energy  of  an  electron  at  different 


SBC.  4]          INTERATOMIC  AND  INTERMOLECULAR  FORCES  369 

points  within  an  atom,  taking  account  of  the  decrease  of  shielding  as  we 
go  closer  to  the  nucleus.  In  (fc),  the  potential  curves  of  two  overlapping 
atoms  are  superposed,  the  resulting  curve  showing  the  potential  energy  of 
an  electron  in  the  combined  molecule.  We  see  that  the  potential  energy 
is  lower  in  the  region  where  they  overlap  than  it  would  be  in  the  cor- 
responding part  of  either  atom  separately.  This  change  of  potential 
energy  would  mean  that  the  electrons  from  both  atoms  were  attracted 
to  this  region  of  overlapping,  so  that  the  tendency  would  be  for  extra 
electronic  charge  to  concentrate  itself  in  this  region.  This  in  turn  would 
decrease  the  total  energy,  for  it  would  mean  the  concentration  of  more 
charge  in  a  region  of  lower  potential  energy.  Thus  this  would  result  in 
an  added  attraction  between  the  atoms.  We  could  regard  this  as  an 
increase  in  the  magnitude  of  the  Coulomb  attraction,  if  we  chose,  giving  a 
much  deeper  minimum  to  the  potential  energy  curve  than  one  finds  in 
Fig.  XXII-4.  This  effect  is  different  from  the  Van  dor  Waals  attraction 
or  correlation  effect,  in  that  it  depends  on  the  average  field  rather  than  on 
the  fluctuating  field,  distorting  or  polarizing  the  charge  distribution  and 
hence  decreasing  the  energy. 

Even  this  effect,  however,  is  not  the  whole  story.  For  we  have 
forgotten  one  essential  fact :  the  electrons  obey  the  Fermi-Dirac  statistics 
or  the  Pauli  exclusion  principle.  Let  us  state  the  Fcrmi-Dirac  statistics 
in  a  very  simple  form,  remembering  the  existence  of  the  electron  spin. 
We  set  up  a  molecular,  or  rather  an  electronic,  phase  space,  in  which 
the  coordinates  of  each  electron  are  given  by  a  point.  This  phase  space 
has  cells  of  volume  A3  in  it.  Then  the  Fermi-Dirac  statistics  states  that 
no  complexion  of  the  system  is  possible  in  which  more  than  one  electron, 
of  a  given  spin  orientation,  is  in  the  same  cell.  Since  two  orientations  of 
the  spin  are  possible,  as  we  saw  in  Chap.  XXI,  Sec.  2,  this  means  that  at 
most  we  can  have  two  electrons  in  a  cell,  one  of  each  spin.  There  is,  in 
other  words,  a  maximum  possible  density  of  electrons  in  phase  space. 
We  can  translate  this  statement  into  one  regarding  the  maximum  density 
of  electrons  in  ordinary  coordinate  space :  for  a  given  range  of  momenta, 
there  is  a  maximum  possible  density  of  electrons  in  coordinate  space.  If 
we  wish  to  pack  more  electrons  into  a  region  of  coordinate  space,  they 
must  have  a  different  momentum  from  those  already  there.  If  the  elec- 
trons already  present  have  a  low  kinetic  energy,  this  means  that  any 
additional  electrons  must  have  higher  kinetic  energy,  and  hence  higher 
total  energy,  in  order  not  to  have  the  same  momenta  as  those  already 
present.  The  exclusion  principle  in  this  form  can  be  used  immediately 
to  discuss  the  electronic  interactions  when  two  atoms  begin  to  overlap. 

We  have  already  talked  about  the  lowering  of  the  potential  energy 
between  two  atoms  when  they  begin  to  overlap,  and  have  illustrated  it  in 
Fig.  XXII-7.  And  we  have  stated  that  there  is  a  tendency  for  electrons 


370  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXII 

to  be  concentrated  in  this  region,  seeking  a  lower  potential  energy  and 
hence  decreasing  the  energy  of  the  whole  system.  But  this  would  involve 
an  increasing  density  of  electrons  in  the  region  between  the  atoms,  and 
from  what  we  have  just  said,  this  might  well  make  difficulties  with  the 
Fermi-Dirac  statistics.  We  now  meet  very  different  situations,  depend- 
ing on  whether  the  electrons  in  the  two  atoms  already  are  in  closed  shells, 
or  not.  First  lot  us  assume  that  they  are  in  closed  shells.  This  can  be 
interpreted  in  terms  of  the  Fermi-Dirac  statistics:  we  choose  our  cells  in 
the  electronic  phase  space  to  coincide  with  the  stationary  states  of  the 
one-electron  problem  of  an  electron  in  a  central  field.  When  we  state  that 
the  electrons  are  all  in  closed  shells,  we  mean  that  there  are  two  electrons 
each,  one  of  each  spin,  in  the  lowest  cells  or  stationary  states,  while  the 
higher  stationary  states  are  empty.  All  the  region  of  space  occupied 
by  electrons  of  either  atom,  then,  is  filled  to  such  a  density  with  electrons 
that  no  additional  charge  can  enter  the  region,  without  having  a  higher 
kinetic  energy  and  total  energy,  than  the  charge  already  there.  Now 
let  us  see  how  this  affects  the  situation.  If  two  atoms  begin  to  overlap,  we 
can  certainly  not  have  the  charge  shifting  into  the  region  between  the 
atoms,  for  this  would  involve  an  increase  of  charge  density.  We  can  not 
even  have  the  charge  of  the  two  atoms  overlap  without  redistribution,  for 
the  same  reason.  For  this  would  involve  such  a  large  density  of  charge 
between  the  atoms  that  the  electrons  would  have  to  increase  their  kinetic 
energy,  and  hence  their  total  energy,  considerably.  The  thing  that  hap- 
pens is  that  some  of  the  charge  actually  shifts  away  from  the  region 
between  the  atoms,  to  the  far  sides  of  the  nuclei.  This  involves  some 
increase  of  kinetic  energy,  for  the  electrons  must  increase  the  charge 
density  everywhere  except  between  the  atoms,  to  make  up  for  the  decrease 
of  density  there;  it  involves  increase  of  potential  energy,  since  electrons 
are  moving  away  from  the  region  of  low  potential  energy  between  the 
nuclei.  Nevertheless,  it  does  not  mean  so  much  increase  of  total  energy 
as  if  the  electrons  piled  up  between  the  atoms.  The  net  effect  of  this 
redistribution  of  charge  is  an  increase  of  energy  and  hence  a  repulsion 
between  the  atoms. 

The  effect  of  which  we  have  spoken,  giving  a  repulsion  between  atoms 
all  of  whose  electrons  are  in  closed  shells,  is  the  origin  of  the  impenetra- 
bility of  atoms  and  of  the  correction  to  the  perfect  gas  law  made  by  Van 
der  Waals  constant  b.  It  is  illustrated  in  Fig.  XXII-8,  where  in  (a)  we 
show  the  charge  distribution  surrounding  two  repelling  atoms,  by  means 
of  contour  lines.  It  is  clear  that  the  charge  has  been  forced  out  of  the 
region  between  atoms,  by  the  effect  of  the  exclusion  principle.  As  a 
matter  of  fact,  we  can  get  similar  effects,  even  if  the  outer  electrons  are 
not  all  in  closed  shells.  Thus,  consider  two  atoms  of  hydrogen,  or  of  an 
alkali  metal,  each  with  one  valence  electron.  Suppose  the  electrons  of 


SEC.  4] 


INTERATOMIC  AND  1NTERMOLECULAR  FOHCKti 


371 


both  atoms  have  their  spins  oriented  in  the  same  way.  Then  as  far  as 
electrons  of  that  spin  are  concerned,  the  shells  are  filled,  although  they  are 
empty  of  electrons  of  the  opposite  spin.  When  the  electrons  begin  to 
overlap,  then,  there  is  the  same  difficulty  about  an  increase  of  charge 


(a)  (b) 

FIG.  XXII-8. — Electronic  charge  density  represented  by  contours*,   foi    (a)   two  repelling 
atoms,  (b)  two  attracting  atoms. 

density  that  there  would  be  with  really  closed  shells,  the  charge  distribu- 
tion becomes  distorted  as  in  Fig.  XXII-8  (a)  and  the  atoms  repel.  There 
really  is  a  mode  of  interaction  of  two  hydrogen  or  two  alkaline  atoms 
leading  to  a  repulsion  of  this  sort.  We  show  it  graphically  in  Fig.  XXII-9 
(a).  In  this  figure,  we  plot  the  total  energy  of  a  pair  of  hydrogen  atoms, 


Interatomic 
distance 


FIG.  XXII-9. — Interaction  energy  of  two  hydrogen  atoms,  as  a  function  of  the  distance  of 
separation,     (a)  repulsive  state,  (b)  attractive  state,  with  molecular  formation. 

as  a  function  of  the  distance  of  separation.  At  large  distances,  the 
energy  is  negative,  on  account  of  the  Van  der  Waals  attraction.  But  at 
small  distances,  when  the  atoms  overlap,  the  curve  (a),  indicating  the 
case  where  the  spins  of  the  two  electrons  are  parallel,  gives  a  repulsion. 
There  is  a  minimum  in  this  curve,  leading  to  a  position  of  equilibrium 
between  the  atoms,  but  it  corresponds  to  a  large  interatomic  distance  and 


372  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XX11 

very  small  binding  energy,  and  does  not  correspond  in  any  way  to  the 
binding  of  the  two  atoms  to  form  a  molecule. 

The  cases  which  we  have  taken  up  so  far  are  those  of  repulsion  between 
atoms,  resulting  from  the  exclusion  principle.  But  this  principle  can  also 
operate,  in  a  somewhat  less  obvious  way,  to  give  an  attraction  between 
atoms  which  is  even  greater  than  the  other  forms  of  attraction  previously 
considered.  In  the  case  of  two  hydrogen  or  alkaline  atoms,  which  we 
have  just  discussed,  it  may  be  that  the  two  electrons  will  have  opposite 
spins.  Then  the  exclusion  principle  does  not  operate  directly;  there  is  no 
obstacle  in  the  way  of  the  electrons  from  the  two  atoms  overlapping,  as 
much  as  they  please,  since  their  spins  are  different.  The  electrons  are 
then  free  to  pile  up  in  the  region  between  the  nuclei,  as  we  have  mentioned 
before,  thus  decreasing  the  potential  energy  and  leading  to  a  binding 
between  the  atoms.  But  the  exclusion  principle,  or  Fermi-Dirac  statis- 
tics, comes  into  wave  mechanics  in  a  more  fundamental  way  than  we  have 
indicated,  a  way  that  can  hardly  be  explained  at  all  without  going  much 
further  into  wave  mechanics  than  we  can.  The  effect  in  this  caso  is 
something  of  the  following  sort:  if  two  electrons  have  the  same  spin,  as  we 
have  seen,  the  exclusion  principle  prevents  them  from  being  in  the  same 
cell  of  phase  space.  But  if  they  have  opposite  spins,  it  operates  just  in 
the  opposite  direction,  making  electrons  tend  to  occupy  the  same  cell 
rather  than  different  cells.  A  hint  as  to  why  this  should  be  so  can  be 
found  from  Chap.  V,  Sec.  6,  where  we  discussed  the  Einstein-Bose  statis- 
tics. We  remember  that  that  form  of  statistics  resulted  merely  from 
the  identity  of  molecules,  and  that  it  could  be  qualitatively  described  as  a 
tendency  for  molecules  to  stick  together  or  condense,  as  if  there  wore 
attractive  forces  between  them.  If  the  exclusion  principle  is  added  to 
the  principle  of  identity,  the  Fermi-Dirac  statistics  results,  leading  to 
something  like  a  repulsion  between  electrons.  Now  two  electrons  of  the 
same  spin  must  obey  the  exclusion  principle,  and  we  have  already  seen 
the  effect  of  the  resulting  repulsion.  But  two  electrons  of  the  opposite 
spin  no  longer  need  to  satisfy  an  exclusion  principle,  as  far  as  their  coor- 
dinates and  momenta  are  concerned,  and  yet  they  still  satisfy  a  principle 
of  identity.  Hence,  in  essence,  they  obey  the  Einstein-Bose  statistics 
and  tend  to  crowd  together  as  closely  as  possible. 

The  effect  of  which  we  have  just  spoken  is  often  called  exchange,  on 
account  of  a  feature  in  the  analytical  calculation  connected  with  it,  in 
which  the  essential  term  relates  to  an  exchange  of  electrons  between  the 
two  atoms.  The  exchange  interaction  results  in  an  additional  piling  up  of 
electrons  in  the  region  of  lowest  potential  energy,  between  the  nuclei,  and 
hence  in  an  addition  to  the  strength  of  binding.  This  is  indicated  in 
Fig.  XXII-8  (6),  where  the  charge  distribution  for  this  case  of  attraction 
is  shown,  and  in  Fig.  XXII-9  (6),  where  we  show  the  potential  energy  of 


SBC.  4]  INTERATOMIC  AND  INTERMOLECULAR  FORCES  373 

this  type  of  interaction.  The  type  of  attraction  which  we  have  in  this 
case  is  what  is  generally  called  homopolar  valence  attraction,  the  word 
"homopolar"  meaning  that  the  two  atoms  in  question  have  the  same 
polarity,  rather  than  one  being  electropositive  and  one  electronegative,  as 
in  attraction  between  a  positive  and  a  negative  ion.  We  can  now  see  the 
various  features  involved  in  it:  there  is  a  tendency,  from  pure  electro- 
statics, for  the  outer  or  valence  electrons  of  the  atoms  to  concentrate  in 
the  region  between  the  atoms;  if  the  electrons  have  the  same  spin  this  is 
prevented  by  the  exclusion  principle,  but  if  they  have  opposite  spin  the 
exclusion  principle  indirectly  operates  to  enhance  the  concentration  of 
charge;  since  this  charge  is  concentrated  in  a  region  of  low  potential 
energy,  the  net  result  is  a  binding  of  the  two  atoms  together;  and  finally, 
the  correlation  effect,  analogous  to  the  Van  der  Waals  attraction,  tends  to 
keep  the  electrons  out  of  each  others'  way,  still  further  decreasing  the 
energy.  All  those  effects  result  in  interatomic  attraction  at  moderate 
distances  of  separation.  As  the  distance  is  further  decreased,  however, 
two  effects  tend  to  produce  repulsion.  First,  there  is  the  simple  effect  of 
the  Coulomb  forces,  discussed  in  Sec.  2:  as  the  nucleus  of  one  atom 
begins  to  penetrate  inside  the  charge  distribution  of  the  other,  the  nuclei 
begin  to  repel  each  other,  the  repulsion  growing  stronger  as  they  approach. 
But  secondly,  all  atoms  but  the  simplest  ones  have  inner,  closed  shells. 
It  is  only  the  outer  electrons,  which  are  not  in  closed  shells,  that  can 
take  part  in  valence  attraction.  When  the  atoms  come  close  enough 
together  so  that  the  closed  shells  begin  to  overlap  each  other,  the  same 
sort  of  repulsion  produced  by  the  exclusion  principle  sets  in  which  we 
have  previously  mentioned.  In  many  cases  this  repulsion  of  closed  shells 
is  the  major  feature  in  producing  the  rise  of  the  potential  energy  curve 
which  is  shown  in  Fig.  XXII-9  (fe). 

We  have  spoken  of  the  effect  of  the  exclusion  principle  and  exchange 
on  the  forces  between  two  atoms.  Now  we  shall  see  what  happens  with 
more  than  two  atoms.  For  the  sake  of  illustration,  let  two  hydrogen 
atoms  be  bound  into  a  molecule  by  the  action  of  exchange  forces,  and 
then  ask  what  happens  when  a  third  hydrogen  atom  approaches  the 
molecule.  The  two  electrons  of  the  first  two  atoms  have  cooperated  to 
form  the  valence  bond  between  them.  One  of  these  electrons  has  one 
spin,  the  other  the  other,  and  they  have  shifted  into  the  region  between 
the  atoms,  filling  that  region  up  to  approximately  the  maximum  density 
allowed  by  the  exclusion  principle,  with  electrons  of  both  spins.  Such  a 
pair  of  electrons,  shared  between  two  atoms,  is  the  picture  furnished 
by  the  quantum  theory  for  the  electron-pair  bond.  But  now  imagine  a 
third  atom,  with  its  electron,  to  approach.  The  spin  of  this  third  elec- 
tron is  bound  to  be  the  same  as  that  of  one  of  the  two  electrons  of  the 
electron-pair  bond.  Thus  this  third  atom  cannot  enter  into  the  attrac- 


374  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXII 

tive,  exchange  type  of  interaction  with  either  of  the  atoms  bound  into 
the  molecule.  Instead,  the  exclusion  principle  will  force  its  electron 
away  from  the  other  atoms,  and  there  will  be  repulsion  between  them. 
This  effect  is  what  is  called  saturation  of  valence:  two  electrons,  and  no 
more,  can  enter  into  an  electron-pair  bond,  and  once  such  a  bond  is 
formed,  the  electrons  concerned  in  it  can  form  no  more  bonds.  It  might 
have  been,  however,  that  one  of  the  atoms  concerned  in  the  original 
molecule  had  two  valence  electrons  which  it  could  share.  In  that  case, 
one  of  them  would  be  used  up  in  forming  a  valence  bond  with  the  first 
hydrogen  atom,  leaving  the  other  one  to  form  a  bond  with  another  hydro- 
gen atom.  In  this  way,  we  can  have  atoms  capable  of  forming  two  or 
more  valence  bonds.  If  all  the  possible  bonds  are  already  formed,  how- 
over,  the  structure  will  act  as  if  all  its  electrons  were  in  closed  shells,  and 
any  additional  atom  or  molecule  approaching  it  will  be  repelled. 

There  is  just  one  case  in  which  the  formation  of  a  bond  by  one  electron 
does  not  prevent  the  same  electron  from  taking  part  in  another  bond. 
This  is  the  case  of  the  metallic  bond.  If  two  sodium  atoms  approach, 
with  opposite  spins,  their  electrons  form  a  valence  bond  botwoon 
them.  But  if  a  third  sodium  atom  approaches,  it  turns  out  that  the  first 
valence  bond  becomes  partly  broken,  so  that  the  valence  electrons  of  the 
first  two  atoms  spend  only  part  of  their  time  in  the  region  between  those 
two  atoms  and  have  part  of  their  time  left  over  to  form  bonds  with  the 
new  atom.  As  more  and  more  atoms  are  added,  this  effect  can  continue, 
the  electrons  forming  bonds  which  are  essentially  homopolar  in  nature 
but  spread  out  throughout  the  whole  group  of  atoms,  holding  them  all 
together.  The  reason  why  metallic  atoms  behave  in  this  way,  while 
nonmetallic  ones  do  not,  is  probably  largely  the  fact  that  the  valence 
electrons  of  metals  are  less  tightly  held  than  in  non metals,  as  we  can  see  in 
Table  XXI-3,  giving  the  ionization  potentials  of  the  elements,  and  con- 
sequently their  orbits  are  larger,  as  we  see  in  Table  XXI-4.  Then  the 
orbit  of  one  atom  overlaps  other  atoms  more  than  in  a  nonmetal,  and  it  is 
easier  for  a  number  of  neighbors  to  share  in  valence  attraction.  The 
conspicuous  features  of  the  metallic  bond  are  two:  first,  there  is  no  satura- 
tion of  valence,  so  that  any  number  of  atoms  can  be  held  together,  forming 
a  crystal  rather  than  a  finite  molecule;  and  secondly,  the  electron  density 
is  not  so  great  as  the  maximum  allowed  by  the  exclusion  principle.  This 
second  fact  makes  it  possible  for  electrons  to  move  from  point  to  point 
without  significant  increase  in  their  energy,  whereas  in  a  molecule  held 
by  valence  bonds  this  is  impossible,  since  the  electron  would  have  to 
acquire  enough  extra  energy  to  rise  to  a  higher  quantum  state,  or  more 
excited  cell  in  phase  space,  before  it  could  enter  regions  which  already  had 
their  maximum  density  of  electrons.  This  free  motion  of  the  electrons 
in  a  metal  is  what  leads  to  its  electrical  conductivity  and  its  typical 
metallic  properties. 


SBC.  5]          INTERATOMIC  AND  1NTERMOLECULAR  FORCES  375 

6.  Types  of  Chemical  Substances. — In  the  preceding  sections  we 
have  made  a  survey  of  the  types  of  interatomic  and  intermolecular 
forces.  Now  we  can  correlate  these  by  making  a  brief  catalog  of  the 
important  types  of  chemical  substances  and  the  sorts  of  forces  found  in 
each  case.  Following  this,  the  remaining  chapters  of  this  book  will  take 
up  each  type  of  substance  in  more  detail,  making  both  qualitative  and 
quantitative  use  of  the  laws  of  interatomic  and  intermolecular  forces 
found  in  each  particular  case,  and  deriving  the  physical  properties  of  the 
substances  as  far  as  possible  from  the  laws  of  force. 

The  simplest  class  of  anhst,fl.m*Psr  in  ^  way,  is  the  class  of  inorpjamn 
salts,  or  ionic  crystals.  J?amiliar  examples  are  NaCl,  NaNO3,  BaS04. 
These  substances  are  definitely  constructed  of  ions,  as  Na+,  Ba++,  Cl~, 
NOa~,  and  SO4 — .  The  ions  act  on  each 'other  by  Coulomb  attraction 
or  repulsion,  and  an  ion  of  one  sign  is  always  surrounded  more  closely  by 
neighbors  of  the  opposite  sign  than  by  others  of  the  same  sign,  so  that  the 
attractions  outweigh  the  repulsions  and  the  structure  holds  together. 
As  the  distance  between  ions  decreases  and  they  begin  to  touch  each  other, 
they  repel,  on  account  of  the  exclusion  principle;  for  as  we  have  seen,  the 
electrons  in  an  ion  form  closed  shells  and  the  repulsion  between  them  is 
the  typical  repulsion  of  closed  shells.  Thus  a  stable  structure  can  be 
formed,  the  electrostatic  attractions  balancing  the  repulsions.  There  is 
nothing  of  the  nature  of  saturation  of  valence;  even  though  two  ions,  as 
Na+  and  Cl~,  may  be  bound  into  a  molecule,  the  electrostatic  effect  of 
their  charges  extends  far  away  from  them,  since  the  molecule  NaCl  has  a 
strong  dipole  moment.  Thus  further  ions  can  be  attracted,  and  the 
tendency  is  to  form  an  extended  structure.  If  the  atoms  in  this  structure 
are  arranged  regularly,  it  is  a  crystalline  solid,  the  most  characteristic 
form  for  ionic  substances.  On  the  other  hand,  at  higher  temperatures, 
where  there  is  more  irregularity,  the  solid  can  liquefy,  and  at  high  enough 
temperatures  it  vaporizes.  It  is  only  in  the  vapor  state  that  we  can  say 
that  the  substance  is  composed  of  molecules;  in  the  liquid  or  solid,  each 
ion  is  surrounded  at  equal  distances  by  a  number  of  ions  of  the  opposite 
sign  and  there  is  no  tendency  of  the  ions  to  pair  off  to  form  molecules. 
The  electrostatic  attractions  met  in  ionic  crystals  are  large,  so  that  the 
materials  are  held  together  strongly  with  rather  high  melting  and  boiling 
points.  We  shall  see  in  a  later  chapter  that  we  can  account  for  their 
properties  satisfactorily  by  quite  simple  mathematical  methods. 

The  ionic  substances  are  those  in  which  an  clectroposjtfivfi  ftlftmftnf,,  »« 
Na.  and  an  electronegative  one?  as  Cl,  are  held  together  by  electrostatic 
forces.  The  other  types  of  substances  are  compounds  of  two  electro- 
negative  elements  or  of  two  electropositive  ones.  The  first  group,  made 
of  electronegative  elements,  is  the  group  of  homopolar  compounds. 
These  are  held  together  by  homopolar  valence  bonds,  coming  from  shared 
electron  pairs,  as  we  have  described  in  the  previous  section.  Since  the 


376  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXII 

bonds  have  the  property  of  saturation,  we  ordinarily  have  molecular 
formation  in  such  cases,  the  molecules  being  of  quite  definitely  deter- 
mined size.  Two  molecules  attract  each  other  only  by  Van  der  Waals 
forces,  and  if  they  are  brought  too  closely  into  contact,  they  repel  each 
other  on  account  of  their  closed  shells.  The  simpler  compounds  of  this 
type,  like  H2,  62,  CO,  etc.,  are  the  materials  most  familiar  as  gases.  The 
Van  der  Waals  forces  holding  the  molecules  together  are  rather  weak, 
while  the  interatomic  forces  holding  the  atoms  together  in  the  molecule 
are  very  strong.  Thus  a  relatively  low  temperature  suffices  to  pull  the 
molecules  apart  from  each  other,  or  to  vaporize  the  liquid  or  solid,  while 
an  extremely  high  temperature  is  necessary  to  dissociate  the  molecule  or 
pull  its  atoms  apart.  As  we  go  to  more  complicated  cases  of  compounds 
held  together  by  homopolar  valence,  we  first  meet  the  organic  compounds. 
They  arise  on  account  of  the  tendency  of  the  carbon  atom  to  hold  four 
other  atoms  by  valence  bonds,  using  its  four  available  electrons  in  this 
way.  The  carbon  atoms,  on  account  of  their  many  valences,  can  form 
chains  and  still  have  other  valence  bonds  available  for  fastening  other 
atoms  to  them;  in  this  way  the  great  variety  of  organic  compounds  is 
built  up.  Another  very  important  class  of  materials  held  together  by 
valence  bonds  contains  the  minerals,  silicates,  and  glasses,  and  various 
refractory  materials,  like  diamond,  carborundum,  or  SiC,  and  so  on.  In 
these  cases,  the  valence  bonds  hold  the  atoms  into  endless  chains,  sheets, 
or  three-dimensional  structures,  so  that  the  materials  form  crystals  in 
their  most  characteristic  form  and  are  held  together  very  tightly.  These 
materials  have  very  high  melting  and  boiling  points,  since  all  the  bonds 
between  atoms  are  the  very  strong  valence  bonds,  rather  than  the  weak 
Van  der  Waal$  f^rnes  as  jp  the  iflolecniar  substances.  _ 

Finally  we  have  the  metals,  made  entirely  of  electropositive  atoms. 
We  *"""»  "»-»"  that  these  atoms  are  hold  together  by  the  metallic  bond, 
similar  to  the  valence  hnnHa,  W.  without,  the  propqrftes  of  saturation. 
Thus  the  metals,  like  the  ionic  crystals  and  the  silicates,  tend  to  form 
indefinitely  large  structures,  crystals  or  liquids,  arid  tend  to  have  high 
melting  and  boiling  points  and  great  mechanical  strength.  We  have 
already  seen  that  the  same  peculiarity  of  the  metallic  bond  which  prevents 
the  saturation  of  valence,  and  hence  which  makes  crystal  formation 
possible,  also  leads  to  metallic  conduction  or  the  existence  of  free  electrons. 

With  this  brief  summary,  we  have  covered  most  of  the  important 
types  of  materials.  In  the  next  chapter  we  shall  make  a  detailed  study  of 
ionic  substances,  and  in  succeeding  chapters  of  the  various  other  sorts  of 
materials,  interpreting  their  properties  in  terms  of  interatomic  and  inter- 
molecular  forces. 


CHAPTP:R  xxm 

IONIC  CRYSTALS 

The  ionic  compounds  practically  all  form  crystalline  solids  at  ordinary 
temperatures,  and  it  is  in  this  form  that  they  have  been  most  extensively 
studied.  The  reason  for  this  crystal  formation  has  been  seen  in  the  pre- 
ceding chapter:  ionic  attractions  are  long  range  forces,  falling  off  only  as 
the  inverse  square  of  the  distance,  so  that  more  and  more  ions  tend  to  be 
attracted  to  a  minute  crystal  which  has  started  to  form,  and  the  crystal 
grows  to  large  size,  held  together  by  electrostatic  attractions  throughout 
its  whole  volume.  Above  the  melting  point,  the  liquids  of  course  are 
held  together  by  the  same  type  of  force,  and  in  the  gaseous  phase  undoubt- 
edly the  same  sort  of  thing  occurs,  one  molecule  tending  to  attract  others, 
>so  that  presumably  there  is  a  strong  tendency  for  the  formation  of  double 
and  multiple  molecules  in  the  gas.  Unfortunately,  however,  the  liquids 
and  gases  of  ionic  materials  have  been  greatly  neglected  experimentally, 
so  that  there  is  almost  no  empirical  information  with  which  to  compare 
any  theoretical  deductions.  For  this  reason,  we  shall  be  concerned  in 
this  chapter  entirely  with  the  solid,  crystalline  phase* of  these  substances, 
but  venture  to  express  the  suggestion  that  further  experimental  study  of 
the  liquid  and  gaseous  phases  would  be  very  desirable.  In  considering 
the  solids,  the  first  step  naturally  is  to  examine  the  geometrical  arrange- 
ment of  the  atoms  or  the  crystal  structure.  Then  we  shall  go  on  to  the 
forces  between  ions,  and  the  mechanical  and  thermal  properties,  first 
taking  up  the  case  of  the  behavior  at  the  absolute  zero,  then  studying 
temperature  vibrations  and  thermal  effects. 

1.  Structure  of  Simple  Binary  Ionic  Compounds. — To  be  electrically 
neutral,  every  ionic  compound  must  contain  some  positive  and  some 
negative  ions.  The  simplest  ones  are  those  binary  compounds  that 
contain  one  positive  and  one  negative  ion.  Obviously  the  positive  ion 
must  have  lost  the  same  number  of  electrons  that  the  negative  one  has 
gained.  Thus  monovalent  metals  form  such  compounds  with  monova- 
lent  bases,  divalent  with  divalent,  and  so  on.  In  other  words,  this  group 
includes  compounds  of  Li+,  Na+,  K+ ,  Rb+,  Cs+,  Cu+,  Ag+,  and  Au+  with 
F-,  C1-,  Br-,  I-;  of  Be++  Mg++  Ca++,  Sr++  Ba++,  Zn++,  Cd++  Hg++ 
with  0—  S—  Se— ,  Te~;  of  B+++  A1+++  Ga+++,  In+++  T1+++  with 

N — ,  P ,  As ,  Sb ,  Bi      .     Even  this  list  contains  some  negative 

ions  which  ordinarily  do  not  really  exist,  as  As      ,  Sb      ,  Bi      .     Never- 
theless some  of  the  compounds  in  question  are  found.     One  can  formally 

377 


378  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 

go  even  further  and  set  up  such  compounds  as  carborundum,  SiC,  as  if 

it  were  made  of  81++++  and  C ,  or  C++++  and  Si .     But  by  the 

time  an  atom  has  gained  or  lost  so  many  electrons,  it  turns  out  that  the 
ionic  description  does  not  really  apply  very  well,  and  we  shall  see  later 
that  such  compounds  are  really  better  described  as  homopolar  compounds. 
The  positive  ions  which  we  have  listed  above  by  no  means  exhaust  the 
list  of  possibilities,  on  account  of  the  fact  that  most  of  the  elements  of 
the  iron,  palladium,  and  platinum  groups  are  found  as  divalent,  or  tri- 
valent  positive  ions,  and  consequently  form  binary  oxides,  sulphides, 
selenides,  and  tellurides  in  their  divalent  form,  and  nitrides  and  phos- 
phides in  their  trivalent  form. 

In  addition  to  these  single  ions, 
there  .are  a  few  complex  positive 
ions,  which  are  so  much  like  metallic 
ions  that  they  can  conveniently  be 
grouped  with  them.     Best  known  of 
these  is  the  ammonium  ion,  NH4+, 
and  somewhat  loss  familiar  is  the 
analogous  phosphonium  ion  PH4+. 
Tic*   The  ammonium  ion  has  ten  elec- 
trons:  seven   from   nitrogen,    one 
from  each  of  the  four  hydrogens, 
FIG.  xxiil-i.— The  sodium  chloride        iess  one  because  it  is  a  positive  ion. 
structure.  n^     ,   .      .,  ,         .      ,    ,, 

lhat  is,  it  has  just  the  same  num- 
ber as  neon,  or  as  Na+.  Similarly  the  phosphonium  ion  has  eighteen, 
like  argon,  or  K+.  The  hydrogen  ions  are  presumably  imbedded  in  the 
distribution  of  negative  charge,  in  a  symmetrical  tetrahedral  arrangement, 
and  do  not  greatly  affect  the  structure,  so  that  these  ions  act  surprisingly 
like  metallic  ions.  We  shall  group  these  compounds  with  the  binary  ones, 
though  really  the  positive  ion  is  complex. 

Most  of  the  binary  ionic  compounds  occur  in  one  of  four  structures, 
and  by  far  the  commonest  is  the  sodium  chloride  structure.  This  is 
shown  in  Fig.  XXIII-1.  It  can  be  described  as  a  simple  cubic  lattice,  in 
which  alternate  positions  are  occupied  by  the  positive  and  negative  ions. 
Each  ion  thus  has  six  nearest  neighbors  of  the  opposite  sign,  and  the 
electrostatic  attraction  between  the  ion  and  its  oppositely  charged  neigh- 
bors holds  the  crystal  together.  This  illustrates  a  principle  which  we 
mentioned  in  the  preceding  chapter  and  which  obviously  must  hold  for 
stability  in  an  ionic  crystal,  namely  that  for  stability  each  ion  must  be 
surrounded  by  as  many  ions  of  the  opposite  charge  as  possible,  and  the 
nearest  ions  of  the  same  sign  must  be  as  far  away  as  possible. 

The  second  common  structure  is  the  caesium  chloride  structure.  In 
this  structure,  ions  of  one  type  are  located  at  the  corners  of  a  cubic  lattice 


SEC'.  I) 


IONIC  CRYSTALS 


379 


and  ions  of  the  other  sign  at  the  centers  of  the  cubes.  In  this  structure 
each  ion  has  eight  neighbors  of  the  opposite  sign.  This  structure  is  shown 
in  Fig.  XXIII-2. 

The  third  and  fourth  structures  are  sometimes  called  the  zmchlcndr 
and  the  wurtzite  structures,  on  account  of  two  forms  of  ZnS.  The  zinc- 
blende  structure  is  also  often  called  the  diamond  structure,  since  it  is 
found  in  diamond  and  some  other  crystals.  The  fundamental  features 
of  both  structures  are  similar:  each  ion  is  tetrahedrally  surrounded  by  four 
ions  of  the  opposite  sign,  as  in  Fig.  XXIII-3  (a).  There  arc  a  number 
of  ways  of  joining  such  tetrahedra  to  form  regular  crystals,  however.  The 
diamond,  or  zincblende,  lattice  is  the  simplest  of  these.  In  the  first 
place,  tetrahedra  like  Fig.  XXIII-3  (a),  can  be  formed  into  sheets  like 


FIG.  XXIII-2. — Tho  caesium  chloride  structure. 

Fig.  XXIII-3  (6)  and  (c),  the  latter  looking  straight  down  along  the 
vertical  leg  of  the  tetrahedron,  so  that  the  ion  at  the  center  and  that 
directly  above  it  coincide  in  the  figure.  Then  in  the  diamond  structure, 
the  next  layer  up  is  just  like  that  shown,  but  shifted  along  so  that  atoms 
like  a,  6,  c  coincide  with  a',  6',  c'.  The  wurtzite  structure,  on  the  other 
hand,  has  the  next  layer  looking  just  like  the  one  shown  in  (6),  as  far  as  its 
projection  is  concerned,  but  actually  being  a  mirror  image  of  it  in  a  hori- 
zontal plane.  When  examined  in  three  dimensions,  the  wurtzite  struc- 
ture proves  to  be  less  symmetrical  than  the  diamond  structure,  in  that  the 
vertical  direction  stands  out  as  a  special  axis  in  the  crystal.  For  this 
reason,  the  length  of  the  vertical  distance  between  ions  in  the  wurtzite 
structure  does  not  have  to  equal  the  other  three  distances,  while  in  the 
diamond  structure  all  distances  must  be  the  same.  The  diamond,  or 
zincblende,  structure  is  shown  in  Fig.  XXIII-3  (d),  and  the  wurtzite 
structure  in  (e).  In  addition  to  these  two,  a  number  of  other  structures 
built  of  tetrahedra  can  exist,  in  which  the  two  types  of  planes,  the  one 


380 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 


Zn 


(e) 

FIG.  XXIII-3.— The  zincblende  and 
wurtzite  structures,  (a)  one  ion  tetra- 
hedrally  surrounded  by  four  others.  (6) 
and  (c)  sheets  formed  from  such  tetrahedra, 
in  perspective  and  plan,  (d)  the  zincblende 
structure  (which  is  identical  with  the  dia- 
mond structure  if  all  atoms  are  alike). 
(e)  the  wurtzite  structure. 


shown  in  (6)  and  its  mirror  image, 
are  arranged  in  various  regular  ways 
through  the  crystal.  Several  of 
these  structures  are  found,  for  in- 
stance, in  different  forms  of  carbo- 
rundum, SiC. 

We  shall  now  list  in  Table  XXIII- 
1  some  of  the  binary  ionic  compounds 
crystallizing  in  the  four  structures 
just  discussed.  Under  each  struc- 
ture, we  shall  arrange  the  compounds 
according  to  their  valence,  starting 
with  monovalent  substances.  We 
note  that  some  compounds  exist  in 
several  polymorphic  forms,  as  ZnS 
in  both  zincblende  and  wurtzite 
structures.  We  tabulate  not  merely 
the  substances  crystallizing  in  each 
form,  but  also  several  quantities 
characterizing  each  substance. 
First  we  give  the  distance  between 
nearest  neighbors,  r0,  in  angstroms. 
This  is  necessarily  the  distance  be- 
tween an  ion  of  either  sign  and  the 
nearest  ions  of  the  opposite  sign,  of 
which  there  are  six  in  the  sodium 
chloride  structure,  eight  in  the 
caesium  chloride,  and  four  in  the 
zincblende  and  wurtzite  structures. 
This  is  followed  by  a  calculated 
value  of  r0,  discussed  in  Sec.  2. 
Finally  we  give  the  melting  point 
where  this  is  known.  This  is  simply 
an  indication  of  the  tightness  of  bind- 
ing, since  a  strongly  bound  material 
has  a  high  melting  point.  We  notice 
that  the  divalent  substances  con- 
sistently have  a  much  higher  melting 
point  than  the  monovalent  ones. 
This  is  a  result  of  the  tighter  binding. 
Since  each  ion  in  a  divalent  crystal 
has  twice  as  great  a  charge  as  in  a 
similar  monovalent  one,  the  inter- 


SBC.  1J  IONIC  CRYSTALS  381 

TABLE  XX 1 1 1-1. — LATTICE  SPACINGS  OF  IONIC  CRYSTALS 


Material 

ro  observed, 
angstroms 

ro  calculated 

Melting  point,  °C. 

Sodium  Chloride  Structure 

LiF 

2.01 
2.57 
2.75 
3.00 

2.31 
2.81 
2.98 
3.23 

2.67 
3.14 
3.29 
3  .  53 

2.82 
3.27 
3.43 
3.66 

3.00 

3.27 
3.45 
3  62 

2.46 
2.77 
2.88 

2.10 
2.60 
2.73 

2.40 
2  84 
2.96 
2.97 

2.58 
3.01 
3.12 
3.33 

2.77 
3.19 
3.30 
3.50 

2.10 
2.60 
2.75 
3.00 

2.35 
2.85 
3.00 
3.25 

2.65 
3.15 
3.30 
3  .  55 

2.80 
3.30 
3.45 
3.70 

3  05 

3.25 
3.40 
3  65 

2.30 
2.80 
2  1)5 

2.15 
2  60 
2.70 

2.40 
2.85 
2.95 
3  15 

2.60 
3.05 
3.15 
3.35 

2.75 
3.20 
3.30 
3.50 

870 
613 
547 
446 

980 
804 
755 
651 

880 
776 
730 
773 

760 
715 
682 
642 

684 

435 
455 
13  1 

2800 
2572 

2430 
882 

1923 

Lid  

LiBr  

Lil  

NaF.  . 
NaCl 
NaBr.  . 
Nal. 

KF  

KC1. 
KBr. 
KI 

RbF  . 
RbCl  . 
RbBr. 
Rbl.         ..      . 

CaF  

NH4C1  .    . 
NH4Br       . 
NHJ 

AgF.     ..       . 
AgCl  
AgBr  . 

MgO  
MgS..        . 
MgSe  .    . 

CaO  
CaS 

CaSe  
CaTe 

SrO.       . 

SrS  ..        . 
SrSe  ... 

SrTe  

BaO  
BaS  
BaSe  

BaTe  

Caesium  Chloride  Structure 

CsCl  .... 
CaBr  

3.56 
3.71 
3.95 

3.34 

3.55 
3.70 
3.95 

3.25 

046 
636 
621 

Csl  .      . 
NH4C1  

382  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 

TABLE  XXIII-1. — LATTICE  SPACINGS  OF  IONIC  CRYSTALS. — (Continued) 


Material 

ro  observed, 
angstroms 

ro  calculated 

Melting  point,  °C. 

Caesium  Chloiide  Stiucture — (Continued) 


NH4Br 
NHJ. 

TICi  . 
TlBr 


CuCl 
CuBr 
CuT 

Bi-8 

BeSe 

BeTe 

ZnS 

ZnSe 

ZnTe 

CdS 

CdSe 

CdTe 

HgS 

HgSe 

HgTe 


3  51 
3  78 

3.33 
3.44 


3.40 
3.65 


Zincblende  Structure 


2.34 
2.46 
2  62 

2  10 
2  18 
2  43 

2  35 
2  45 
2  04 

2  52 
2. 02 
2  80 

2  53 
2.62 
2.79 


2.30 
2.45 
2  70 

2  10 
2  20 
2  40 

2  35 
2  45 
2.65 

2  50 
2  60 
2.80 

2  50 
2.60 
2  80 


430 
460 


422 
504 
005 


1800 


1750 


Wurtzite  Stiucture  (Fir.sl  distance  is  that  to  neighbor  along  axis,  second  to  three  neighbo 

in  same  layer) 


NH4F             .                                «, 

2  63,  2.76 

2  75 

BeO 

1.64,  1.60 

1  65 

2570 

ZnO 

1.94,2.04 

1.90 

ZnS 

2.36,  2.36 

2.35 

1850 

CdS                                                       .    . 

2.52,  2.56 

2.50 

1750 

CdSe 

2.63,  2.64 

2.60 

Data  regarding  crystal  stiucture,  here  and  in  other  tables  in  this  book,  are  taken  from  the  "Struk- 
turbencht,"  issued  as  a  supplement  to  the  Zettachrift  fdr  Krixtallographie  in  several  volumes  from  1931 
onward  This  is  the  (standard  reference  for  crystal  structure  data. 

ionic  forces  are  four  times  as  great,  with  correspondingly  large  latent  heats 
of  fusion.  Since  the  entropy  of  fusion  is  not  very  different  for  a  divalent 
crystal  from  what  it  is  for  a  monovalent  one,  this  means  that  the  melting 
point  of  a  divalent  crystal  must  be  several  times  as  large  as  for  a  mono- 
valent one,  as  Table  XXIII-1  shows  it  to  be. 

2.  Ionic  Radii. — The  first  question  that  we  naturally  ask  about  the 
crystals  is,  what  determines  the  lattice  spacings?    Examination  of  the 


SBC.  2]  IONIC  CRYSTALS  383 

experimental  values  shows  that  these  spacings  are  very  nearly  additive; 
that  is,  we  can  assign  values  to  radii  of  the  various  ions,  such  that  the 
sums  of  the  radii  give  the  distance  between  the  corresponding  ions  of  the 
crystals.  A  possible  set  of  ionic  radii  is  given  in  Table  XXIII-2.  Using 

TABLE  XXIII-2.— IONIC  RADII 
(Angstroms) 


Be++ 
0.20 

0  80 

Mg+H 

Na+ 

F- 

o— 

0.70 

1  05 

1  30 

1.45 

Ca++ 

K+ 

ci- 

S~ 

Zn+4 

Cu+ 

0.95 

1  35 

1.80 

1.90 

0.45 

0.50 

Sr++ 

Rb+ 

Br~ 

So— 

Cd-*  + 

Ag+ 

1  15 

I  50 

1.95 

2  00 

0.60 

1.00 

Ba^+ 

Csf 

I- 

Te— 

Hg++ 

1   30 

1  75 

2.20 

2  20 

0.60 

1  45 

these  radii,  which  as  will  be  observed  are  smoothed  off  to  0  or  5  in  the  last 
place,  the  values  r0  calculated  of  Table  XXIII-1  are  computed.  The 
agreement  between  calculated  and  observed  lattice  spacing  is  surely 
rather  remarkable.  There  have  been  a  good  many  discussions  seeking  to 
show  the  reasons  for  the  small  errors  in  the  table,  the  departures  from 
additivity.  In  particular,  there  are  good  reasons  for  thinking  that 
different  radii  should  be  used  for  the  sodium  chloride  structure,  where 
every  ion  is  surrounded  by  six  neighbors,  from  those  used  in  the  zinc- 
blende  and  wurtzite  structures,  where  there  are  four  neighbors.  But  the 
comparative  success  of  the  calculations  of  Table  XXIII-1,  where  both 
types  of  structure  are  discussed  by  means  of  the  same  radii,  shows  that 
these  corrections  are  comparatively  unimportant  and  the  significant  fact- 
is  that  the  agreement  is  as  good  as  it  is. 

There  is  one  point  in  which  our  assumed  valuqp  for  ionic  radii  are  not 
uniquely  determined.  The  observed  interionic  spacings  can  determine 
only  the  sum  of  the  radii  for  a  positive  and  negative  ion  of  the  same 
valency.  We  can  add  any  constant  to  all  the  radii  of  the  positive  ions 
of  one  valency,  subtract  the  same  constant  from  the  radii  of  the  negative 
ions,  without  changing  the  computed  results.  On  the  other  hand,  the 
difference  between  the  assumed  radii  of  two  ions  of  the  same  sign  cannot 
be  changed  without  destroying  agreement  with  experiment.  We  have 
chosen  this  arbitrary  additive  constant  in  such  a  way  as  to  make  the 
positive  ion,  as  K+,  an  appropriate  amount  smaller  than  the  negative 
ion,  such  as  Cl~,  which  contains  the  same> number  of  electrons,  considering 
the  tendency  for  extra  negative  electrons  to  be  repelled  from  an  atom, 
making  negative  ions  large,  positive  ones  small.  Our  estimate  is  probably 


384 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 


not  reliable,  however,  and  the  absolute  values  should  not  be  taken  seri- 
ously as  representing  in  any  way  the  real  radii  of  the  ions.  In  particular, 
the  ion  of  Be++  is  pretty  certainly  not  so  small  as  its  extremely  small 
radius,  0.20  A,  would  suggest. 

It  is  interesting  to  compare  the  radii  of  Table  XXIII-2  with  those  of 
Table  XXI-4.  It  will  be  seen  that  though  they  are  of  the  same  order  of 
magnitude,  the  ionic  radii  of  Table  XXIII-2  are  several  times  the  radii 
of  the  corresponding  orbits  in  Table  XXI-4.  Remembering  that  Table 
XXI-4  gives  the  radius  of  maximum  density  in  the  shell,  we  see  that 
the  region  occupied  by  electrons,  given  by  the  ionic  radii,  is  several  times 
the  sphere  whose  radius  is  the  radius  of  maximum  density.  This  is  surely 
a  natural  situation,  since  the  charge  density  falls  off  rather  slowly  from  its 
maximum. 

In  Table  XXIII-2,  we  can  interpolate  between  the  monovalent  posi- 
tive and  negative  ions  to  get  radii  for  the  inert  gas  atoms.  Thus  we  find 
approximately  the  following:  No  1.1  A,  A  1.5  A,  Kr  1.7  A,  Xe  1.95  A.  It 
is  interesting  to  compute  the  volumes  of  the  inert  gas  atoms  which  we 
should  get  in  this  way,  and  compare  with  the  volumes  which  we  find  for 
them  from  the  constant  b  of  Van  der  Waals'  equation.  For  neon,  for 
instance,  the  volume  from  Table  XXIII-2  would  be 

^r(l.l)3  =  5.53  X  10~24  cc.  per  atom 

=  5.53  X  10~24  X  6.03  X  1023  cc.  per  mole  =  3.33  cc.  per  mole. 

In  Table  XXIII-3  we  give  the  volumes  computed  this  way,  the  values  of 
TABLE  XXIII-3. — VOLUMES  OF  INERT  GAS  ATOMS 


Volume  from 

i, 

b 

Volume  of 

ionic  radius 

volume 

liquid 

Ne  

3  33 

17.1 

5  1 

16.7 

A        

8  6 

32  2 

3  8 

28  1 

Kr 

12  5 

39  7 

3  2 

38  9 

Xe 

18  8 

50.8 

2  7 

47  5 

The  volumes  computed  from  ionic  radii  are  interpolated  as  described  in  the  text  from  the  ionic  radii 
of  Table  XXIII-2.  Values  of  b  are  taken  from  Table  XXI V-l.  Volumes  are  in  cubic  centimeters  per 
mole. 

Van  der  Waals  6,  computed  from  the  critical  pressure  and  temperature, 
the  ratio  of  6  to  the  volume  computed  from  the  ionic  radius,  and  finally 
the  molecular  volume  of  the  liquid.  From  the  table  we  see  that  the 
values  of  b,  and  the  volumes  of  the  liquid,  agree  fairly  closely  with  each 
other,  and  are  three  to  five  times  the  volume  of  the  molecule,  as  computed 
from  Table  XXIII-2.  Since  the  liquid  is  a  rather  closely  packed  struc- 
ture, this  seems  at  first  sight  a  little  peculiar.  The  explanation,  however, 
is  not  complicated.  The  molecules  really  are  not  hard,  rigid  things  but 


SEC.  3]  IONIC  CRYSTALS  385 

are  quite  compressible.  Thus  when  they  are  held  together  loosely,  by 
small  forces,  they  have  fairly  large  volumes,  while  when  they  are  squeezed 
tightly  together  they  have  much  smaller  volumes.  Now  the  ions  used  in 
computing  Table  XXIII-2  are  held  together  by  strong  electrostatic  forces, 
equivalent  to  a  great  many  atmospheres  external  pressure.  Thus  their 
atoms  are  greatly  compressed,  as  we  saw  in  the  preceding  paragraph;  by 
interpolating,  we  find  volumes  for  the  inert  gas  atoms  in  a  very  com- 
pressed state.  On  the  other  hand,  the  real  inert  gas  liquids  are  held 
together  only  by  weak  Van  der  Waals  forces,  which  cannot  squeeze  the 
atoms  to  nearly  such  a  compressed  state.  Thus  it  is  reasonable  that 
the  volumes  computed  from  the  radii  of  Table  XXIII-2  should  be  much 
smaller  than  the  volumes  of  the  liquids.  These  remarks  lead  to  a  conclu- 
sion regarding  the  divalent  ions.  Being  doubly  charged,  the  forces  in  the 
divalent  crystals  are  much  greater  than  in  the  monovalcnt  ones,  as  we 
have  mentioned  earlier,  and  the  ions  are  correspondingly  more  squeezed. 
Thus  the  sizes  of  the  divalent  ions  in  Table  XXIII-2  are  really  too  small 
in  comparison  with  the  monovalent  ones,  and  we  cannot  get  a  correct  idea 
of  the  relative  sizes  of  the  ions  by  studying  Table  XXIII-2. 

3.  Energy  and  Equation  of  State  of  Simple  Ionic  Lattices  at  the  Abso- 
lute Zero. — The  structure  of  the  ionic  lattices  is  so  simple  that  we  can 
make  a  good  deal  of  progress  toward  explaining  their  equations  of  state 
theoretically.  In  this  section  we  shall  consider  their  behavior  at  the 
absolute  zero  of  temperature.  Then  the  thermodynamic  properties  can 
be  derived  entirely  from  the  internal  energy  as  a  function  of  volume,  as 
discussed  in  Chap.  XIII,  Sec.  3,  and  particularly  in  Eq.  (3.4)  of  that 
chapter.  The  internal  energy  at  the  absolute  zero  is  entirely  potential 
energy,  arising  from  the  forces  between  ions.  As  we  saw  in  the  preceding 
chapter,  these  forces  are  of  two  sorts.  In  the  first  place,  there  are  the 
electrostatic  forces  between  the  charged  ions,  repulsions  between  like  ions 
and  attractions  between  oppositely  charged  ones.  The  net  effect  of  these 
forces  is  an  attraction,  for  each  ion  is  closer  to  neighbors  of  the  opposite 
sign,  which  attract  it,  than  to  those  of  the  same  sign,  which  repel  it.  In 
addition  to  this  electrostatic  force,  there  are  the  repulsive  forces  between 
ions,  resulting  from  the  exclusion  principle,  vanishingly  small  at  large 
distances,  but  increasing  so  rapidly  at  small  distances  that  they  prevent 
too  close  an  approach  of  any  two  ions. 

First  we  take  up  the  electrostatic  forces.  We  shall  consider  only 
the  sodium  chloride  structure,  though  other  types  are  not  essentially 
more  difficult  to  work  out.  Let  the  charge  on  an  ion  be  ±  zet  where  z  =  1 
for  monovalent  crystals,  2  for  divalent,  etc.  We  shall  now  find  the 
electrostatic  potential  energy  of  the  crystal,  by  summing  the  terms 
±z*e*/d  for  all  pairs  of  ions  in  the  crystal,  where  d  is  the  distance  between 
the  ions.  We  start  by  choosing  a  certain  ion,  say  a  positive  one,  and 


386 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 


summing  for  all  pairs  of  which  this  is  a  member.  Let  the  spacing  between 
nearest  neighbors  of  opposite  sign  be  r.  Assume  the  ion  we  are  picking 
out  is  at  the  origin,  and  that  the  $,  y,  z  axes  point  along  the  axes  of  the 
crystal.  Then  other  ions  will.be  found  for  x  =  n\r,  y  =  n2r,  z  =  n8r, 
where  n\,  w2,  w3  are  any  positive  or  negative  integers.  It  is  easy  to  see 
that  if  n\  +  n2  +  n$  is  even  the  other  ion  will  be  positive,  and  if  it  is  odd 
it  will  be  negative;  for  this  means  that  increasing  one  of  the  three  integers 
by  unity,  which  corresponds  to  a  translation  of  r  along  one  of  the  three 
axes,  will  change  the  sign  of  the  ion,  which  is  characteristic  of  the  sodium 
chloride  structure.  The  distance  from  the  origin  to  the  ion  at  nir,  n2r, 
n$r  is  of  course  r\Ai?  +  n\  +  n|,  so  that  the  potential  energy  arising  from 
the  pair  of  ions  in  question  is  ±z2e2/(rv/^i  +  ™!  +  ™1)>  where  the  sign 
is  +  if  n\  +  r?2  +  n3  is  even,  —  if  it  is  odd.  Now  there  will  be  a  number 
of  ions  at  the  same  distance  from  the  origin,  and  since  the  potential  is  a 
scalar  quantity,  these  will  all  contribute  equal  amounts  to  the  potential 
energy  and  can  be  grouped  together.  We  can  easily  find  how  many  such 
ions  there  are.  They  all  arise  from  the  same  set  of  numerical  values  of 
Wi,  ^2,  fta,  but  arranged  in  the  different  possible  orders,  with  all  possible 
combinations  of  sign.  If  rii,  n2,  n3  are  all  different  and  different  from 
zero,  there  are  six  ways  of  arranging  them,  and  each  can  have  either  sign, 
so  there  are  eight  possible  combinations  of  signs,  making  48  equal  terms. 
If  two  out  of  the  three  n's  are  equal  in  magnitude,  but  not  zero,  there  are 
only  three  arrangements,  but  eight  sign  combinations,  making  24  equal 
terms.  If  all  three  arc  equal  in  magnitude,  there  are  still  the  eight  com- 
binations of  signs,  making  8  terms.  If  one  of  the  rz's  is  zero,  there  is  no 
ambiguity  of  sign  connected  with  it,  so  that  there  are  only  half  as  many 
possible  terms,  and  if  two  of  the  n's  arc  zero  there  are  only  a  quarter  as 
many  terms.  By  use  of  these  rules,  we  can  set  up  Table  XXIII-4,  giving 

TABLE  XXIII-4. — CALCULATION  OP  ELECTROSTATIC  ENERGY,   SODIUM  CHLORIDE 

STRUCTURE 


n\n^n^ 

Number  of 
terms 

Distance 

Contribution  to  potential  energy 

1  0  0 

6 

rVT 

(z*e*/r)  X  (-6/VT) 

=     -6.00C 

1  1  0 

12 

rV2 

(z*e*/r)  X  (12/V2) 

8.48* 

1  1  1 

8 

r\/3 

(zV/r)  X  (-8/\/3) 

=     -4.62( 

200 

6 

r\/4 

<«V/r)  X  (6/V4) 

3.  OCX 

2  1  0 

24 

r\/5 

(zW/r)  X  (-24/V6) 

=  -10.73( 

2  1  1 

24 

r\/6 

(«V/r)  X  (24/\/6) 

9.80C 

220 

12 

r\/8 

frV/r)  X  (12A/8) 

4.244 

2  2  1 

24 

r\/9 

(rfe»/r)  X  (-24/V9) 

-     -8.00C 

222 

8 

r\/12 

(iV/r)  X  (8A/I5) 

=       2.31C 

SBC.  3]  IONIC  CRYSTALS  387 

the  values  of  n\n^n^  arranged  in  order  so  that  n\  £:  n*  ^  ns;  the  number 
of  neighbors  associated  with  various  combinations  of  these  n's;  the 
distance;  and  the  total  contribution  of  all  these  neighbors  to  the  potential 
energy.  This  table  can  be  easily  extended  by  analogy  to  any  desired 
distance. 

On  examining  Table  XXIII-4,  we  see  that  the  terms  show  no  tendency 
to  decrease  as  the  distance  gets  greater,  though  they  alternate  in  sign.     It 
is  plainly  out  of  the  question  to  find  the  total  potential  energy  simply  by 
adding  terms,  for  the  series  would  not  converge.     But  we  can  adopt  a 
device  that  brings  very  rapid  convergence. 
As  shown  in  Fig.  XXIII-4,  we  set  up  a 
cube,  its  faces  cutting  through  planes  of 
atoms.     Then  if  we  count  each  ion  on  a 
face  of  the  cube  as  being  half  in  the  cube, 
each  one  on  an  edge  as  being  a  quarter 
inside,  and  each  one  at  a  corner  as  being 
one-eighth  inside,  the  total  charge  within 
the  cube  will  be  zero.     Enlarging  the  cube 
by  a  distance  d  all  around  will  then  add 

a  volume  that  contains  a  net  charge  of  zero,  FIO.  xxni-4.—  Cube  of  ions 
and  so  contributes  fairly  little  to  the  total  ,tak!n  from  sodmm  cUoride  type  of 

.    ,          ,  .    .          T         ,  ,      .„   lattice. 

potential  at  the  origin.     In  other  words,  if 

we  set  up  the  total  potential  energy  of  interaction  of  the  ion  at  the  origin, 
and  all  ions  in  such  a  cube,  the  result  should  converge  fairly  rapidly  as  the 
size  of  the  cube  is  increased.  This  is  in  fact  the  case.  If  the  cube  extends 
from  —  r  to  +r  along  each  axis,  the  points  100  will  be  counted  as  half  in 
the  cube;  those  at  110  will  be  one  quarter  in;  and  those  at  111  one-eighth 
in.  Thus  the  contribution  of  these  terms  to  the  potential  energy  will  be 

6.000   ,8.485       4.620 


If  it  extends  from  —  2r  to  2r,  the  points  100,  110,  111  will  be  entirely 
inside  the  cube,  those  at  200,  210,  211  will  be  half  inside,  those  at  220,  221 
one  quarter  inside,  and  those  at  222  one-eighth  inside.  Thus  for  the 
potential  energy  we  have 


-6.000  +  8.485  -  4.620  +  -  + 


The  next  approximation  is  found  similarly  to  be  —1.714,  and  successive 
terms  oscillate  slightly  but  converge  rapidly,  to  the  value  —1.742,  the 
correct  value,  which  as  we  see  is  very  close  to  the  value  (3.2). 


388  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 

As  we  have  just  seen,  the  sum  of  potential  energy  terms  between  one 
positive  ion  and  all  its  neighbors  is 

~2f>2 

-1.742^--  (3.3) 

The  number  1.742  is  often  called  Madelung's  number,  since  it  was  first 
computed  by  Madelung.1  We  should  have  found  just  the  same  answer 
if  we  had  started  with  a  negative  ion  instead  of  a  positive  one,  since 
the  signs  of  all  charges  would  have  been  changed,  and  each  term  involves 
the  product  of  two  charges.  Now  to  get  the  total  energy,  we  must  sum 
over  all  pairs  of  neighbors.  Let  there  be  N  molecules  in  the  crystal 
(that  is,  N  positive  ions  and  N  negative  ions).  Then  each  of  the  positive 
ions  contributes  the  amount  (3.3)  to  the  summation,  and  each  of  the 
negative  ions  contributes  an  equal  amount,  so  that  at  first  sight  we  should 

/         z*c*\ 
say  that  the  total  energy  was  —  2AT(  1.742  —  )•     This  is  incorrect,  how- 

ever, for  in  adding  up  the  terms  in  this  way  we  have  counted  each  term 
twice.  Each  pair  of  ions,  say  ion  a  and  ion  6,  has  been  counted  once  when 
ion  a  was  the  one  at  the  origin,  b  at  another  point,  and  then  again  when  b 
was  at  the  origin.  To  correct  for  this,  we  must  divide  our  result  by  two, 
obtaining 

Electrostatic  energy  =  -7VM.742—  V  (3.4) 

Next  we  consider  the  repulsive  forces  between  ions.  The  only  thing 
we  can  say  about  these  is  that  they  are  negligible  for  large  interionic  dis- 
tance, and  get  very  large  as  the  distance  becomes  small.  A  simple 
function  having  this  property  is  l/dm,  where  d  is  the  distance  between 
ions,  m  is  a  large  integer.  We  shall  tentatively  use  this  function  to  repre- 
sent the  repulsions.  To  give  results  agreeing  with  experiment,  it  is  found 
that  m  must  be  of  the  order  of  magnitude  of  8  or  10  in  most  cases.  Thus, 
let  the  potential  energy  between  two  positive  ions  at  distance  d  apart  be 
a++/dm,  between  two  negatives  a  __  /dm,  and  between  a  positive  and  a 
negative  a+_/dm.  The  coefficients  a  are  all  assumed  to  be  positive, 
leading  to  repulsions.  Then  we  can  compute  the  total  repulsive  potential 
energy,  just  as  we  have  computed  the  electrostatic  attraction,  only  now 
the  series  converges  so  rapidly  that  we  do  not  have  to  adopt  any  special 
methods  of  calculation.  Thus  for  the  sum  of  all  pairs  of  ions  in  which 
one  is  a  positive  ion  at  the  origin,  we  have 

(3-5) 


1  For  other  methods  of  computation,  see  for  instance  M.  Born,  "Problems  of 
Atomic  Dynamics,"  Series  II,  Lecture  4,  Massachusetts  Institute  of  Technology,  1926. 


SBC.  3]  IONIC  CRYSTALS  389 

To  have  a  specific  case  to  consider,  let  us  take  w  =  9,  which  works  fairly 
satisfactorily  for  most  of  the  crystals.     Then  the  series  (3.5)  becomes 

0.530a++  +  0.0571a+_  +  0.0117a++  +  0.0171a+_ 
+  0.00750++  +  0.0010a++  +  0.0012a+_  •  •  •  ) 

0.550a++)-      (3.6) 


There  is  a  similar  formula  for  the  sum  of  all  pairs  in  which  one  ion  is  a 
negative  one  at  the  origin,  with  a  __  in  place  of  a++.  Then,  as  before,  we 
can  get  the  total  energy  by  multiplying  each  of  these  formulas  by  N/2. 
That  is,  the  total  repulsive  energy  is 

Repulsive  energy  =  AT^)[6.075a+_  +  O.OH»(«y  +  a.-jj 

-  £  (3.7) 

where  A  is  a  constant.     More  generally, 

^ 
Repulsive  energy  =  —  •  (3.8) 

We  can  now  combine  the  electrostatic  energy  from  Eq.  (3.4)  and  the 
repulsive  energy  from  Eq.  (3.8)  to  obtain  the  total  internal  energy  at  the 
absolute  zero, 

f/0=  -Ni.7*££  +  A.  (3.9) 

To  make  connection  with  our  discussion  of  the  equation  of  state  in  Chap. 
XIII,  we  should  expand  Eq.  (3.9)  in  power  series  about  r0,  the  value  of  r 
at  which  UQ  has  its  minimum  value.  First  we  have 

df/o       *71  -An**e*  4 

-r-  =  Nl.  742—  ~  --  w-^ri 

ttT  T2  f.m+1 

=  0  when  r  =  r0.  (3.10) 

From  Eq.  (3.10)  we  can  write  A  in  terms  of  other  quantities,  finding 


We  then  have 


A  =  N1.742z*e*-  (3.11) 

m  ^       ' 


U.  -  -tfl.74a  -  .  (3.12) 

TOr 


390  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 

Expanding  Eq.  (3.12)  in  power  series  in  r  —  r0,  we  have 

U0  -  -#1.742—  2  (  1  -  1) 
ro  \        m/ 


Equation  (3.13)  is  of  the  form  of  Eq.  (3.4),  Chap.  XIII, 


[700  +  ATc7ips  -  9(P?  -  Pto  '  '  '     '     (3-14) 


To  see  the  significance  of  c  in  this  equation,  it  will  be  remembered  that  the 
volume  per  molecule  is  given  by 

jj  =  cr».  (3-15) 

In  this  case,  consider  a  cube  of  edge  r,  with  a  positive  or  negative  ion  at 
each  corner.  There  are  8  ions  at  the  corners,  each  counting  as  if  it  were 
one  eighth  inside  the  cube,  so  that  the  cube  contains  just  one  ion,  or  half  a 
molecule.  In  other  words,  V/N  =  2r3,  or  c  =  2,  for  the  sodium  chloride 
structure.  It  will  also  be  remembered  that  the  quantities  PJ  arid  P£  in 
Eq.  (3.15)  are  coefficients  in  the  expansion  of  the  pressure  as  a  function 
of  volume,  at  the  absolute  zero  of  temperature,  as  shown  in  Eq.  (1.5)  of 
Chap.  XIII.  These  can  be  found  from  experiment  by  Eqs.  (1.10)  of 
Chap.  XIII.  Identifying  Eqs.  (3.13)  and  (3.14),  we  can  then  solve  for 
t/oo,  the  energy  at  zero  pressure  at  the  absolute  zero,  P?  andPJj,  finding 

(3.16) 
^  =  ^~r(m-l),  (3.17) 

*-^%("-»(«  +  W-  <3-18> 

4.  The  Equation  of  State  of  the  Alkali  Halides.— The  alkali  halides, 
the  fluorides,  chlorides,  bromides,  and  iodides  of  lithium,  sodium,  potas- 
sium, rubidium,  and  caesium  have  been  more  extensively  studied  experi- 
mentally than  any  other  group  of  ionic  crystals.  For  most  of  these 
materials,  enough  data  are  available  to  make  a  fairly  satisfactory  com- 
parison between  experiment  and  theory.  The  observations  include  the 
compressibility  and  its  change  with  pressure,  at  room  temperature,  from 
which  the  quantities  ai(T),  a2(T)  of  Eq.  (1.1),  Chap.  XIII,  can  be  found 


t/oo  =  -tfl.742^Yl  -  A 


SEC.  4]  IONIC  CRYSTALS  391 

for  room  temperature;  very  rough  measurements  of  the  change  of  com- 
pressibility with  temperature,  giving  the  derivative  of  ai  with  respect  to 
temperature;  the  thermal  expansion,  giving  the  derivative  of  ao  with 
respect  to  temperature;  and  the  specific  heat.  There  are  two  sorts  of 
comparison  between  theory  and  experiment  that  can  be  given.  In  the 
first  place,  we  have  found  a  number  of  relations  between  experimental 
quantities,  not  involving  the  detailed  theory  of  the  last  section,  which  we 
can  check.  Secondly,  we  can  test  the  relations  of  the  last  section  and  see 
whether  they  are  in  agreement  with  experiment. 

The  relations  between  experimental  quantities  mostly  concern  the 
temperature  effects.  First,  let  us  consider  the  specific  heat.  In  Chap. 
XV,  Sec.  3,  we  have  seen  that  it  should  be  fairly  accurate  to  use  a  Debye 
curve  for  the  specific  heat  of  an  alkali  halido,  using  the  total  number  of 
ions  in  determining  the  number  of  characteristic  frequencies  in  that 
theory.  It  is,  in  fact,  found  that  the  experimental  values  fit  Debye 
curves  accurately  enough  so  that  we  shall  not  reproduce  them.  We  can 
then  determine  the  Debye  temperatures  from  experiment,  and  in  Table 
XXIII-5  we  give  these  values  for  NaCl  and  KC1,  the  two  alkali  halides 

TABLE  XXIII-5. — DEBYE  TEMPERATURES  FOR  ALKALI  HALIDES 


NaCl,  °  abs. 

KC1,  °  abs. 

O/>  from  specific  heat  ... 
Qr>  from  elastic  constants  
Bo  from  residual  rays  

281 
305 
277 

230 
227 
227 

Data  are  taken  from  the  article  by  Schrodinger,  "Spezifische  Warme,"  in  "Handbuch  der  Physik," 
Vol.  X,  Springer,  1926.  This  volume  contains  a  number  of  articles  bearing  on  topics  taken  up  in  this 
book  and  is  useful  for  reference. 

which  occur  in  a  crystalline  form  in  nature  and  for  which  most  measure- 
ments have  been  made.     But  from  Eqs.  (3.1),  (3.5),  and  (3.9),  Chap. 

XIV,  we  have  information  from  which  the  Debye  temperature  can  be 
calculated  from  the  elastic  constants  and  the  density.     These  constants 
are  known  for  NaCl  and  KC1,  and  in  the  table  we  also  give  the  calculated 
Debye  temperature  found  from  the  elastic  constants.     Finally,  in  Chap. 

XV,  Sec.  4,  we  have  seen  that  the  frequency  of  the  residual  rays  should 
agree  with  the  Debye  frequency.     In  the  table  we  give  the  observed 
frequency  of  the  residual  rays,  in  the  form  of  a  characteristic  temperature. 
We  see  that  the  agreement  between  the  three  values  of  Debye  temperature 
in  Table  XXIII-5  is  remarkably  good,  indicating  the  general  correctness 
of  our  analysis  of  the  vibrational  problem.    As  a  matter  of  fact,  the  agree- 
ment is  better  than  we  could  reasonably  expect,  on  account  of  the  approx- 
imations made  in  Debye's  theory,  and  there  are  many  other  crystals  for 
which  it  is  not  so  good,  so  that  we  may  lay  the  excellent  agreement  here 
partly  to  coincidence.  ' 


392  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 

Next  we  may  consider  the  equation  of  state.  From  the  compressi- 
bility and  its  change  with  pressure  we  have  the  quantities  ai,  a2  of  Eq. 
(1.1),  Chap.  XIII,  as  we  have  mentioned  before,  and  from  the  thermal 
expansion  we  have  the  derivative  of  ao  with  respect  to  temperature,  but 
not  its  value  itself.  Since  the  thermal  expansion  of  most  of  the  materials 
has  not  been  measured  as  a  function  of  temperature  we  cannot  integrate 
the  derivative  to  find  values  of  a0.  The  quantity  a0  comes  in  only  as  a 
small  correction  term  in  applications,  however,  and  if  we  are  willing  to 
assume  Griineisen's  theory  we  can  calculate  it  to  a  sufficiently  good 
approximation.  From  Eq.  (1.9)  or  (1.10)  of  Chap.  XIII,  we  can  find  a0 
from  the  thermal  pressure  and  the  compressibility.  The  thermal  pressure 
PO,  the  pressure  necessary  to  reduce  the  volume  to  the  volume  at  the 
absolute  zero,  is  given  by  Eq.  (4.12)  of  Chap.  XIII.  For  the  present 
case,  if  there  are  N  molecules,  2N  atoms,  and  6AT  degrees  of  freedom  of  the 
atoms,  and  if  we  assume  according  to  Gruneisen  that  all  the  7/8  are  equal 
to  7,  this  equation  gives 


where  v  is  a  suitable  mean  of  the  natural  frequencies  Vj.  The  Eq.  (4.1) 
is  the  limiting  case  suitable  for  high  temperatures,  where  the  thermal 
expansion  is  constant,  and  can  be  regarded  as  the  integral  of  Eq.  (4.16), 
Chap.  XIII.  If  we  assume  as  a  rough  approximation  that  v  can  be 
replaced  by  the  Debye  frequency,  Eq.  (4.1)  leads  to 


Po  =  ^Fl  T  -  ^ 

v  o 

6Nk     /„,      9D 


p  ,r         D  ,,0, 

ao  =  aiP0  =  -y^yx(  T  -  -^l  (4.2) 

where  6Nk  is  the  heat  capacity  at  high  temperatures,  x  the  compressi- 
bility, GD  the  Debye  temperature.  Equation  (4.2)  should  hold  for 
temperatures  considerably  above  half  the  Debye  temperature  and  should 
be  fairly  accurate  at  temperatures  as  high  as  the  Debye  temperature, 
where  the  specific  heat  is  fairly  constant.  From  Table  XXIII-5,  we  see 
that  the  Debye  temperatures  for  these  materials  are  of  the  order  of 
magnitude  of  room  temperature,  so  that  we  should  expect  Eq.  (4.2)  to 
be  fairly  accurate  at  room  temperature  where  the  observations  have  been 
made. 

Using  the  approximation  (4.2)  and  measured  values  of  ai  and  a*  at 
room  temperature,  we  can  use  Eqs.  (1.10),  Chap.  XIII,  to  find  PO,  Pi,  and 
P2.  We  find,  as  a  matter  of  fact,  that  the  term  in  a0,  in  Eq.  (1.10)  for 
Pi,  is  a  small  correction  term,  so  that  to  a  gojod  approximation  Pi  and  P2 


SEC.  4] 


IONIC  CRYSTALS 


393 


can  be  found  directly  from  the  observed  compressibility  and  its  change 
with  pressure.     In  Table  XXIII-6,  we  give  values  of  PI  and  P^  computed 

TABLE   XXIIT-6. — QUANTITIES   CONCERNED   IN   EQUATION   OF  STATE  OF  ALKALI 

HALIDES 


Pi 

P2 

7 
(Grttn- 
eisen) 

7 
(Eq. 

4.3) 

m 

P2 
calculated 

7 
(Eq. 
4.4) 

LiF  

0.652  X  1012 

2.41  X  1012 

1.34 

3.02 

5.80 

1.72  X  1012 

1.97 

LiCl...     . 

0.293 

0  815 

1  52 

2  11 

6  75 

0  819 

2  12 

LiBr  

0.232 

0.635 

1.70 

2.06 

6.95 

0.655 

2.16 

NaCl  

0.238 

0.600 

1  63 

1.85 

7.66 

0.700 

2  27 

NaBr  

0.197 

0.476 

(1  56) 

1.75 

7.97 

0.590 

2.33 

KF  

0.302 

0  885 

1.45 

2.26 

7.90 

0.900 

2.32 

KC1  

0.178 

0  402 

1.60 

1.59 

8.75 

0.557 

2.46 

KBr 

0  149 

0  341 

1  68 

1  62 

8  82 

0  468 

2  47 

KI  

0.117 

0.259 

1.63 

1.54 

9.15 

0.373 

2.52 

RbBr..   .. 

0.126 

0.268 

(1.37) 

1.46 

8.82 

0.395 

2.47 

Rbl  . 

0  104 

0  226 

(1.41) 

1.50 

9  37 

0  335 

2.56 

Values  of  Pi  and  P*  are  computed  from  data  of  J.  C.  Slater,  Phys.  Rev.,  23, 488  (1024).  Values  of  y 
by  Gruneisen's  method  are  taken  from  the  article  by  Gruneisen,  "Zustand  des  festen  Korpers,"  in 
"Handbuch  der  Physik,"  Vol.  X,  Springer,  1926.  Values  of  m,  Pz  calculated  and  the  two  calculated 
values  of  7,  are  found  as  described  in  the  text. 

in  this  way,  for  those  of  the  alkali  halides  for  which  suitable  measurements 
have  been  made.  We  can  now  make  a  calculation  that  will  serve  to  check 
Gruneisen's  theory  of  thermal  expansion.  In  the  first  place,  from  Eq. 
(4.16),  Chapter  XIII  or  from  Eq.  (4.2)  above,  we  see  that  y  can  be  com- 
puted from  the  thermal  expansion,  specific  heat,  compressibility,  and 
density.  But  if  we  assume  Debye's  theory  and  neglect  the  variation  of 
Poisson's  ratio  with  volume,  we  have  seen  in  Chap.  XIV,  Eq.  (4.6),  that 
we  can  write  y  in  terms  of  P\  and  P*: 


(4.3) 


This  gives  us  two  independent  ways  of  computing  7,  and  if  they  agree 
with  each  other  we  can  conclude  that  Gruneisen's  theory  is  fairly  accurate. 
In  Table  XXIII-6,  we  use  the  thermal  expansion,  specific  heat,  and 
volume  per  mole  of  the  crystals  at  room  temperature,  in  order  to 
compute  y  by  Gruneisen's  theory.  In  the  next  column  we  give  y  com- 
puted by  Eq.  (4.3).  It  will  be  seen  that  the  two  sets  of  numbers  agree  in 
order  of  magnitude,  and  for  most  of  the  crystals  they  are  in  rather  close 
quantitative  agreement.  Putting  this  in  another  way,  if  we  knew  merely 


394  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 

the  compressibility  and  its  change  with  pressure,  w^  could  make  a  good 
calculation  of  thcTtherma^^^  the  thermaTJ^pansion 

and  compressihilHy  we  could  calculate  jthei  Changs  of '£gngp^ssjil^^Y_yth 
pressure.  The  only  serious  discrepancies  come  with  thejluosdes,  the 
most^incompressible  of  the  crystals,  and  it  will  be  foundjin  latcjr^clmprers 
that  thliT'situation  holds  also  for  metals:  the  more  incompressible  the 
crystal*  the  poorer  the  agreement  between  the  7  computed  from  the 
clastic  congtants^'and  that  found  from  the  thermal  expansion.  Experi- 
mentally, the  change  of  compressibility  with  pressure  is  greater  than  we 
should  conclude  from  the  thermal  expansion,  or  conversely  the  thermal 
expansion  is  less  than  we  should  suppose  from  the  change  of  compressi- 
bility with  pressure.  The  reason  for  this  discrepancy  is  not  understood, 
but  it  presumably  arises  from  inaccuracies  in  Griineisen's  assumptions, 
since  there  is  no  indication  that  the  experimental  error  could  be  great 
enough  to  explain  the  lack  of  agreement  between  theory  and  experiment. 
The  comparisons  with  experiment  which  we  have  made  so  far  do  not 
involve  the  assumptions  of  the  preceding  section  about  interatomic  forces. 
We  shall  now  see  how  far  those  assumptions  are  correct.  From  Eqs. 
(3.17)  and  (3.18)  we  can  find  values  of  Pi  and  P2  at  the  absolute  zero,  in 
terms  of  r0,  which  we  can  take  from  experiment,  and  the  one  parameter 
m.  For  approximate  purposes,  we  can  replace  the  values  of  PI  and  P2 
at  the  absolute  zero  by  the  values  at  room  temperature.  Then  we  can 
ask  whether  it  is  possible  to  find  a  single  value  of  m  that  will  reproduce 
both  PI  and  P2.  To  test  this,  we  have  used  PI  to  find  a  value  of  m  and 
then  have  substituted  this  in  Eq.  (3.18)  to  compute  a  value  of  P2,  compar- 
ing this  computed  value  with  experiment.  These  computed  values  are 
given  in  Table  XXIII-6,  and  it  is  seen  that  the  values  agree  as  to  order  of 
magnitude  but  not  in  detail.  In  jpther  jvyords^  pur  assumptioji.jtha.t  the 
repulsive  potential  varies  inversely  as  a  power  of  r  is  not  very  accurate, 
and  to  do  better  one  would  have  .to.  use  a  function  with  an  extr^diappsable 
constant.  In  the  table  we  give  values  of  m,  and  it  is  seen  that  they  are 
in  the  neighborhood  of  9  for  most  of  the  crystals,  as  we  have  stated  earlier. 
Using  the  values  (3.17)  and  (3.18)  for  Pi  and  P2,  and  Eq.  (4.3),  we  at  once 
find 

7=^  +  1.  (4.4) 

Values  of  7  computed  in  this  way  are  tabulated  in  Table  XXIII-6,  and 
it  is  seen  that  the  agreement  with  the  value  found  from  the  thermal  expan- 
sion is  only  moderately  good,  much  poorer  than  that  found  with  values 
computed  by  Eq.  (4.3)  from  the  experimental  values  of  Pa/Pi.  In  other 
words,  if  we  had  a  theoretical  formula  for  the  repulsive  potential  which 
gave  a  better  value  for  the  change  of  compressibility  with  pressure  than 


SBC.  4] 


IONIC  CRYSTALS 


395 


the  inverse  power  function,  it  would  at  the  same  time  give  a  better  value 
for  the  thermal  expansion. 

In  the  preceding  paragraph,  we  have  seen  that  the  potential  energy 
curve  (3.9)  derived  from  theory,  gives  qualitative  but  not  vory  good  quan- 
titative agreement  with  experiment  for  the  change  of  compressibility  with 
pressure,  and  the  thermal  expansion.  From  Eq.  (3.16),  we  can  also  use 
this  curve  to  find  the  energy  of  the  crystal  at  the  absolute  zero  and  zero 
pressure,  C/oo.  The  negative  of  this  quantity  is  the  heat  of  dissociation 
of  such  a  crystal  into  ions,  which  we  may  call  D.  Of  course,  a  crystal 
would  not  really  dissociate  in  this  way  if  it  were  heated.  It  would 
dissociate  into  neutral  molecules,  for  example  of  NaCl,  or  possibly  into 
atoms  of  Na  vapor  and  molecules  of  CU,  instead.  Nevertheless,  thermo- 
chemical  measurements  are  available  from  which  we  can  get  experimental 
values  for  D.  We  may  imagine  that  we  go  from  the  crystal  to  the  ionized 
gas  in  the  following  steps,  each  of  which  is  understood  experimentally: 
(1)  we  vaporize  the  crystal,  obtaining  NaCl  molecules,  the  necessary 
energy  being  found  from  the  heat  of  vaporization,  which  has  been  meas- 
ured; (2)  we  dissociate  the  NaCl  molecules  into  atoms,  the  energy  being 
the  heat  of  dissociation  of  the  diatomic  molecule,  as  used  in  the  Morse 
curve;  (3)  we  ionize  the  Na  atom  to  form  a  positive  Na+  ion,  the  energy 
being  the  ionization  potential;  (4)  we  add  the  electron  so  obtained  to  the 
chlorine  atom,  obtaining  a  Cl~  ion,  releasing  an  amount  of  energy  that 
is  called  the  electron  affinity  of  the  chlorine  ion.  By  adding  the  amounts 
of  energy  required  for  all  these  processes,  we  find  experimental  values  that 


TABLE  XXI 11-7.—  LATTICE  KNEHGJES  OF  THE  ALKALI 
(Kilogrnm-enlorics  per  mole) 


Observed 

Calculated 

LiF    .       . 
LiCl 

240 
199 

238 
191 

LiHr  

NaCl  . 
NaBr        .                                     ... 

KF    .       .. 
KC1 

188 

183 
175 

190 
165 

180 

179 
169 

189 
163 

KBr  

159 

156 

RbBr  

154 

149 

Rbl  

145 

141 

The  observed  values  are  taken  from  Landolt-Bornstein's  Tables,  Dritter  Erg&nzungsband,  p.  2870, 
Dritter  Teil.  Calculated  values  are  found  by  Eq.  (3.16),  using  numerical  values  from  Tables  XXIII- 1 
and  XXIII-6. 


396 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 


•should  agree  with  the  values  calculated  from  Eq.  (3.16).  In  Table 
XXIII-7  we  give  a  number  of  these  values  and  the  calculated  ones.  The 
units  are  kilogram  calories  per  mole.  The  agreement  between  theory 
and  experiment  in  these  values  is  quite  striking  and  is  one  of  the  most 
satisfactory  results  of  the  theory  of  ionic  crystals.  It  is  not  hard  to  see 
why  the  agreement  here  is  so  much  better  than  in  the  calculation  of  change 
of  compressibility  with  pressure.  The  heat  of  dissociation  depends  on 
the  value  of  C/o  as  a  function  of  F,  while  the  compressibility  depends 
essentially  on  the  second  derivative  of  this  curve,  and  the  change  of 
compressibility  with  pressure  on  the  third  derivative.  It  is  a  well-known 
fact  that  differentiating  exaggerates  the  errors  of  a  curve  which  is  almost, 
but  not  quite,  correct.  It  thus  seems  likely  that  the  energy  of  Eq.  (3.16) 
is  really  quite  accurate,  but  that  its  second  and  third  derivatives  are 
slightly  in  error. 

6.  Other  Types  of  Ionic  Crystals. — In  the  preceding  sections  we  have 
been  talking  about  simple  binary  crystals,  formed  from  a  positive  and  a 
negative  ion  of  the  same  valency.  Of  course,  there  are  many  other  types 
of  ionic  crystals,  and  we  shall  not  take  up  the  other  sorts  in  nearly  such 
detail.  We  shall,  however,  list  a  number  of  crystal  structures,  with  the 

substances  crystallizing  in  them. 
First  we  may  mention  the  fluorite 
structure,  named  for  fluorite,  CaF2, 
which  crystallizes  in  it.  This  is  one 
of  the  simplest  crystals,  having 
twice  as  many  ions  of  one  sort  as 
of  the  other.  The  structure  is 
shown  in  Fig.  XXIII-5.  It  can  be 
considered  as  a  cube  of  calcium  ions, 
with  an  ion  at  the  center  of  each 
face  of  the  cube  as  well  as  at  the 
corners,  and  inside  this  a  cube  of 
8  fluorine  ions.  It  is  really  better, 
however,  to  consider  the  neighbors  of  each  ion.  As  we  see  from  the 
figure,  each  fluorine  is  surrounded  tetrahcdrally  by  4  calciums.  On 
the  other  hand,  each  calcium  is  at  the  center  of  a  cube  of  8  equally 
distant  fluorine  ions.  Thus  each  calcium  has  twice  as  many  fluorine 
neighbors  as  each  fluorine  has  calciums,  as  the  chemical  formula  demands. 
It  is  plain  that  molecules  have  no  more  independent  existence  in  such  a 
structure  than  they  do  in  sodium  chloride. 

In  Table  XXIII-8  we  give  the  crystals  that  exist  in  the  fluorite  struc- 
ture and  the  distances  between  nearest  neighbors.  In  addition,  we 
tabulate  the  sum  of  the  ionic  radii  of  Table  XXIII-2.  Though  these 
were  computed  from  binary  compounds  of  elements  of  equal  valency,  they 


FIG.  XXIII-5. — Fluorite  structure. 


5] 


IONIC  CRYSTALS 


397 


give  fairly  good  results  for  the  ionterionic  distances  even  in  these  rather 
different  compounds. 

TABLE  XXIII-8. — SUBSTANCES  CRYSTALLIZING  IN  FLTJORITE  STRUCTURE 


Substance 

Distance, 
angstroms 

Distance 
computed 

CaF2...    . 

2.36 

2.25 

SrF2  

2.50 

2.42 

SrCl2  

3.02 

2.92 

BaF2... 

2.68 

2.60 

CdF2  .  .  . 

2.34 

PbF2.  .  . 

2.57 

Ce02  

2.34 

PrO2.    .. 

2.32 

ZrO2.   .                                                    .      . 

2.20 

ThO2    .    . 

2.41 

Li20 

2.00 

2.2, 

Li2S.. 

2.47 

2  70 

Na2S.    . 

2.83 

2  95 

Cu2S. 

2.42 

Cu2Sc.   ... 

2.49 

In  addition  to  the  fluorite  structure,  there  are  a  number  of  other 
structures  assumed  by  similar  compounds.     We  shall  not  attempt  to 
enumerate  or  describe  them.     Some   of 
them  are  considerably  more  complicated 
than    the    fluorite    structure,    but    they 
resemble  it  in  that  there  is  no  semblance 
of  separate  molecules.    Each  positive  ion 
is  surrounded  by  a  number  of  negative 
ions  and  each  negative  by  a  number  of 
positives,  at  equal  distances,  so  that  it  is  in 

no  sense  correct  to  say  that  one  ion  is  FIO.  XXIII-G.  -  Tho  rairite 
bound  to  one  or  two  neighbors  more  than  ture- 

to  others. 

It  is  rather  interesting  to  consider  the  crystal  structure  of  substances 
containing  more  complicated  negative  ions.  Simple  examples  are 
nitrates,  sulphates,  and  carbonates.  These  are  all  similar  to  each  other 
in  that  the  negative  ion  exists  as  a  structure  by  itself,  like  an  ionized 
molecule,  while  the  positive  and  negative  ions  are  arranged  in  a  lattice 
without  suggestion  of  molecular  structure,  as  in  the  other  ionic  crystals. 
Thus  in  the  calcite  structure,  CaC03,  the  CO8 —  ion  exists  as  a  triangular 
structure,  with  the  carbon  in  the  middle,  the  oxygens  around  the  corners 
of  the  triangle.  This  structure  is  built  of  hexagonal  units,  as  shown  in 
Fig.  XXIII-6,  with  a  COa —  ion  at  the  center,  surrounded  by  six  Ca++ 


398 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIII 


ions.  Units  like  that  of  Fig.  XXIII-6,  and  others  which  are  the  mirror 
image  of  it  in  a  horizontal  plane  through  the  carbonate  ion,  are  built  up 
into  a  crystal.  The  substances  which  crystallize  in  this  structure  are 
tabulated  in  Table  XXIII-9.  Several  distances  are  necessary  to  describe 

TABLE  XXIII-9. — SUBSTANCES  CRYSTALLIZING  IN  CALCITE  STRUCTURE 


Substance 

C-0  distance, 
angstroms 

C-metal 
distance 

0-metal 
distance 

CaCO8  
MgCO3 

1.24 

3.21 
2  92 

2.37 

ZnCOa  

2.93 

MnCO3.   .                

1.27 

3  00 

2  14 

FeCO,..              .         ... 
NaNO3  

1  27 
1.27 

3.01 
3.25 

2.18 
2  40 

the  structure,  and  these  cannot  be  found  so  accurately  from  x-ray  methods 
as  in  the  simpler  crystals.  In  a  few  cases  they  are  not  known,  and  in  any 
case  they  are  not  very  certain.  We  tabulate  the  carbon  to  oxygen  (or 
nitrogen  to  oxygen)  distance,  giving  the  size  of  the  negative  ion,  and  also 
the  distances  from  positive  ion  to  carbon  and  oxygen.  We  see  that 
while  the  lattice  spacing  depends  on  the  positive  ion,  the  carbonate  or 
nitrate  ion  is  of  almost  the  same  size  in  each  case,  forming  a  practically 
independent  unit. 

The  sulphate  ion,  in  sulphates,  is  a  tctrahedral  structure,  with  the 
sulphur  in  the  center,  the  four  oxygens  at  the  corners  of  a  regular  tetra- 
hedron surrounding  it.  The  sulphur-oxygen  distance  is  about  1.40  A 
in  all  the  compounds.  Examples  are  CaS04  and  BaSC>4.  These  form 
different  lattices,  rather  complicated,  but  as  we  should  expect  they  are 
structures  formed  of  positive  metallic  ions  and  negative  sulphate  ions, 
each  ion  being  surrounded  by  a  number  of  ions  of  the  opposite  sign. 
There  are  a  number  of  other  compounds  crystallizing  in  the  BaSO4  struc- 
ture: BaS04,  SrS04,  PbS04,  (NH4)C104,  KC104,  RbClO4,  CsClO4,  T1C104, 
KMnO4. 

6.  Polarizability  and  Unsymmetrical  Structures. — In  discussing  the 
energy  of  ionic  crystals,  we  have  assumed  that  the  only  forces  acting  were 
electrostatic  attractions  and  repulsions,  and  the  repulsions  on  account 
of  the  finite  sizes  of  ions.  But  under  some  circumstances  there  can  also 
be  forces  and  changes  of  energy  arising  from  the  polarizability  of  the 
ions.  Of  course,  just  as  in  Chap.  XXII,  we  can  have  Van  der  Waals 
attractions  between  ions,  but  this  is  ordinarily  a  small  effect  compared 
to  the  electrostatic  attraction  and  can  be  neglected.  There  can  be 
other,  larger  effects  of  polarizability,  however.  We  remember  that 
according  to  Sec.  3  of  Chap.  XXII,  an  atom  or  ion  in  an  electric  field  E 


SBC.  6]  IONIC  CRYSTALS  399 

acquires  a  dipole  moment  aE,  where  a  is  the  polarizability  of  the  ion. 
Furthermore,  the  force  on  the  resulting  dipole  is  equal  to  the  dipole 
moment,  times  the  rate  of  change  of  electric  field  with  distance.  This 
force  and  the  resulting  term  in  the  energy  can  be  large  if  a  polarizable  ion 
is  in  an  external  field,  such  as  can  arise  from  other  ions.  Now  in  most 
of  the  structures  we  have  considered,  this  does  not  occur.  In  the  sodium 
chloride,  caesium  chloride,  zincblende,  wurtzite,  and  fluorite  structures, 
each  ion  is  surrounded  by  ions  of  the  opposite  sign  in  such  a  symmetrical 
way  that  the  electric  field  at  each  ion  is  zero,  so  that  it  is  not  polarized. 
But  the  calcite  and  barium  sulphate  structures  are  quite  different.  There 
the  oxygens  in  the  carbonate  or  sulphate  ions  are  by  no  means  surrounded 
symmetrically  by  other  ions,  and  there  is  a  strong  field  acting  on  them. 
Furthermore,  they  are  very  polarizable,  and  the  result  is  a  large  added 
attraction  between  the  parts  of  the  complex  ion.  Adopting  an  ionic 
picture  of  the  structure  of  the  C03 —  and  other  ions,  we  should  have  it 
made  of  C++++  and  three  O  's,  giving  the  net  charge  of  two  negative 
units.  Similarly  NOa"  would  be  made  of  N+f~+H~  and  three  O — X  and 
SC>4  of  S+4"f"l"f  +  and  four  0  's.  Each  of  the  ions,  in  these  cases,  would 
form  a  closed  shell,  the  carbon  and  nitrogen  being  like  helium,  the  sulphur 
and  oxygen  like  neon.  The  central  ion  of  the  complex  ion  in  each  case 
would  be  very  strongly  positively  charged  and  would  polarize  the  oxygen 
very  strongly,  adding  greatly  to  its  attraction  to  the  carbon,  nitrogen,  or 
sulphur.  We  shall  not  try  to  estimate  the  effect  of  this  added  attraction 
at  present,  but  we  can  easily  get  evidence  of  it.  Thus  we  have  mentioned 
that  the  sulphur-oxygen  distance  in  the  sulphates  is  about  1.40  A.  On  the 
other  hand,  the  0~  radius,  from  Table  XXIII-2,  is  about  1 .45  A.  Of 
course,  S~H~f+~H~  would  have  an  extremely  small  radius,  but  still  we  should 
expect  that  the  sulphur-oxygen  distance  would  be  something  like  1.50  A 
in  the  absence  of  extra  attraction.  Even  more  striking  is  the  carbon- 
oxygen  distance  in  the  carbonates,  about  1.27  A,  well  below  the  ionic 
radius  of  oxygen  alone.  These  facts  suggest  that  some  additional  attrac- 
tion is  acting  between  the  ions  in  question,  decreasing  the  distance  of 
separation.  One  way  of  interpreting  this  added  attraction  is  the  polariza- 
tion effect  we  have  mentioned.  In  the  next  chapter,  we  shall  see  that 
another  interpretation  is  to  suppose  that  the  ions  are  not  really  as  highly 
charged  as  the  ionic  picture  would  suggest,  but  that  in  addition  there  are 
homopolar  bonds  between  the  atoms  making  up  these  complex  ions.  The 
homopolar  binding  would  in  this  interpretation  furnish  the  extra  attrac- 
tive force  resulting  in  the  small  spacing  between  atoms.  Thus  we  are 
not  perhaps  forced  to  think  about  polarizability  at  all  in  such  a  case. 
There  are  many  cases,  however,  where  it  is  definitely  important,  and 
the  effect  of  polarization  can  be  calculated  easily  from  known  polariza- 
bilities  and  charge  distributions. 


CHAPTER  XXIV 
THE  HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS 

Ionic  compounds,  as  we  have  seen  in  the  preceding  chapter,  exist  most 
characteristically  in  crystalline  solids,  for  the  electrostatic  forces  that  hold 
them  together  extend  out  in  all  directions,  binding  the  ions  together  into  a 
structure  that  has  no  trace  of  molecular  formation.  Compounds  held 
together  by  homopolar  bonds,  on  the  contrary,  form  definitely  limited 
molecules,  which  are  bound  to  each  other  only  by  the  relatively  weak 
Van  der  Waals  forces.  Thus  their  most  characteristic  form  is  the  gaseous 
phase  in  which  the  molecules  have  broken  apart  from  each  other  entirely. 
The  ordinary  gases  with  which  we  are  familiar,  and  the  ordinary  liquids, 
belong  to  this  group  of  compounds.  They  arc  the  only  group  to  which  the 
idea  of  the  molecule,  so  common  in  chemistry,  really  applies.  We  shall 
begin  our  discussion  by  taking  up  some  of  the  familiar  molecular  com- 
pounds and  discussing  the  homopolar  bonds  which  hold  their  atoms 
together,  and  the  nature  of  homopolar  valence.  Then  we  shall  go  on  to  a 
discussion  of  tho  Van  der  Waals  constants  of  these  substances,  as  indicat- 
ing their  behavior  in  the  gaseous  and  liquid  phases,  and  finally  we  shall 
take  up  the  solid  forms  of  the  homopolar  substances. 

1.  The  Homopolar  Bond.  —  The  Qrincipa^glements  sometimes  form- 

Te:  H.  F.  CL  Br. 


In  these  groups  of  elements  we  must  add  four,  three,  two,  or  one  electron 
respectively  to  form  a  closed  shell.  But  if  two  such  elements  combine 
together,  where  are  the  extra  electrons  to  come  from?  There  is  no 
positive  ion  losing  electrons  and  ready  to  donate  them  to  help  form  nega- 
tive ions.  The  expedient  which  these  elements  adopt  in  their  effort  to 
form  closed  shells  is  the  sharing  of  electrons,  as  we  have  discussed  in 
Chap.  XXII.  An  electron  can  sometimes  be  held  by  two  atoms  in  com- 
mon, spending  part  of  its  time  on  one,  part  on  the  other,  and  part  in  the 
region  between;  in  so  doing  it  helpslfill  UP  the  shells  of  both  atoms'.  There 
is  just  one  conspicuous  rule  that  holds  for  almost  all  such  bonds,  and 
that  is  tbftt  ordinarily  two  electrons  are  shared  in  a  similar  way,  the  two 
together  forming  what  is  called  a  homopolar  or  electron-pair  bond.  The 
reason  why  two  cooperate,  as  we  saw  in  Chap.  XXII,  is  essentially  the 
electron  spin  in  conjunction  with  the  exclusion  principle. 

In  the  first  place,  we  can  symbolize  the  process  of  forming  a  homopolar 
bond  by  a  simple  device  used  by  G.  N.  Lewis,  to  whom  many  of  the  ideas 
of  homopolar  binding  are  due.  Most  of  the  elements  forming  this  type  of 
bond  are  trying  to  complete  a  shell,  or  subshell,  of  eight  electrons,  as  we 

400 


SEC.  1]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  401 

have  explained  in  Chap.  XXII.  Lewis  indicates  the  eight  electrons  by 
eight  dots  surrounding  the  symbol  of  the  element.  Thus,  for  instance,  the 
neutral  fluorine  atom,  which  has  only  seven  electrons  in  its  outer  shell, 

would  be  symbolized  as  :  F . .    This  does  not  have  a  completed  shell.     But 

by  combining  two  such  atoms,  we  can  form  the  structure  :F:F:,  contain- 
ing fourteen  electrons  but  sharing  two  of  them  in  a  homopolar  bond,  so 
that  each  atom  in  a  sense  has  a  completed  shell.  It  is  clear  from  this 
symbolization  that  the  halogons,  F,  Cl,  Br,  I,  can  form  one  homopolar 
bond;  the  divalent  elements  0,  S,  Se,  Te,  can  form  two;  and  so  on.  In 
this  symbolization,  hydrogen  takes  a  special  place,  for  by  adding  electrons 
it  forms  a  completed  helium  shell  of  two  electrons.  It  thus  forms  one 
homopolar  bond,  and  in  this  type  of  bonding  it  is  in  many  respects  analo- 
gous to  a  halogen.  In  such  a  way,  for  instance,  we  can  indicate  the 

structure  of  hydrogen  chloride  as  H :  Cl : ,  the  electron  pair  sharod  between 

the  two  atoms  helping  to  fill  up  the  hydrogen  shell  of  two,  and  the  chlorine 
shell  of  eight.  Similarly  and  illustrating  also  the  valences  of  0,  N,  and 

C,  we  may  write  water,  ammonia,  and  methane  respectively  as  H:O:, 

H 
H 

H :  N :  H,  and  H :  C :  H.     In  each  of  these  compounds,  we  observe  that  the 

H  H 

total  number  of  electrons  indicated  is  just  the  number  furnished  by  the 
outer  shells  of  the  atoms  entering  into  the  molecule.  Thus,  in  NH3 
the  nitrogen  furnishes  five  electrons,  each  of  the  hydrogens  three,  making 
eight  in  all. 

Hydrogen  is  in  a  very  special  position,  in  that  it  forms  a  closed  shell 
(heliumlike)  by  adding  an  electron,  and  also  a  closed  shell  (the  nucleus 
without  any  electrons)  by  losing  an  electron.  It  can  act,  in  the  language 
of  ions,  like  either  a  univalent  positive  or  a  univalent  negative  element. 
This  gives  the  possibility  of  an  ionic  interpretation  of  most  of  the  hydrogen 

compounds.     Thus  we  may  symbolize  hydrogen  chloride  as  H+(:C1:)~, 

and  water  as  H+(:6:)—  H+,  or  as  H+(:6:)~.    We  shall  see  later  that 

H 

there  is  good  reason  to  think,  however,  that  in  most  of  these  cases  the 
homopolar  way  of  writing  the  compound  is  nearer  the  truth  than  the  ionic 
method. 

The  elements  carbon,  nitrogen,  and  oxygen  have  a  peculiarity  rarely 
shown  by  other  elements:  they  form  sometimes  what  is  called  a  double 


402  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHA*.  XXIV 

bond,  and  in  the  cases  of  carbon  and  nitrogen  a  triple  bond.  This  means 
that  two  or  three  pairs  of  electrons,  rather  than  one,  may  be  shared 
between  a  pair  of  atoms.  This  is  seen  in  its  simplest  form  in  the  molecules 
O2  and  N2.  If  two  oxygens  shared  only  one  pair  of  electrons,  they  would 
not  achieve  closed  shells;  they  must  share  two  pairs,  so  that  two  electrons 
of  each  atom  count  in  the  shell  of  the  other  as  well.  Similarly  two  nitro- 
gen atoms  must  share  three  pairs.  We  can  symbolize  these  compounds 

by  :  O : :  O : ,  where  both  pairs  of  electrons  between  the  O's  are  counted 
in  each  group  of  eight,  and  by  :N: :  :N:.  Compounds  having  double  or 
1  riple  bonds  generally  have  a  rather  unsaturated  nature;  they  tend  to  add 
more  atoms,  breaking  down  the  bonds  into  single  ones  and  using  the 
valences  left  over  in  order  to  attach  the  other  atoms.  A  familiar  example 
is  the  group  of  compounds  acetylene  C2H2,  ethylene  C2H4,  and  ethane 
C2H6,  in  which  the  last  named  is  the  most  stable.  These  are  symbolized 

HH 
H:C:::C:H,  ij-C:;C-H>  H:C:C:H,  formed  with  triple,  double,  and 

HH 
single  bonds  respectively. 

One  can  derive  a  good  deal  of  information  about  the  three-dimensional 
structure  of  a  molecule  in  space  from  the  nature  of  the  homopolar  bonds, 
and  it  must  not  be  supposed  that  the  chemical  formulas,  written  as  we 
have  been  writing  them  in  a  plane,  express  the  real  shape  of  the  molecule. 
We  have  written  them  in  each  case  so  as  to  approximate  the  shape  as 
closely  as  possible,  but  in  many  cases  have  not  succeeded  very  well.  In 
general,  the  four  pairs  around  an  atom  tend  to  be  arranged  in  the  only 
symmetrical  way  they  can  be,  namely  at  the  four  corners  of  a  regular 
tetrahedron.  The  vectors  from  the  center  to  the  corners  of  a  tetrahedron 
form  angles  of  109.5°  with  each  other,  often  called  the  tetrahedral  angle, 
and  in  a  great  many  cases  it  is  found  that  when  two  or  more  atoms  are 
attached  to  another  atom  by  homopolar  bonds,  the  lines  of  centers 
actually  make  approximately  this  angle  with  each  other.  We  shall  dis- 
cuss this  more  in  detail  in  the  next  section,  in  which  we  take  up  the  struc- 
tures of  a  number  of  homopolar  molecules. 

2.  The  Structure  of  Typical  Homopolar  Molecules. — Many  of  the 
homopolar  molecules  are  among  the  most  familiar  chemical  substances. 
In  this  section  we  shall  describe  a  few  of  them,  discussing  the  nature  of 
their  valence  binding  and  giving  information  about  their  shape  and  size. 
For  the  diatomic,  and  some  of  the  polyatomic,  molecules,  this  information 
is  derived  from  band  spectra.  In  other  cases,  it  is  found  by  x-ray  diffrac- 
tion studies  of  the  solid,  using  the  fact  that  homopolar  molecules  generally 
are  very  similar  in  the  solid  and  gaseous  phases,  or  by  electron  diffraction 
with  the  gas.  We  begin  with  some  of  the  diatomic  molecules  listed  in 
Table  IX-1,  including  H2,  C12,  Br2,  I2,  NO,  02,  N2,  CO,  HC1,  and  HBr. 


SBC.  2]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  403 

The  first  molecule  in  the  list,  and  the  simplest  diatomic  molecule,  is 
hydrogen,  H2.  Its  structure  of  course  is  H:H,  the  two  electrons  being 
shared  to  simulate  a  helium  structure  about  each  atom.  The  internuclear 
distance1  0.75  A  is  the  smallest  internuclear  distance  known  for  any 
compound,  as  is  natural  from  the  small  number  of  electrons  in  hydrogen. 
As  we  have  seen,  hydrogen  acts  a  little  like  a  halogen  when  it  forms  homo 
polar  bonds,  and  we  might  consider  next  the  molecules  F2,  C12,  Br2,  I2. 
We  have  already  mentioned  their  bonding,  by  a  single  electron  pair 
bond,  in  the  previous  section.  The  internuclear  distances  in  these 
molecules  are  large,  being  1.98  A  for  C12,  2.28  A  for  Br2,  and  2.66  A  for 
I2,  as  we  saw  in  Table  IX-1.  It  is  interesting  to  compare  these  inter- 
nuclear distances  with  the  ionic  radii,  from  Table  XXIII-2.  There  we 
found  radii  of  1.80  A,  1.95  A,  and  2.20  A  for  C1-,  Br~,  and  I~.  If  these 
radii  represented  the  sizes  of  the  atoms  in  the  diatomic  molecules,  the 
internuclear  distances  would  be  twice  the  radii,  or  3.60  A,  3.90  A,  and 
4.40  A,  almost  twice  the  observed  distances.  This  is  an  illustration  of  the 
fact,  which  proves  to  be  quite  general,  that  interatomic  distances  in 
homopolar  binding  are  decidedly  less  than  in  ionic  binding.  The  reason  is 
simple.  While  the  sharing  of  a  pair  of  electrons  is  in  a  sense  a  way  of 
building  up  a  closed  shell  of  electrons,  still  the  shell  is  really  not  filled  to 
capacity.  There  is,  so  to  speak,  a  soft  place  in  the  shell  just  where  the 
bond  is  located,  and  the  atoms  tend  to  pull  together  closer  than  if  the  shell 
were  really  filled. 

The  remaining  molecules  in  our  list  are  NO,  02,  N2,  CO,  HC1,  and 
HBr.  The  first  of  these,  NO,  is  the  most  peculiar  compound  in  the  list 
and  one  of  the  most  peculiar  of  the  known  compounds.  We  note  that 
nitrogen  supplies  five,  and  oxygen  six,  outer  electrons  to  the  compound, 
making  a  total  of  eleven,  an  odd  number.  It  is  quite  obvious  that  an  odd 
number  of  electrons  cannot  form  closed  shells,  electron  pairs,  or  anything 
else  associated  with  stable  molecules.  As  a  matter  of  fact,  out  of  all  the 
enormous  number  of  known  chemical  compounds,  only  a  handful  have  an 
odd  number  of  electrons,  and  NO  is  almost  the  only  well-known  one  of 
these.  We  shall  make  no  effort  to  explain  it  in  terms  of  ordinary  valence 
theory,  for  it  is  in  every  way  an  exception,  though  it  can  be  understood 
in  terms  of  atomic  theory. 

Oxygen,  with  a  double  bond,  and  nitrogen,  with  a  triple  one,  have 
already  been  discussed.  The  internuclear  distances,  from  band  spectra, 
are  1.20  A  for  oxygen,  and  1.09  A  for  nitrogen.  In  line  with  what  we  have 
just  said  about  the  halogens,  it  is  interesting  to  notice  that  the  inter- 
nuclear distance  in  oxygen  is  a  great  deal  less  than  the  double  radius  of 
O — .  That  ionic  radius  was  1.45  A,  so  that  if  it  represented  the  size  of 

1  See  Sponer,  "  Moleklilspektren  und  ihre  Anwendungen  auf  chemische  Probleme," 
Vol.  I,  Springer,  1935,  for  interatomic  distances  of  diatomic  and  polyatomic  molecules 
in  this  chapter. 


404  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 

the  oxygen  in  O2,  the  internuclear  distance  would  be  2.90  A,  more  than 
twice  the  actual  distance.  In  the  case  of  02,  with  its  double  bond,  the 
tendency  of  the  atoms  to  pull  together  in  homopolar  binding  is  particu- 
larly pronounced,  for  the  shell  is  even  less  nearly  filled  than  with  a  single 
bond  and  can  be  even  more  compressed  by  the  interaction  forces.  In 
nitrogen  with  its  triple  bond,  the  interatomic  distance  is  even  smaller,  in 
line  with  this  fact.  The  next  molecule  on  the  list,  CO,  does  not  fit  in  very 
well  with  our  rules.  A  clue  to  its  structure  is  provided  by  the  fact  that 
it  has  4  +  6  ==  10  outer  electrons,  just  like  N2,  and  that  in  many  of  its 
properties  it  strongly  resembles  N2.  We  have  stated  that  the  internuclear 
distance  in  nitrogen  was  1.09  A;  in  CO  it  is  1.13  A.  The  suggestion  is 
very  natural  that  in  a  sense  a  triple  bond  is  formed  in  this  case  also,  with 
the  structure  :  C : : :  0 : ,  though  a  triple  bond  is  not  usually  formed  by 
oxygen. 

We  have  already  discussed  the  structure  of  the  next  two  molecules, 
HC1  and  HBr,  whose  valence  properties  are  indicated  by  the  symbols 

H:C1:,  H:Br:.  The  internudear  distances  are  1.27  A  and  1.41  A  respec- 
tively. We  note,  as  before,  how  much  smaller  these  are  than  the  values 
given  by  ionic  radii.  We  have  no  ionic  radius  for  H,  but  for  Cl~  we  have 
1.80  A,  and  for  Br~  the  distance  is  1.95  A.  The  internuclear  distances  in 
these  cases  are  actually  less  than  the  ionic  radii.  This  is  good  evidence 
for  the  homopolar,  rather  than  the  ionic,  nature  of  these  compounds. 
Another  reason  comes  from  the  magnitudes  of  the  electric  dipole  moments 
of  these  two  molecules,  which  are  found  to  be  1.03  X  10~18  and 
0.78  X  10~18  e.s.u.-cm.  If  the  molecules  were  really  ionic,  we  should 
expect  that  electrically  they  would  consist  of  a  unit  positive  charge  on 
the  hydrogen  nucleus,  and  a  net  negative  charge  of  one  unit,  spherically 
symmetrical,  and  therefore  acting  as  if  it  were  on  the  chlorine  or  bromine 
nucleus.  That  is,  there  would  be  charges  of  one  electronic  unit  located 
1.27  A  and  1.41  A  apart  respectively.  This  would  give  dipole  moments, 
equal  to  the  product  of  charge  and  displacement,  of 

(4.8  X  10-10)  X  (1.27  X  10~8)  =  6.1  X  10~18, 

and  of  6.8  X  10~~18  units,  respectively.  The  observed  dipole  moment  of 
HC1,  as  we  have  seen,  is  only  about  one-sixth  of  this  value,  and  of  HBr 
about  one-ninth  of  the  value  given  by  the  polar  model.  To  explain  this, 
we  must  assume  that  the  negative  charge  is  not  located  symmetrically 
about  the  Cl  or  Br  but  is  displaced  toward  the  hydrogen.  This  is  what  we 
should  expect  if  there  is  really  a  homopolar  bond,  for  then  the  shared 
electrons  would  be  in  the  neighborhood  of  the  hydrogen,  displaced  in 
that  direction  from  the  halogen  ion.  The  dipole  moments  then  furnish 
arguments  for  the  homopolar  nature  of  the  bond  and  for  thinking  that  it 


SBC.  2]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  405 

is  less  polar  in  HBr  than  in  HC1.  There  is  still  another  way  of  regarding 
this  situation :  we  may  make  use  of  the  polarizability  of  the  halide  ion.  A 
hydrogen  nucleus  close  to  a  halide  ion  would  produce  an  extremely  strong 
electric  field,  which  would  polarize  the  ion,  changing  it  into  a  dipole  with 
the  negative  charge  displaced  slightly  toward  the  positive  hydrogen  ion. 
This  dipole  moment  would  tend  to  cancel  the  moment  produced  by  the 
two  undistorted  ions,  so  that  the  net  moment  would  be  less  than  the  figure 
6.1  X  10~18  calculated  above.  As  a  matter  of  fact,  calculations  by 
Debye1  show  that  the  resulting  dipole  moment  calculated  in  this  way 
would  be  of  the  right  order  of  magnitude.  We  should  not  regard  this 
calculation  as  indicating  that  our  homopolar  shared  electron  theory  is  not 
accurate,  however.  For  in  this  case  the  polarizability  is  simply  a  rather 
crude  way  of  taking  account  of  the  shifting  of  electric  charge  which  we 
can  describe  more  precisely  as  electron  sharing  between  the  atoms.  The 
situation,  however  we  describe  it,  is  essentially  this:  that  the  electrons, 
instead  of  being  arranged  in  a  spherically  symmetrical  way  about  the 
halogen  nucleus,  tend  to  be  somewhat  displaced  toward  the  hydrogen,  so 
that  it  also  is  partly  surrounded  by  electrons,  rather  than  acting  as  an 
entirely  isolated  ion. 

We  have  now  completed  our  list  of  diatomic  molecules.  Next  wo 
might  well  take  up  various  hydrides:  H2O,  NH3,  CH4,  H2S,  PH3,  SiH4. 
We  have  already  given  structural  formulas  for  H2O,  NH3,  and  CH4;  the 
others  are  analogous,  H2S  resembling  H2O,  PH3  being  like  NH3,  and  SiH4 
like  CH4.  The  hydrogens  are  bound,  on  the  homopolar  theory,  by 
single  bonds,  and  the  angles  made  by  the  radius  vectors  to  different 
hydrogens  are  very  closely  the  tetrahedral  angle  109.5°.  Thus  H2O  is  a 
triangular  molecule,  the  hydrogen-oxygen  distance  being  0.96  A,  and  tho 
H-O-H  angle  being  104.6°.  NH3  is  pyramidal,  the  nitrogen-hydrogen 
distance  being  1.01  A  and  the  H-N-H  angle  109°.  CH4  is  tetrahedral,  the 
carbon-hydrogen  distance  being  1.1  A.  H2S  is  triangular,  like  water,  the 
sulphur-hydrogen  distance  1.35  A  and  the  angle  92.1°,  rather  smaller 
than  we  should  have  supposed.  PH3  is  presumably  pyramidal  like  NH3 
but  the  distances  and  angles  do  not  seem  to  be  known.  SiH4  is  tetra- 
hedral but  again  the  distances  are  not  known. 

There  are  only  a  few  other  common  inorganic  molecules  to  be 
mentioned.  CC>2  is  a  linear  structure,  with  valences  symbolized  by 

:O:  :C:  :O:.  That  is,  there  are  double  bonds  between  the  carbon  and 
each  oxygen.  The  C-0  distance  is  1.16  A,  slightly  greater  than  the 
value  1.13  A  found  in  CO,  where  there  is  a  triple  bond.  N2O  is  also  a 

linear  molecule,  presumably  with  the  structure  :  N : :  N : :  0 : ,  again  formed 
with  double  bonds  like  CO2,  which  has  the  same  number  of  outer  electrons. 

1  See  Debye,  "Polare  Molekel,"  Sec.  14,  Hirzel,  1929. 


406  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 

The  distance  between  end  atoms  is  2.38  A,  slightly  greater  than  the  value 
2  X  1.18  =  2.32  A  between  end  atoms  in  C02.  Here  again  we  see  the 
resemblance  between  N2  and  CO,  in  that  they  form  similar  molecules 
when  another  oxygen  atom  is  added.  The  molecule  802  is  a  triangular 
molecule  shaped  something  like  water.  Its  structure  presumably  is 

:0:S:.    This  is  the  first  example  we  have  met  of  a  case  where  an  atom 


is  not  surrounded,  even  with  the  aid  of  shared  electrons,  by  a  closed 
shell:  the  sulphur  has  only  six  electrons  around  it.  We  shall  come  to 
other  examples  later,  when  we  talk  about  inorganic  radicals.  The 
sulphur-oxygen  distance  in  802  is  1.37  A.  Carbon  disulphide  CS2 
resembles  CO2  in  being  a  linear  molecule,  with  carbon-sulphur  distance  of 
1.6  A,  decidedly  larger  than  the  value  1.16  A  in  C02,  in  accordance  with 
the  fact  that  sulphur  is  a  larger  atom  than  oxygen. 

The  molecules  taken  up  so  far  are  all  very  simple,  composed  of  very 
few  atoms.  The  more  complicated  examples  of  homopolar  molecular 
compounds  are  found  almost  entirely  in  the  field  of  organic  chemistry. 
We  shall  postpone  a  discussion  of  organic  compounds  until  the  next 
chapter,  since  they  form  a  field  by  themselves.  Before  closing  our  discus- 
sion, however,  we  shall  take  up  a  different  sort  of  homopolar  structure, 
namely,  a  few  inorganic  negative  ions,  formed  very  much  like  molecules. 
The  most  important  ones  are  NOa~,  COa  ,  804  ,  C104~,  mentioned 
in  the  preceding  chapter.  The  first  two,  as  stated  in  Chap.  XXIII,  Sec. 
5,  are  triangular  structures  in  a  plane,  the  N  or  C  being  in  the  center,  the 
oxygens  at  the  corners.  Each  has  24  electrons  (when  we  take  account 
of  the  negative  charge  011  the  ion),  so  that  it  was  possible  in  the  last 
chapter  to  treat  them  as  ionic  structures,  the  oxygen  having  a  closed 
shell  of  eight  electrons,  the  nitrogen  or  carbon  having  no  outer  electrons. 

..       :6:  ..      :6: 

A  structure  much  nearer  the  truth,  however,  is  :O:N  \\    and  :0:C  *.'.  . 

"        :0:  "      :0: 

These  structures  differ  from  the  ionic  one  in  that  we  have  indicated  two 
electrons  from  each  oxygen  as  being  shared  with  the  central  nitrogen  or 
carbon.  This  case  resembles  that  of  SO2  in  the  preceding  paragraph,  in 
that  one  of  the  atoms,  in  this  case  the  central  one,  has  only  six  rather 
than  eight  electrons  surrounding  it.  Another  molecule  showing  similar 
structure  is  SO3.  We  have  stated  in  the  preceding  chapter,  Table 
XXIII-9,  that  the  C-0  or  N-O  distance  in  the  carbonates  and  nitrates  is 
about  1.27  A.  This  is  decidedly  greater  than  the  C-0  distance  in  CO, 
which  is  1.13  A,  and  in  C02,  1.16  A,  but  in  the  present  case  there  is  only 
a  single  bond,  rather  than  the  triple  or  double  bonds  found  in  those  two 


SEC.  3]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  407 

compounds.  The  agreement  is  close  enough  so  that  it  is  quite  plain  that 
the  bonds  in  these  cases  are  homopolar  and  not  ionic,  as  was  stated  in 
Chap.  XXIII,  Sec.  6.  In  the  sulphate  and  perchlorate  ions,  S04~  and 
C1O4~,  there  are  enough  electrons,  32  outer  ones  per  ion,  to  form  com- 
plete shells  around  the  central  atoms.  The  valence  can  be  symbolized 


:  O  :  S  :  O  :  and  :  O  :  Cl  :  O  :  ,  and  the  compounds  can  be  described  as  having 
":6:"  "  :6:" 

single  valence  bonds  between  the  central  atom  arid  each  oxygen.  They 
have  a  tetrahedral  form,  the  sulphur-oxygen  distance  in  the  sulphates 
being  about  1.40  A. 

3.  Gaseous  and  Liquid  Phases  of  Homopolar  Substances.  —  We  have 
already  mentioned  that  the  gaseous  phase  is  the  most  characteristic  one 
for  molecular  substances  with  homopolar  binding.  We  shall  begin, 
therefore,  by  examining  the  Van  dor  Waals  constants  a  and  6  for  a  con- 
siderable number  of  homopolar  substances.  The  constant  b  will  give  us 
information  about  the  dimensions  of  the  molecules,  information  that  we 
can  correlate  with  the  known  interatomic  distances,  and  a  will  lead  to 
information  about  the  strength  of  the  Van  der  Waals  attraction. 

In  Table  XXIV-1,  we  give  Van  der  Waals  constants  for  quite  a  series 
of  gases,  arranged  in  order  of  their  &'s.  We  include  not  merely  the  gases 
mentioned  in  the  preceding  section,  but  the  inert  gases,  for  comparison, 
and  then  a  considerable  number  of  organic  substances,  which  as  we  have 
mentioned  furnish  the  largest  and  most  characteristic  group  of  homopolar 
substances.  In  the  table,  the  units"  of  a  are  (dynes  per  square  centimeter) 
times  (cubic  centimeters  per  mole)2.  The  units  of  6  are  cubic  centimeters 
per  mole.  These  constants  are  obtained  from  the  critical  pressure  and 
temperature  by  Eq.  (2.5),  Chap.  XII.  They  do  not,  therefore,  have  just 
the  same  significance  as  the  a  and  6  of  Eq.  (5.3),  Chap.  XII,  for  those  are 
the  constants  that  would  lead  to  agreement  between  Van  der  Waals' 
equation  and  experiment  at  low  density,  while  the  values  we  use  are  suit- 
able to  pressures  and  temperatures  around  the  critical  point,  which  will 
disagree  with  the  other  ones  unless  Van  der  Waals'  equation  is  really 
applicable  over  the  whole  range  of  pressures  and  temperatures.  We 
note  from  Eq.  (2.6)  of  Chap.  XII,  that  if  Van  der  Waals'  equation  were 
correct,  we  should  have  Vc/3  =  6,  where  Ve  is  the  observed  critical 
volume.  To  test  this  relation,  Table  XXIV-1  lists  observed  values  of 
Fc/3.  We  see  that  these  values  do  not  agree  very  closely  with  the  values 
of  6,  as  was  stated  in  Chap.  XII,  though  they  are  not  widely  different.  In 
addition  to  these  quantities,  we  also  tabulate  the  molecular  volume  of 
the  liquid,  in  cubic  centimeters  per  mole,  for  the  lowest  temperature  for 
which  figures  are  available,  and  also  the  dipole  moment  for  dipole  mole- 


408  INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXIV 

TABLE  XXI V-l. — VAN  DER  WAALS  CONSTANTS  FOB  IMPERFECT  GASES 


Gas 

Formula 

a 

ft 

TV3 

Molec- 
ular vol- 
ume of 
liquid 

Electric 
moments 

Neon           

Ne 

0  21  X  1012 

17   1 

14  7 

16  7 

0  X  10~» 

Helium 

He 

0  035 

23  f> 

20  5 

27  4 

o 

Hydrogen 
Nitric  oxide                  .... 
Water                  .... 
Oxygen                                    . 

H2 
NO 
H2O 
O2 

0  25 
1  36 
5.53 
1  40 

26  5 
27  8 
30  4 
32  2 

21  6 
19.1 
18.9 
24  8 

26  4 
23  7 
18.0 
25  7 

0 
1.85 

o 

Argon           
Ammonia         
Nitrogen 

A 
NHs 

N2 

1.36 
4  22 
1   36 

32.2 
36.9 
38  3 

26  1 
24.2 
30  0 

28  1 
24.5 
32  8 

0 
1.44 

o 

Carbon  monoxide 

CO 

1  50 

39  7 

30  0 

32  7 

0  10 

Krypton 

Kr 

2  35 

39  7 

36  0 

38  9 

o 

Hydrogen  chloride 

HC1 

3  72 

40  7 

20  8 

30  g 

1  03 

Nitrous  oxide             ... 
Carbon  dioxide            ..... 

N;O 
COs 

3  61 
3.64 

41.1 
42  5 

32.3 
32  8 

44  0 
41  7 

0.25 
0 

Methane 

CH4 

2  28 

42  6 

32  9 

49  5 

o 

Hydrogen  sulphide   . 
Hydrogen  bromide  

H2S 
HBr 

4  49 
4  51 

42.7 
44   1 

35.4 
37  5 

0.93 
0  78 

Xe 

4.15 

50  8 

38  0 

47  5 

0 

Acetylene   .          .... 
Phosphtne                 .... 

C2H2 
P1I3 

4  43 
4  69 

51.3 
51  4 

37.5 
37  7 

50.2 
49  2 

0 
0  55 

Chlorine       

Ch 

6.57 

56  0 

41  0 

41  .2 

0 

Sulphur  dioxide 

SOa 

6  80 

56   1 

41  0 

43  8 

1  61 

Ethylene                

C2H4 

4  46 

56  1 

42  3 

49  3 

o 

Silicon  hydride        

SiH4 

4.38 

57  6 

47 

0 

Methylamine     
Ethane       .          

CHaNH2 
CHa  —  CHs 

7  23 
5.46 

59.6 
63  5 

47  6 

44.5 
54  9 

1.31 
0 

Methyl  chloride          
Methyl  alcohol 
Methyl  ether       
Carbon  bisulphide     

CHaCl 
CHaOH 
(CH3)20 

CS2 

7.56 
9  65 
8.17 
11  75 

64.5 
66.8 
72  2 
76  6 

45.4 
39.0 

67  5 

49  2 
40.1 

59  0 

1.97 
1.73 
1.29 

Dimethylamine       
Propylene        

(CHa)aNH 
CsHe 

9  77 
8.49 

79.6 
82  4 

66  2 
69  0 

o 

Ethyl  alcohol 

CaHsOII 

12  17 

83  8 

41  0 

57  2 

1  63 

Propane             

CHs—CHz  —  CHd 

8.77 

84   1 

75  3 

o 

Chloroform     .         

CHCh 

15.38 

102 

77   1 

80.2 

1.05 

Acetic  acid          .             . 
Trimethylamine              .  .    . 

CHaCOOH 

(CHa)aN 

17.81 
13  20 

106 
108 

57.0 

56.1 
89  3 

iso-Butane     .  .            .... 

CH(CHa)3 

13.10 

114 

96  3 

Benzene 

Cell  6 

18  92 

120 

85  5 

86  7 

o 

n-Butane                  
Ethyl  ether     

CH3(CH2)2CH3 

(C2H5)2O 

14.66 
17.60 

122 
134 

94  0 

96  5 
100 

0 
1.2 

Triethylamirie     ... 

(C2H6)3N 
CioH« 

27.5 
40.3 

183 
193 

139 
112 

0.69 

w-Octane                  • 

CHa(CH2)aCHa 

37  8 

236 

162 

162 

o 

Decane  

CIIa(CHa)8CHi 

49.1 

289 

195 

o 

The  unit  of  pressure  in  the  constants  above  is  the  dyne  per  square  centimeter,  the  unit  of  volume  is 
cubic  centimeters  per  mole.  The  electric  moments  are  expressed  in  absolute  electrostatic  units.  Data 
for  Van  der  Waals  constants  and  volumes  are  taken  from  Landolt's  Tables;  for  the  electric  moments  from 
Debye,  "  Polare  Molekeln,"  Leipzig,  1929. 


SBC.  3]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  409 

cules,  in  electrostatic  units;  as  we  have  seen  in  Chap.  XXII,  this  has  a 
bearing  on  the  Van  der  Waals  attraction. 

The  first  thing  which  we  shall  consider  in  connection  with  the  con- 
stants of  Table  XXIV-1  is  the  set  of  b  values.  It  will  be  remembered  that 
b  represents  in  some  way  the  reduction  in  the  free  volume  available  to  a 
molecule,  on  account  of  the  other  molecules  of  the  gas.  That  is,  it  should 
bear  considerable  resemblance  to  the  actual  volume  of  the  molecules. 
According  to  statistical  mechanics,  we  have  seen  in  Chap.  XII,  Sec.  5, 
that  the  b  appropriate  to  the  limit  of  low  densities  should  be  four  times 
the  volume  of  the  molecules,  but  this  prediction  is  not  very  accurately 
fulfilled  by  experiment.  The  reason  no  doubt  is  that  that  prediction  was 
based  on  the  assumption  of  rigid  molecules,  whereas  we  have  scon  in  this 
chapter  and  the  preceding  one  that  molecular  diameters  really  depend  a 
great  deal  on  the  amount  of  compression  produced  by  various  forms  of 
interatomic  attraction.  This  was  made  particularly  plain  in  Table 
XXIII-3,  where  wo  computed  volumes  of  the  inert  gas  atoms  by  using 
radii  interpolated  between  the  ionic  radii  of  the  neighboring  positive 
and  negative  ions  and  compared  these  volumes  with  the  b  values  and  the 
volumes  of  the  liquids.  We  found,  as  a  matter  of  fact,  that  the  b  values 
were  from  three  to  five  times  the  computed  volumes  of  the  molecules,  in 
fair  agreement  with  the  prediction  of  statistical  mechanics,  but  wo  found 
that  the  volumes  of  the  liquids,  in  which  the  molecules  are  held  together 
only  by  Van  der  Waals  forces,  were  of  the  same  order  of  magnitude  as  the 
b  values,  indicating  that  the  Van  der  Waals  forces,  being  very  weak 
compared  to  ionic  forces,  cannot  compress  the  molecules  very  much.  To 
see  if  this  situation  is  general,  we  give  the  molecular  volume  of  the  liquid 
for  each  gas  in  Table  XXIV-1.  A  glance  at  the  table  will  show  the 
striking  parallelism  between  the  volume  of  the  liquid  and  the  constant  b. 
For  the  smaller,  lighter  molecules,  b  is  of  the  same  order  of  magnitude 
as  the  volume  of  the  liquid,  as  a  rule  somewhat  larger,  but  not  a  great  deal 
larger.  For  the  heavier  molecules,  there  seems  to  be  a  definite  tendency 
for  b  to  be  larger  than  the  volume  of  the  liquid,  but  even  here  it  is  not  so 
great  as  twice  as  large. 

The  actual  magnitude  of  the  constants  6  is  of  considerable  interest. 
In  the  first  place,  we  may  ask  how  much  of  the  volume  of  a  gas,  under 
ordinary  conditions  of  pressure  and  temperature,  is  occupied  by  the 
molecules.  One  mole  of  a  gas  at  atmospheric  pressure  and  0°C.  occupies 
22.4  1.,  or  22,400  cc.  The  molecules,  under  the  same  circumstances, 
occupy  the  volume  tabulated  in  Table  XXIV-1,  in  cubic  centimeters. 
For  the  common  gases,  these  volumes  are  of  the  order  of  30  or  40  cc.  This 
is  the  order  of  magnitude  of  two-tenths  of  1  per  cent  of  the  actual  volume, 
so  that  it  is  correct  to  say  that  most  of  the  volume  occupied  by  a  gas  is 
really  empty.  On  the  other  hand,  even  under  these  circumstances,  the 


410  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 

molecules  are  not  very  far  apart  in  proportion  to  their  size.  We  may 
take  an  extreme  case  of  helium,  where  at  normal  pressure  and  temperature 
the  molecules  would  occupy  some  23  cc.,  or  about  one  one-thousandth  of 
the  volume.  To  get  an  idea  of  the  spacing,  we  may  imagine  the  atoms 
spaced  out  uniformly,  each  one  in  the  center  of  a  cube  (though  of  course 
actually  they  will  be  distributed  at  random).  Then  a  cube  of  the  volume 
of  the  atom  would  have  one  one-thousandth  the  volume  of  one  of  these 
cubes  containing  an  atom,  but  the  side  of  the  small  cube  would  be  one- 
tenth  [=  (TWIT)^]  *he  s^e  °f  the  large  cube,  meaning  that  the  average 
distance  between  atoms  is  only  about 'ten  times  the  diameter  of  a  single 
atom.  This  is  for  a  small  molecule;  with  the  larger  molecules  in  the  table, 
the  molecules  are  twice  as  large  or  more  in  diameter  and  are  correspond- 
ingly closer  together  in  proportion.  We  can  imagine  that  under  these 
circumstances  real  gases  depart  quite  appreciably  from  perfect  gas  condi- 
tions. Furthermore,  it  is  natural  that  collisions  between  atoms  are 
frequent  and  that  they  are  of  great  importance  in  many  phenomena. 
This  is  all,  however,  for  one  atmosphere  pressure.  A  pressure  of  10"6  mm. 
of  mercury  can  easily  be  obtained  in  the  laboratory.  This  is  about 
10~9  atm.  and  corresponds  to  atoms  spaced  a  thousand  times  farther 
apart,  or  something  like  10,000  atomic  diameters  apart.  It  is  clear  that 
a  gas  in  this  condition  must  be  very  much  like  a  perfect  gas  and  that 
deviations  from  the  gas  law  and  interactions  between  molecules  can  be 
neglected. 

It  is  interesting  to  ask  how  much  space,  on  the  average,  is  occupied 
by  each  atom  or  molecule.  In  a  gram  molecular  weight,  as  we  have  said 
before,  there  are  about  6.03  X  1023  molecules.  Thus  if  we  divide  b  by 
this  figure,  we  shall  get  the  volume  per  molecule.  It  is  obvious  from  the 
table  that  the  molecules  with  many  atoms  have  much  larger  volumes  than 
those  with  few  atoms,  and  it  appears  very  roughly  that  the  volume  is 
something  like  12  cc.  per  mole  per  atom,  a  figure  which  we  could  get  by 
dividing  the  b  for  a  particular  molecule  by  the  number  of  atoms  in  the 
molecule.  Variations  of  more  than  100%  from  this  figure  are  seen  in  the 
table,  but  still  it  is  correct  as  to  order  of  magnitude.  This  gives 
12/(6.03  X  1023)  =  2  X  10~23  cc.  as  the  volume  assigned  to  an  atom. 
This  is  the  volume  of  a  cube  2.7  X  10~8  cm.  on  a  side;  this  figure  seems 
reasonable,  being  of  the  order  of  magnitude  of  the  dimensions  of  most  of 
the  atoms,  within  a  factor  of  two  at  most. 

From  what  we  have  said,  the  values  of  the  Van  der  Waals  constants 
b  for  the  gases  of  our  table  look  very  reasonable.  Next  we  can  consider 
their  a's.  In  Eq.  (3.6)  of  Chap.  XXII,  we  have  seen  that  the  Van  der 
Waals  interaction  energy  between  molecules  of  polarizability  a,  mean 
square  dipole  moment  /z*,  at  a  distance  r,  is 

Energy-  -JJ2L*.  (3.1) 


SBC.  3]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS 


411 


In  Eq.  (5.3),  Chap.  XII,  we  have  seen  that  the  Van  der  Waals  a  for 
molecules  of  radius  r0/2,  with  an  intermolecular  attractive  potential  of 


s 


(3.2) 

) 

where  NQ  is  Avogadro's  number.     Using  Eqs.  (3.1)  and  (3.2),  we  can 
now  write  an  explicit  formula  for  the  Van  der  Waals  a.     It  is 


(3.3) 

The  terms  of  Eq.  (3.3)  are  all  things  that  can  be  estimated.  From  Table 
XXI V-l,  we  see  that  the  permanent  dipole  moments  of  dipole  molecules 
are  of  the  order  of  magnitude  of  2  X  10~18  absolute  units,  and  we  may 
expect  the  root  mean  square  fluctuating  moments  of  other  molecules  to 
be  of  the  same  order  of  magnitude.  The  polarizability  can  be  found  from 
the  measured  dielectric  constant,  as  we  have  explained  in  Chap.  XXII, 
Sec.  3.  In  addition  to  these  quantities,  we  need  the  volume  of  the  sphere, 
^TrrJ,  which  appears  in  the  denominator  of  Eq.  (3.3).  This  will  certainly 
be  of  the  general  order  of  magnitude  of  the  molecular  volume,  and  for  the 
present  very  crude  calculation  we  may  take  it  to  be  the  same  as  6,  which 
we  have  tabulated  in  Table  XXI V-l.  (For  air  we  use  a  value  inter- 
mediate between  oxygen  and  nitrogen.)  We  have  now  computed  values 
of  ^  which,  substituted  into  the  formula  (3.3),  will  give  the  correct  value 
of  a,  and  have  tabulated  these  in  Table  XXIV-2.  We  see  that  the 
values  of  M  necessary  to  give  the  observed  a  values  are  of  the  order  of 

TABLE  XXIV-2. — CALCULATIONS  CONCERNING  VAN  DER  WAALS  ATTRACTIONS 


Gas 

Dielectric 
constant  e 

AT0a,  cubic 
centimeters 

M 

H2  
Air  

1.000264 
.000590 

0.470 
1.05 

1  .  7  X  10-"18 
3  1 

CO  

.000690 

1.23 

3.1 

CO2    

.000985 

1  75 

4  2 

CH4  

.000944 

1  .68 

3.4 

C2H4  

.00131 

2.33 

4  7 

The  dielectric  constant  c  is  given  for  gas  at  0°C.,  one  atmosphere  pressure. 
gas  occupies  2.24  X  10*  cc.  at  this  pressure  and  temperature,  we  have 


Thus,  since  a  mole  of 


-  2.24  X  10* 


using  Eq.  (3.4),  Chap.  XXII.  The  value  of  *i  is  calculated  from  Eq.  (3.3),  as  described  in  the  text,  and 
represents  the  dipole  moment  necessary  to  explain  the  observed  Van  der  Waals  a. 

magnitude  which  we  expected  to  find.  As  we  should  naturally  expect, 
they  increase  as  we  go  to  larger  and  more  complicated  molecules.  Direct 
calculations  of  /*,  or  rather  of  the  whole  Van  der  Waals  force,  have  been 


412  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 

made  for  a  few  of  the  simple  gases  like  hydrogen  and  helium,  with  fairly 
close  agreement  with  the  experimental  values.  It  seems  likely,  therefore, 
that  our  explanation  of  these  attractions  is  quite  close  to  the  truth. 

The  explanation  just  given  for  the  magnitude  of  the  Van  der  Waals 
attractions  must  be  modified  for  strongly  polar  molecules,  those  having 
large  dipole  moments.  From  Chap.  XXII,  Sec.  3,  we  remember  that  in 
such  cases  there  is  an  extra  term  in  the  polarizability,  on  account  of  the 
orientation  of  the  molecule  in  an  external  field.  We  mentioned  that  this 
would  increase  the  Van  der  Waals  attraction,  because  pairs  of  molecules 
would  tend  to  orient  each  other  into  the  position  of  maximum  attraction, 
and  suggested  that  it  might  result  in  a  Van  der  Waals  attraction  several 
times  as  great  as  for  nonpolar  molecules.  Examination  of  Table  XXIV-1 
shows  that,  in  fact,  the  strongly  polar  molecules  have  constants  a  which 
are  much  greater  than  those  of  nonpolar  molecules  near  them  in  the  list. 
Thus  water  has  an  a  value  about  four  times  those  of  its  neighbors,  and 
ammonia  about  three  times.  We  can  understand  in  detail  what  happens 
by  considering  the  most  conspicuous  case,  water.  The  crystal  structure 
of  ice  is  well  known  and  will  be  described  in  the  next  section.  It  is  a 
molecular  crystal,  in  which  each  oxygen  is  surrounded  tetrahedrally  by 
four  other  oxygens.  Between  each  pair  of  oxygens  is  a  hydrogen.  Each 
oxygen  thus  has  four  hydrogens  near  to  it.  But  two  of  these  four  hydro- 
gens are  close  to  the  oxygen,  forming  with  it  a  water  molecule,  with  its 
two  hydrogens  at  an  angle,  just  like  the  water  molecule  in  the  gas.  The 
other  two  hydrogens  are  attached  to  two  of  the  four  neighboring  oxygens, 
forming  part  of  their  water  molecules.  This  structure  puts  each  of  the 
hydrogens  of  one  molecule  near  the  oxygen  of  another,  so  that  their  oppo- 
site electrical  charges  can  attract  each  other,  helping  to  hold  the  crystal 
together.  This  arrangement  undoubtedly  persists  to  a  large  extent  in 
the  liquid  and  even  to  some  extent  in  the  gas,  though  it  undoubtedly 
decreases  as  the  temperature  is  raised,  for  at  high  temperatures  the  mole- 
cules tend  to  rotate,  spoiling  any  effect  of  orientation.  And  it  is  this 
extra  attraction,  on  account  of  the  particular  orientations  of  the  molecules, 
which  results  in  the  very  largo  value  of  a  for  water.  Similar  explanations 
hold  for  the  other  molecules  with  large  dipole  moments,  but  examination 
of  their  structure  shows  that  the  others  cannot  form  such  tightly  bound 
structures  as  water. 

There  is  another  feature  of  Table  XXIV-1  that  bears  out  the  unusually 
large  forces  between  dipole  molecules,  and  that  is  the  molecular  volumes 
of  the  liquids.  If  the  polar  molecules  have  unusually  large  attractive 
forces,  we  should  expect  that  these  forces,  which  after  all  hold  the  liquid 
together,  would  bind  it  particularly  tightly,  so  that  the  liquids  would  be 
unusually  dense.  Consistent  with  this,  we  note  that  water  and  ammonia 
conspicuously,  and  some  of  the  other  polar  liquids  to  a  lesser  extent,  have 


SBC.  3]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  413 

molecular  volumes  for  their  liquids  decidedly  smaller  than  their  b  values, 
while  the  nonpolar  molecules  have  molecular  volumes  rather  closely  equal 
to  their  b's.  To  put  these  facts  in  somewhat  more  striking  form,  if  water 
had  no  dipole  moment  we  should  expect  it  to  have  a  density  only  about 
two-thirds  what  it  does,  and  we  should  expect  the  intermolecular  forces 
to  be  so  small  that  it  would  boil  many  degrees  below  zero,  as  its  neighbors 
NO  and  02  in  the  table  do,  and  to  be  known  to  us  as  a  permanent  gas 
difficult  to  liquefy. 

The  extra  intermolecular  force  in  water  resulting  from  dipole  attrac- 
tion, which  we  have  just  discussed,  is  closely  tied  up  with  one  of  the  most 
remarkable  properties  of  water,  its  ability  to  dissolve  and  ionize  a  great 
many  ionic  compounds.  We  have  seen  in  the  last  chapter  that  it  requires 
a  large  amount  of  energy  to  pull  a  crystal  of  an  ionic  substance  apart  into 
separated  ions.  Such  a  process  surely  will  not  occur  naturally;  if  we 
examined  the  equilibrium  between  the  solid  and  the  ionic  gas,  the  heat  of 
evaporation  would  be  so  enormous  that  at  ordinary  temperatures  there 
would  be  practically  no  vapor  pressure.  If  an  ion  is  introduced  into 
water,  however,  there  is  a  strong  binding  between  the  water  molecules 
and  the  ion,  corresponding  to  a  large  negative  term  in  the  energy,  with  the 
result  that  the  heat  of  solution,  or  the  work  required  to  break  up  the 
crystal  into  ions  and  dissolve  the  ions  in  water,  is  a  small  quantity.  In 
other  words,  referring  back  to  the  type  of  argument  met  in  Chap.  XVII, 
an  ion  is  about  as  strongly  attracted  to  a  water  molecule  as  to  the  ions  of 
opposite  sign  in  the  crystal  to  which  it  normally  belongs;  this  is  the 
necessary  condition  for  solubility.  We  now  ask,  why  is  the  ion  so 
strongly  bound  to  the  water  molecules?  The  reason  is  simply  that  as  we 
have  seen  the  hydrogens  of  the  water  molecule  are  positively  charged,  the 
oxygen  negatively,  so  that  a  positive  ion  can  locate  itself  near  the  oxygens 
of  a  number  of  neighboring  water  molecules,  a  negative  ion  near  a  number 
of  hydrogens,  which  attract  it  electrostatically  approximately  as  much  as 
two  ions  would  attract  each  other,  or  as  the  negative  oxygen  of  one  water 
molecule  would  attract  the  positive  hydrogens  of  its  neighboring  water 
molecules.  This  effect  is  particularly  strong  in  water,  for  the  same  reason 
that  the  Van  der  Waals  binding  is  strong  in  water;  similar  effects  are 
found  in  a  lesser  extent  in  liquid  ammonia,  which  is  also  a  powerful  ioniz- 
ing solvent,  with  many  of  the  same  properties  as  water. 

We  have  mentioned  several  times  that  the  characteristic  of  the 
molecular  compounds  is  that  the  Van  der  Waals  forces  between  molecules 
are  small  compared  to  the  valence  forces  holding  the  atoms  together  to 
form  a  molecule.  Thus  the  substances  vaporize  at  a  low  temperature, 
whereas  their  molecules  do  not  dissociate  chemically  to  any  extent  except 
at  very  high  temperatures.  For  instance,  the  dissociation  H2  ^  2H  is 
a  typical  example  of  chemical  equilibrium,  to  be  handled  by  the  methods 


414 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 


of  Chap.  X,  and  the  forces  holding  the  atoms  together  are  the  sort  taken 
up  in  Chap.  IX.  We  saw  in  Table  IX-1  that  the  heat  of  dissociation  of  a 
hydrogen  molecule  was  103  kg.-cal.  per  gram  mole,  such  a  large  value  that 
the  dissociation  is  almost  negligible  at  any  ordinary  temperature.  On  the 
other  hand,  the  latent  heat  of  vaporization,  the  heat  required  to  pull  the 
molecules  of  the  liquid  or  solid  apart  to  form  the  gas,  is  only  0.256  kg.-cal. 
per  gram  mole  in  this  case,  so  that  a  temperature  far  below  0°C.  will 
vaporize  hydrogen.  To  illustrate  how  general  this  situation  is,  Table 
XXIV-3  gives  the  latent  heat  of  vaporization  and  the  heat  of  dissociation 

TABLE  XXIV-3. — LATENT  HEAT  OF   VAPOKIZATION   AND   HEAT  OF  DISSOCIATION 


Substance 

Latent  heat, 
kg.-cals.  per 
gram  mole 

Heat  of  dis- 
sociation, 
kg.-cals.  per 
gram  mole 

H2  

0.220 

103 

O2  

2.08 

117 

N2    

1.69 

170 

CO                     

I  90 

223 

CO2              

6  44 

NH8  

7.14 

90 

HC1  

4.85 

102 

H2O  

11.26 

118 

Latent  heats  are  from  Landolt's  Tables,  and  in  each  case  are  for  as  low  a  temperature  as  possible, 
since  the  latent  heat  of  vaporization  decreases  with  increasing  temperature,  going  to  zero  at  the  critical 
point.  Heats  of  dissociation  are  from  Sponer,  "  Molekulspektren  und  ihre  Anwendungen  auf  chemische 
Probleme,"  Berlin,  1935,  some  of  them  having  been  quoted  m  Table  IX-1. 

of  a  few  familiar  gases.  The  heat  of  dissociation  in  each  case  is  the  energy 
required  to  remove  the  most  loosely  bound  atom  from  the  molecule.  We 
see  that  in  each  case  the  latent  heat  is  only  a  few  per  cent  of  the  heat  of 
dissociation;  water  is  a  distinct  exception  on  account  of  its  high  latent 
heat,  ten  per  cent  of  the  heat  of  dissociation,  which  of  course  is  tied  up 
with  the  large  Van  der  Waals  attraction  arising  from  the  dipole  moments 
of  the  molecules. 

4.  Molecular  Crystals. — The  molecules  which  we  have  been  discussing 
in  this  chapter  are  tightly  bound  structures,  held  together  by  strong 
homopolar  forces.  Mathematically,  these  forces  can  be  described 
approximately  by  Morse  curves,  as  discussed  in  Sec.  1,  Chap.  IX.  On 
the  other  hand,  the  forces  holding  one  molecule  to  another  are  simply  the 
Van  der  Waals  forces,  which  we  have  spoken  about  in  Chap.  XXII,  and 
which  are  very  much  weaker  than  homopolar  forces,  as  we  saw  from 
Table  XXIV-3.  It  thus  comes  about  that  the  crystals  of  these  materials 
consist  of  compact  molecules,  spaced  rather  widely  apart.  Since  the 
forces  between  molecules  are  so  weak,  the  crystals  melt  at  low  tempera- 


SBC.  4]       HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  415 

tures,  are  very  compressible,  and  are  easily  deformed  or  broken,  in  con- 
trast to  the  ionic  crystals  with  their  considerable  mechanical  strength,  low 
compressibility,  and  high  melting  points.  In  this  section  we  shall  discuss 
the  structure  of  a  few  of  the  molecular  crystals. 

We  start  with  the  inert  gases.  The  atoms  of  these  substances  are 
spherical,  and  we  should  naturally  expect  that  their  crystals  would  be 
formed  simply  by  piling  the  spheres  on  top  of  each  other  in  the  closest 
manner  possible.  This  is,  in  fact,  the  case.  There  are  two  alternative 
lattices,  corresponding  to  the  closest  packing  of  spheres.  Of  these,  the 
inert  gases  choose  the  type  called  the  face-centered  cubic  structure.  This 
structure  is  shown  in  Fig.  XXI V-l.  In  the  first  place,  we  can  regard  the 
structure  as  arising  from  a  simple  cubic  lattice,  as  shown  in  (a).  There 
is  an  atom  at  each  corner  of  the  cube  and  in  the  center  of  each  face. 
Comparison  with  Fig.  XXIII-1,  showing  the  sodium  chloride  structure, 


(«)  (b)  (c) 

FIG.  XXI V-l. — The  face-centered  cubic  structure,  (a)  atoms  at  the  corners  of  a  cube 
and  the  centers  of  the  faces.  (6)  the  same  atoms  connected  up  in  planes  perpendicular  to 
the  cube  diagonal,  (c)  view  of  successive  planes  looking  along  the  cube  diagonal,  illustrat- 
ing the  close-packed  nature  of  the  structure. 

will  show  that  in  the  latter  type  of  structure  the  positive  ions  by  them- 
selves, or  the  negative  ions  by  themselves,  form  a  face-centered  cubic 
structure.  Another  way  of  looking  at  the  face-centered  cubic  structure 
can  be  understood  from  Fig.  XXI V-l  (6).  In  this,  we  have  drawn  the 
atoms  just  as  in  (a)  but  have  connected  them  up  differently,  so  as  to  form 
two  parallel  triangles  of  six  atoms  each,  oppositely  oriented,  with  two 
extra  atoms.  If  we  had  considered,  not  simply  the  cube  of  (a)  but  the 
whole  crystal,  each  of  these  triangles  would  have  been  part  of  a  whole 
plane  of  atoms.  These  same  six  atoms  are  shown  by  the  heavy  circles  in 
(c),  where  we  now  look  down  along  the  normal  to  the  planes  of  the  tri- 
angles; that  is,  along  the  body  diagonal  of  the  cube,  shown  dotted  in 
(a).  In  (c),  we  have  drawn  the  circles  representing  the  atoms  large 
enough  so  that  they  touch,  as  if  they  were  closely  packed  spheres.  The 
next  layer  of  spheres  is  shown  in  (c)  by  dotted  lines,  and  we  see  that  it  is  a 
layer  similar  to  the  first  but  shifted  along,  atoms  of  the  second  layer  fitting 
into  every  other  one  of  the  depressions  between  atoms  in  the  first  layer. 
The  third  layer,  of  which  only  one  atom  is  shown  in  (6),  fits  on  top  of  (c) 


416 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 


in  a  similar  manner,  the  one  atom  shown  in  (6)  going  in  the  common  center 
of  the  triangles  of  (c).  After  these  three  layers,  the  structure  repeats,  the 
fourth  layer  being  just  like  the  one  drawn  in  heavy  lines  in  (c).  From 
this  description  in  terms  of  Fig.  XXI V-l,  (c),  it  is  clear  that  the  face- 
centered  structure  is  a  possible  arrangement  for  close  packed  spheres. 
Each  atom  has  twelve  equally  spaced  neighbors,  six  in  the  same  plane 
in  (c),  three  each  in  the  planes  directly  above  and  directly  below. 

As  we  have  stated,  the  inert  gases  crystallize  in  the  face-centered 
cubic  structure.  The  distances  between  nearest  neighbors  are  given  in 
Table  XXIV-4.  In  this  table  we  give  also  the  volume  of  the  crystal  per 

TABLE  XXIV-4. — CRYSTALS  OF  INERT  GASES 


Interatomic 
distance,  A 

Volume,  cc. 
per  mole 

Volume  of 
liquid,  cc. 
per  mole 

Ne  

3  20 

14  0 

16.7    ' 

A  

3.84 

24  3 

28  1 

Kr  ...      . 

3.94 

26.4 

38  9 

Xe  

4  37 

35.8 

47  5 

mole,  computed  in  a  simple  way  from  the  crystal  structure,  and  finally  we 
give  tho  volume  of  the  liquid  per  mole,  from  Table  XXI V-l,  for  compari- 
son. We  see  that  the  volumes  of  the  crystals  are  somewhat  but  not  a 
great  deal  less  than  the  volumes  of  the  liquids,  as  we  should  probably 
expect,  since  in  the  liquids  the  same  atoms  are  packed  in  a  less  orderly 
arrangement.  The  interatomic  distances  in  these  crystals  are  much 
greater  than  interatomic  distances  in  any  cases  where  the  atoms  are  held 
by  either  homopolar  or  ionic  bonds.  This  has  already  been  commented 
on  in  Sec.  3  and  has  been  explained  by  saying  that  the  large  attractive 
forces  of  the  homopolar  or  ionic  bonds  pull  atoms  together,  essentially 
compressing  them,  so  that  they  get  much  closer  together  than  when  held 
only  by  the  weak  Van  der  Waals  forces. 

The  inert  gases  are  the  only  strictly  spherical  molecules,  but  a  number 
of  the  other  gases  have  molecules  nearly  enough  spherical  to  pack  together 
in  similar  ways.  The  hydrogen  molecule,  except  at  the  very  lowest 
temperatures,  is  in  continual  rotation,  so  that  while  it  is  not  spherical  at 
any  instant,  still  it  fills  up  a  spherical  volume  on  the  average,  the  volume 
swept  out  by  its  two  atoms  when  they  are  pivoted  at  the  midpoint  of  the 
line  joining  them  and  are  free  to  rotate  in  any  plane  about  this  point. 
Hydrogen  molecules,  then,  pack  as  rigid  spheres,  but  they  adopt  the  other 
method  of  close  packing,  the  so-called  hexagonal  close-packed  structure. 
This  is  shown  in  Fig.  XXIV-2.  It  starts  with  the  same  layer  of  atoms 
shown  by  the  heavy  lines  in  Fig.  XXI V-l  (c),  then  has  the  dotted  layer  of 


SBC.  4]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS 


417 


Fia.     XXIV-2.- -Hexagonal 
riose-paekcd  structure. 


(c),  but  after  these  two  layers  it  has  another  layer  like  the  first  one,  and 
so  on,  having  just  two  alternating  types  of  layer  instead  of  three  as  in  the 
face-centered  cubic  structure.  Part  of  the  structure  is  shown  in  perspec- 
tive in  Fig.  XXIV-2.  This  indicates  plainly  the  hexagonal  unit  from 
which  the  structure  takes  its  name.  As  we  have  stated,  hydrogen  crystal- 
lizes in  this  form,  a  hydrogen  molecule  being  at  each  lattice  point.  The 
distance  between  molecules,  on  centers,  is  3.75  A,  the  volume  per  molo 
is  21.7  cc.,  to  be  compared  with  26.3  cc.  in  the  liquid. 

The  molecules  N2  and  CO,  though  they  are  rather  far  from  spherical, 
still  crystallize  approximately  though  not  exactly 
like  close-packed  spheres.  Their  crystal  is  a 
slightly  distorted  face-centered  cubic  structure, 
one  dumbbell-shaped  molecule  being  located  at 
each  lattice  point.  The  molecules  of  those  sub- 
stances do  not  rotate  enough  to  simulate  a 
spherical  shape  at  ordinary  temperature,  but  in- 
stead they  oscillate  about  definite  directions  in 
space.  These  directions  are  determined  for  the 
various  molecules  in  the  crystal  in  a  rather  com- 
plicated though  regular  way,  there  being  several 
different  orientations  for  different  molecules. 
The  molecules  are  spaced  about  3.96  A  apart  on  centers,  resulting  in  a 
volume  of  27.0  cc.  per  mole,  both  for  N2  and  CO,  compared  to  32.8  cc.  per 
mole  in  the  liquid  for  N«,  and  33  cc.  for  CO. 

Many  molecules  that  are  not  spherical  still  rotate  enough  at  high 
temperatures  so  that  they  simulate  spheres,  like  the  hydrogen  molecule. 
In  some  cases,  such  molecules  oscillate  about  definite  directions  at  low 
temperatures,  as  with  N2  and  CO,  but  simulate  spheres  at  higher  tempera- 
tures when  they  are  rotating  with  more  energy.  In  such  cases,  the 
substance  has  two  crystal  forms,  and  there  is  a  transition  from  one  to 
the  other  at  a  definite  temperature.  The  low  temperature  phase  is  likely 
to  be  complicated  in  structure,  with  the  molecules  pointing  in  definite 
directions,  while  the  high-temperature  phase  is  one  of  the  close-packed 
structures.  Hydrogen  chloride  HC1  is  a  case  in  point.  Below  98°  abs., 
the  molecules  are  hindered  from  rotating  and  the  structure  is  a  compli- 
cated one  which  has  not  been  completely  worked  out.  At  this  tempera- 
ture there  is  a  transition,  and  above  98°  the  molecules  rotate  freely  and 
the  substances  crystallize  in  a  face-centered  cubic  structure.  In  many 
cases,  where  we  might  expect  such  a  transition,  it  does  not  occur  in  the 
available  temperature  range.  Thus  we  should  expect  that  hydrogen, 
which  shows  free  rotation  under  ordinary  conditions,  might  conceivably 
show  a  transition  to  another  structure  with  hindered  rotation  at  low 
enough  temperatures,  while  CO  and  N2,  with  hindered  rotation  at  ordi- 


418  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIV 

nary  temperatures,  might  have  a  transition  to  a  state  of  free  rotation  at 
high  enough  temperatures.  But  the  necessary  temperatures  might  be 
above  the  melting  point,  in  which  case  the  transition  could  not  really  be 
observed. 

The  diatomic  molecules  which  show  hindered  rotation  in  the  solid 
generally  have  quite  complicated  molecular  crystals.  This  is  true,  for 
instance,  of  the  halogens.  CU  forms  a  crystal  composed  of  molecules, 
each  of  interatomic  distance  1.82  A  (compared  to  1.98  A  in  the  gas), 
arranged  in  a  complicated  way  which  we  shall  not  describe.  Iodine  I2 
forms  a  layer  lattice.  In  Fig.  XXIV-3  we  show  one  of  the  layers,  showing 


\) 


FIG.  XXIV-3.  —  Layer  of  molecules  in  la  structure. 


the  diatomic  molecules  arranged  in  two  sorts  of  rows.  The  spacing 
between  atoms  in  a  molecule  is  2.70  A  compared  to  2.66  A  in  the  gas. 
The  spacing  between  different  molecules  in  the  same  row  is  4.79  A; 
between  rows,  4.89  A  on  centers;  between  layers,  3.62  A.  This  structure 
is  typical  of  the  sort  that  one  finds  in  other  cases. 

Among  the  hydrides,  water  has  received  more  attention  than  the 
others  and  is  fairly  well  understood.  The  hydrides  are  hard  to  analyze 
by  x-ray  diffraction  methods,  because  the  hydrogens  are  not  shown  by  the 
x-ray  photographs;  we  must  use  other  evidence  to  find  where  they  are 
located.  As  we  have  mentioned  earlier,  we  find  that  each  oxygen  is 
tetrahedrally  surrounded  by  four  other  oxygens.  Between  each  pair  of 
oxygens  is  a  hydrogen,  two  of  the  four  hydrogens  near  an  oxygen  being 
joined  to  it  to  form  a  water  molecule,  the  other  two  being  attached  to  two 
of  the  four  neighboring  oxygens  to  form  part  of  their  molecules.  This 
structure,  as  we  have  mentioned  in  Sec.  3,  puts  each  hydrogen  of  one 
molecule  near  the  oxygen  of  another,  so  that  their  opposite  electrical 


SBC.  4]        HOMOPOLAR  BOND  AND  MOLECULAR  COMPOUNDS  419 

charges  can  attract  each  other,  helping  to  hold  the  crystal  together  almost 
as  if  it  were  an  ionic  crystal.  In  Fig.  XXIV-4  we  show  a  layer  of  the 
crystal,  indicating  the  oxygens  by  spheres,  with  vectors  drawn  out  to  the 
hydrogens.  Three  neighbors  of  the  upper  molecules  in  the  layer,  which 
we  have  drawn,  are  shown;  the  fourth  lies  directly  above  and  is  not 
shown.  The  molecules  are  spaced  2.76  A  on  centers.  It  is  interesting  to 
notice  that  this  is  slightly  less  than  twice  the  ionic  radius  1.45  A  of  oxygen 
given  in  Table  XXIII-2,  showing  that  ice  is  not  entirely  different  from  an 
ionic  crystal.  The  molecules  have  such  a  spacing  that  the  volume  of  the 
solid  is  20.0  cc.  per  mole,  compared  to  18  cc.  per  mole  for  the  liquid,  agree- 


FIG.  XXIV-4  — Layer  of  ice  structure. 

ing  with  the  well-known  fact  that  ice  is  less  dense  than  water.  This  is 
because  the  molecules  in  ice  are  unusually  loosely  packed.  In  water,  the 
individual  molecules  are  actually  farther  apart,  about  2.90  A  on  centers 
compared  to  2.76  A  in  ice,  but  they  are  packed  so  much  more  efficiently 
that  there  are  more  molecules  in  less  space  and  a  greater  density.  In 
addition  to  showing  how  each  oxygen  is  surrounded  by  four  others,  Fig. 
XXIV-4  shows  the  hexagonal  structure  which  is  so  characteristic  of  ice 
crystals  and  which  is  well  known  from  the  form  of  snow  flakes. 


CHAPTER  XXV 
ORGANIC  MOLECULES  AND  THEIR  CRYSTALS 

In  Chap.  XXIV,  we  have  been  talking  about  substances  held  together 
by  homopolar  bonds.  We  have  had  a  rather  small  list  of  compounds  to 
work  on;  but  we  have  hardly  touched  the  most  fertile  field  for  discussing 
the  homopolar  bond.  Organic  chemistry  of  course  presents  the  best 
organized  and  most  extensive  field  for  the  theory  of  homopolar  valence. 
The  carbon  atom  can  form  four  single  bonds,  which  tend  to  be  oriented 
toward  the  four  corners  of  a  tetrahedron,  and  this  furnishes  the  funda- 
mental fact  on  which  the  chemistry  of  the  aliphatic  or  chain  compounds  is 
based.  The  great  difference  between  organic  and  inorganic  chemistry  is 
the  way  in  which  more  and  more  carbons  can  be  bonded  together  to  form 
great  chains,  resulting  in  molecules  of  great  complexity.  These  carbon 
chains  form  the  framework  of  the  organic  compounds,  the  other  atoms 
merely  being  attached  to  the  carbons  in  most  cases.  In  the  first  section, 
we  discuss  the  ways  in  which  carbons  can  be  joined  together. 

1.  Carbon  Bonding  in  Aliphatic  Molecules. — In  the  first  place,  two 
carbon  atoms  can  join  together,  as  for  instance  in  ethane,  by  a  single 

H  H 

bond.     Thus  ethane  has  the  structure  H:C:C:H,  as  indicated  in  Sec.  1, 

HH 

Chap.  XXIV.  The  carbon-carbon  distance  in  this  case  is  about  1 .54  A,  a 
value  approximately  correct  for  the  carbon-carbon  distance  in  all  aliphatic 
molecules  with  single  bonds.  The  hydrogens  are  arranged  around  the 
carbons  so  that  the  three  hydrogens  and  one  carbon  surrounding  either 
carbon  have  approximately  tetrahedral  symmetry.  The  carbon-hydro- 
gen distance  is  presumably  about  1.1  A,  as  in  methane.  Unfortunately 
this  carbon-hydrogen  distance  is  almost  impossible  to  determine  accu- 
rately, since  the  hydrogen  atom  represents  too  small  a  concentration  of 
electrons  to  be  shown  in  x-ray  or  electron  diffraction  pictures.  Each  of 
the  CH3  groups  is  able  to  rotate  almost  freely  about  the  axis  joining  the 
two  carbons,  as  shown  in  Fig.  XXV-1.  Thus  there  is  no  fixed  relation 
between  the  positions  of  the  hydrogens  on  one  carbon  and  those  on 
another. 

If  more  than  two  carbon  atoms  join  together  to  form  a  chain,  they 
necessarily  form  a  zigzag  structure,  on  account  of  the  tetrahedral  angle 
between  bonds.  Thus  in  Fig.  XXV-2  we  show  propane 

420 


SBC.  1] 


ORGANIC  MOLECULES  AND  THEIR  CRYSTALS 


421 


and  butane  CHaCI^CEUCHs.  The  carbon-carbon  distances  as  before 
are  about  1.54  A  and  the  carbon-hydrogen  distances  about  1.1  A.  On 
account  of  this  zigzag  nature,  the  chains  with  an  even  number  of  carbons 
act  differently  from  those  with  an  odd  number,  and  there  is  an  alternation 


FIG.  XXV-l.—  The  ethane  molecule,  CH3-OH3. 

in  physical  properties  as  we  go  up  the  scries  of  chain  compounds,  the 
even-numbered  compounds  falling  on  one  curve,  the  odd-numbered  on 
another.  Chains  of  practically  indefinite  length  can  be  built  up,  and  it  is 
interesting  to  see  how  the  physical  properties  of  the  substances  change 


Propane 


Fia.  XXV-2.- 


Bufane 
-The  molecules  of  propane,  CH»CH2CH3,  and  butane,  CH8CH2CH2CH8. 


as  the  chains  get  longer  and  longer.  For  example,  in  Fig.  XXV-3  we  show 
the  melting  point  and  boiling  point  of  the  chain  compounds  as  a  function 
of  the  number  of  carbon  atoms  in  the  chain.  The  alternation  of  which  we 
have  spoken  is  obvious  in  the  melting  points,  though  not  in  the  boiling 


422 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXV 


points.  This  can  be  explained  as  follows.  In  the  solids,  as  we  shall 
mention  later,  the  zigzag  molecules  are  arranged  with  their  carbons  all 
in  a  plane,  so  that  the  line  joining  carbons  to  their  neighbors  makes  a 
sort  of  saw-tooth  shape.  Then  the  molecules  with  an  odd  number  of 
carbons,  in  which  the  lines  joining  the  two  end  carbons  point  in  different 
directions  at  the  two  ends  of  the  chain,  are  definitely  different  from  those 
with  an  even  number,  in  which  the  end  lines  point  in  the  same  direction 
at  the  two  ends  of  the  chain.  Thus  we  can  expect  the  solids  to  show  an 
alternation  in  properties.  But  in  the  liquid,  or  the  gas,  the  possibility  of 
*ree  rotation  about  a  carbon-carbon  bond  results  in  a  great  flexibility  of 


300  - 


-200 


345 


6    7    8    9    10  11   12  13  14  15  16  17 
Number  of  Carbons 


18 


FIG.  XXV-3. — Melting  and  boiling  points  of  chain  compounds,  as  a  function  of  the  number 

of  carbon  atoms  in  the  chain. 

the  chain.  It  can  turn  and  twist,  forming  anything  but  an  approximately 
straight  chain,  and  the  result  is  that  the  two  ends  will  be  oriented  quite 
independently  of  each  other.  Thus  there  will  be  no  average  difference,  in 
the  liquid  or  gas,  between  the  chains  with  even  and  odd  numbers  of  car- 
bons. Now  the  melting  point  measures  the  equilibrium  between  liquid 
and  solid,  so  that  the  alternation  in  the  properties  of  the  solid  will  show  in 
the  curve,  but  the  boiling  point  depends  only  on  liquid  and  gas  and  will 
not  show  the  alternation. 

In  addition  to  this  feature,  Fig.  XXV-3  is  interesting  in  that  it  shows 
that  the  melting  and  boiling  points  of  the  chain  hydrocarbons  increase 
rapidly  as  the  chain  gets  longer.  Not  only  that,  but  the  viscosity  of  the 
liquids  goes  up  as  carbons  are  added  to  the  chain.  These  effects  are 
qualitatively  reasonable.  Anything  tending  to  hold  molecules  together 
tends  to  increase  the  melting  and  boiling  points.  Now  a  chain  hydro- 
carbon,  as  far  as  the  carbons  are  concerned,  is  much  like  a  string  of 


SBC.  1]  ORGANIC  MOLECULES  AND  THEIR  CRYSTALS  423 

methane  molecules  fastened  together  like  a  string  of  beads.  In  a  very 
rough  way,  the  valence  forces  hold  the  carbons  and  their  hydrogens 
together  tightly,  whereas  the  Van  der  Waals  forces,  the  only  ones  opera- 
tive in  methane,  hold  that  substance  together  only  very  loosely.  It  is 
only  reasonable,  then,  that  the  boiling  points  of  these  long  chain  hydro- 
carbons should  be  higher  than  for  the  short  ones.  The  increased  viscosity 
of  the  liquids  is  also  reasonable.  After  all,  a  liquid  made  of  long  flexible 
chains  will  certainly  get  tangled  up,  just  as  a  mass  of  threads  will  get 
tangled  and  knotted.  Quite  literally,  if  the  chains  are  long  enough,  the 
molecules  will  tie  each  other  up  in  knots  and  prevent  flow  of  one  molecule 
past  another. 

H      H      H      H      H      M 

H— C  —  C—  C  —  C  —  C  —  C  —  H 
I        I        I        I        I        I 
H        H       H       H       H       H 

H  H 

H-C-H  H-C-H 

H        |       H       H       H  H       H        I        H       H 

I        I        I        I        I  I        I        I        I        I 

H-C  —  C- C  —  C—  C-H  H-C  —  C—  C  —  C—  C— H 
I        I       I        I        I  I        I        I        I        I 

HHHHH  HHHHH 


H  H 


H       H  HH      H 

I  I  I  I  I  I  I 

H—  C—  C—  C  —  C—  H  H—  C  —  C  —  C—  C—  H 


K  J 

Flo.  XXV-4.  —  Isomers  of  hexane,  C«HU. 

We  have  spoken  of  the  simple  chain  hydrocarbons,  in  which  there  is  a 
single  chain  of  carbons,  with  their  attached  hydrogens.  These  arise  when 
each  carbon  of  the  chain,  except  the  end  ones,  is  joinod  to  only  two  other 
carbons  and  two  hydrogens.  But  there  is  nothing  to  prevent  a  carbon 
being  bonded  to  three  or  to  four  other  carbons.  In  other  words,  branch 
chains  can  be  formed.  In  this  way,  new  branching  compounds  are 
formed,  which  in  general  will  have  the  same  chemical  composition  as  some 
one  of  the  simple  chains,  but  of  course  will  be  rather  different  in  physical 
properties.  Two  such  compounds,  having  the  same  number  of  atoms  of 
each  element  but  with  different  arrangements,  are  called  isomers.  As 
an  illustration,  Fig.  XXV-4  shows  the  formulas  for  the  isomers  of  hexane. 
Of  course,  on  account  of  the  possibility  of  rotation  about  C-C  bonds,  these 


424  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXV 

molecules  can  assume  many  complicated  shapes.  When  we  begin  to  get 
complicated  side  chains,  however,  the  possibility  of  rotation  is  somewhat 
diminished,  since  different  parts  of  the  molecule  can  get  into  each  other's 
way  with  some  orientations.  This  effect  is  called  steric  hindrance,  and  it 
operates  to  stiffen  the  molecule  to  gome  extent.  It  can  hardly  stiffen  a 
long  chain  to  any  great  extent,  however,  and  the  various  branches  of  a 
hydrocarbon  with  long  branches  are  presumably  very  flexible. 

If  the  branching  process  extends  very  far,  there  is  no  reason  why  the 
extremity  of  one  branch  cannot  join  onto  another,  forming  a  closed  loop. 
The  simplest  compound  in  which  this  occurs  is  cyclohexane,  shown  in 
Fig.  XXV-5.  A  geometrical  investigation,  made  most  easily  with  a 
model,  will  show  that  with  less  than  the  six  carbon  atoms  of  cyclohexane 


FIG.  XXV-5. — The  cyclohexane  mole-  FIG.  XXV-6. — The  diamond 

cule  (CH2)e»  structure. 

it  involves  considerable  distortion  of  the  tetrahedral  bond  angles  to  form 
a  closed  ring,  but  that  six  carbons  can  join  up  with  no  distortion  of  the 
tetrahedral  angles.  The  atoms  are  rigidly  held  in  position  by  their  bond- 
ing, in  this  case,  so  that  cyclohexane  is  a  much  more  rigid  molecule  than 
the  flexible  chain  hydrocarbons. 

The  branching  process  and  the  formation  of  closed  loops  can  continue 
much  further  than  it  is  carried  in  cyclohexane.  The  ultimate  in  this 
line  is  the  diamond.  This  is  the  structure  obtained  when  every  carbon 
is  joined  tetrahedrally  to  four  other  carbons.  It  makes  a  continuous  lat- 
tice filling  space,  as  shown  in  Fig.  XXV-6,  which  is  essentially  the  same 
as  the  zincblende  structure  of  Fig.  XXIII-3  except  that  it  is  formed  of 
only  one  type  of  atom.  The  type  of  rigidity  present  in  cyclohexane  is 
found  here  in  its  most  extreme  form.  The  structure  is  braced  in  every 
direction,  and  the  result  is  that  diamond  is  the  hardest  and  most  rigid 
material  known.  One  can  trace  out  hexagons  like  cyclohexane  in  the 


SEC.  2]  ORGANIC  MOLECULES  AND  THEIR  CRYSTALS  425 

diamond  crystal;  one  only  has  to  replace  the  hydrogens  in  cyclohexane  by 
carbons  and  continue  the  lattice  indefinitely  to  get  diamond.  Not  only 
the  arrangement  but  the  lattice  spacing  of  diamond  is  the  same  as  in  the 
aliphatic  chain  compounds:  the  carbon-carbon  distance  in  diamond  is 
1.54  A,  just  as  in  the  chains. 

The  possibility  of  carbon  chains  is  what  leads  to  the  richness  of 
organic  chemistry.  A  diamond  is  really  a  molecule  of  visible  dimensions, 
held  together  by  just  the  same  forcoa  actirig  in  small  molecules.  There  is 
no  reason  why  there  cannot  be  all  intermediate  stages  between  the  small 
molecules  made  of  a  few  atoms  which  we  usually  think  about  and  mole- 
cules of  enormous  size.  Obviously  carbon  atoms  can  link  themselves 
together  in  innumerable  ways,  if  we  only  have  enough  of  them.  The 
organic  chemists  have  discovered  a  very  great  number  of  kinds  of  mole- 
cules, but  there  seems  no  reason  why  they  cannot  go  on  forever  without 
exhausting  the  possibilities.  For  by  the  time  a  structure  is  built  up  of 
carbon  atoms,  many  atoms  in  length,  one  end  can  no  longer  be  expected 
to  know  what  the  other  end  is  doing.  There  is  no  reason  why  one  mole- 
cule cannot  add  chains  in  one  way,  another  in  another,  and  form  a  con- 
tinually increasing  variety  of  new  molecules.  One  gets  to  the  point  very 
easily  where  it  hardly  pays  to  speak  about  molecules  of  a  single  type  at 
all,  but  where  one  may  have  chains  of  indefinite  length  and  things  of  that 
sort.  Such  situations  are  presumably  met  in  problems  of  living  matter, 
where  many  molecules  are  of  almost  microscopic  size.  We  shall  meet  one 
other  field  in  which  we  have  similar  chain  formation,  and  therefore  a  great 
variety  of  compounds:  the  silicates,  which  form  a  chain  of  alternating 
silicons  and  oxygens.  As  the  carbon  chain  leads  to  the  great  variety  of 
materials  in  organic  chemistry  and  living  matter,  so  we  shall  see  that  the 
silicon-oxygen  chain  leads  to  the  great  variety  of  materials  in  the  field  of 
mineralogy. 

2.  Organic  Radicals. — The  carbon  chains  form  the  skeleton,  so  to 
speak,  of  aliphatic  organic  compounds.  But  in  place  of  the  hydrogens 
which  are  attached  to  them  in  the  simple  hydrocarbons,  there  are  many 
organic  radicals  which  can  be  bonded  to  the  carbons  in  various  positions 
of  the  molecule,  thus  greatly  increasing  the  complexity  of  the  possible 
compounds.  We  shall  mention  only  a  few  of  the  simple  radicals  in  this 
section. 

In  the  first  place,  a  single  electronegative  atom  can  act  like  an  organic 
radical,  being  substituted  for  a  hydrogen.  The  monovalent  halogens, 
F,  Cl,  Br,  I,  replace  hydrogen  freely.  Like  hydrogen,  they  can  form  a 
single  homopolar  bond  with  carbon.  For  instance,  methyl  chloride  has 

H 

the  structure  H :  C :  Cl : .    Like  methane,  it  is  tetrahedral.     The  carbon- 
H  " 


426 


INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXV 


hydrogen  distance  is  presumably  about  1.1  A,  as  in  methane,  and  the 
carbon-chlorine  distance  is  1.77  A.  Similarly  two,  three,  or  all  four,  of 
the  hydrogens  can  be  replaced  by  a  halogen  atom,  not  necessarily  all 
the  same  halogens.  In  these  compounds,  the  carbon-halogen  distances 
are  always  approximately  the  following: 

Carbon-fluorine 1.36  A 

Carbon-chlorine. .         .      .  1.77  A 

Carbon-bromine 1.93  A 

Carbon-iodine 2.28  A.     (2.1) 

These  distances  are  not  far  from  the  ionic  radii  of  Table  XXIII-2,  which 
were  1.30,  1.80,  1.95,  2.20  A  respectively  for  F",  C1-,  Br~,  I".  Since 
the  carbon  certainly  has  nonvanishing  dimensions,  this  means  that  in 
these  bonds  there  is  considerable  shrinkage  from  the  atomic  distances  in 
ionic  crystals,  but  not  so  much  shrinkage  as  in  some  other  cases,  so  that 
we  should  not  be  surprised  to  find  that  the  halogen  atoms  have  quite  a 
little  of  the  properties  of  negative  ions.  As  a  matter  of  fact,  these  com- 
pounds have  rather  strong  dipole  moments:  in  CH3C1,  for  instance,  we  see 
from  Table  XXI V-l  that  the  moment  is  1.97  X  10~18  absolute  units, 
corresponding  to  about  0.23  of  an  electron  at  the  distance  1.77  A.  We 
may  conclude,  then,  that  the  halogen  atoms  pull  the  electrons  that  they 
share  with  the  methyl  or  other  organic  group  rather  strongly  toward 
them,  so  that  they  have  quite  a  little  the  structure  of  negative  ions.  In 
the  matter  of  physical  properties,  we  can  see  from  Table  XXIV-1  that 
replacing  hydrogen  by  halogens  increases  both  Van  der  Waals  a  and  6,  as 
we  should  expect  from  the  fact  that  the  halogen  atoms  are  much  bigger 
than  hydrogen.  Thus  for  the  series  CH4,  CH3C1,  CH2C12,  CHC18,  CC14, 
we  have  the  properties  shown  in  Table  XXV-1.  The  Vs  increase  fairly 

TABLE  XXV-1. — PROPERTIES  OF  SUBSTITUTED  METHANES 


a 

b 

Boiling 
point,  °C 

CH4  ... 
CH8C1                .                    

2.28  X  1012 
7.56 

42.6 
64.5 

-161.4 
-  23  7 

CH2C12  

CHCla  

15.38 

102 

61  2 

ecu  

20.65 

138 

76  0 

regularly  as  more  chlorines  are  added,  and  the  amount  of  increase  per 
chlorine  is  not  far  from  28  cc.  per  mole,  which  is  half  the  b  value  for 
C12  (56.2,  from  Table  XXIV-1).  The  increase  in  the  a's  and  &'s  leads  to 
an  increase  in  boiling  points,  as  is  shown,  and  as  is  natural  with  larger 
and  heavier  molecules. 


SBC.  2]  ORGANIC  MOLECULES  AND  THEIR  CRYSTALS  427 

Divalent  and  trivalent  as  well  as  monovalent  electronegative  atoms 
can  also  attach  themselves  to  organic  carbon-hydrogen  chains.  Thus,  in 
particular,  oxygen  plays  a  very  important  part  in  organic  compounds.  If 
an  oxygen  atom  attaches  itself  by  a  single  bond  to  a  carbon,  it  has  another 
bond  free,  with  which  to  attach  itself  to  something  else.  This  second 
bond  may  go  to  a  hydrogen,  in  which  case  we  have  the  organic  OH  group, 

H 
forming  an  alcohol.     Thus  methyl  alcohol  is  H:C:0:.     The  OH  group, 

HH 

like  the  halogens,  though  it  does  not  exist  as  a  separate  ion  in  the  organic 
molecules,  still  has  a  considerable  tendency  to  draw  negative  charge  to  it, 
pulling  the  shared  electrons  away  from  the  methyl  group.  Thus  the 
dipole  moment  of  methyl  alcohol  is  1.73  X  10~18,  almost  as  large  as  that 
of  methyl  chloride  CH8C1.  Instead  of  being  bound  to  one  organic  group 
and  one  hydrogen,  as  in  the  alcohols,  the  oxygen  may  join  to  two  simple 
organic  groups,  forming  an  ether,  like  dimethyl  ether  (CH3)O(CH3), 
diethyl  ether  (C2H5)O(C2H5),  etc.  Here  the  organic  groups  come  off 
from  the  oxygen  more  or  less  at  tetrahedral  anglos.  The  oxygen  in  the 
ethers  also  has  some  tendency  to  draw  negative  charge  to  itself,  so  that 
the  ethers  have  a  dipole  moment,  though  not  quite  so  large  as  in  the 
alcohols.  It  is  plain  -from  this  discussion  that  the  alcohols  and  ethers 
can  in  a  way  be  derived  from  water  by  replacing  one  or  both  of  the 
hydrogens  by  methyl,  ethyl,  or  more  complicated  groups.  The  carbon- 
oxygen  distances  in  these  compounds  are  about  1.44  A  and  the  angles 
between  the  bonds  are  roughly  the  tetrahedral  angle,  though  there  are 
considerable  variations  both  in  distance  and  in  angle  from  one  compound 
to  another. 

As  the  alcohols  and  ethers  can  be  derived  from  water  by  replacing 
one  or  both  of  the  hydrogens,  so  the  amines  come  from  ammonia  by 
replacing  one,  two,  or  three  of  the  hydrogens  by  organic  groups.  In 
Table  XXI V-l,  for  instance,  we  give  properties  of  methylamine  CHsNH2, 
dimethylamine  (CHs^NH,  and  trimethylamine  (CHa^N.  Evidently 
a  complicated  set  of  compounds  can  be  built  up  in  this  way,  using  more 
and  more  complicated  groups  to  tie  to  the  nitrogen  atom. 

One  of  the  most  important  organic  radicals  is  the  carboxyl  group, 
COOH.  This  group  attaches  itself  by  a  single  bond  to  any  carbon  atom, 
forming  an  organic  acid.  For  instance,  acetic  acid  has  the  structure 

H     .0: 
H:C:C"      .    That  is,  one  oxygen  is  held  to  the  carbon  by  a  double 

H:O:H 

bond,  the  other  by  a  single  bond,  so  that  this  latter  can  also  attach  itself 
to  a  hydrogen.  Even  simpler  is  formic  acid,  HCOOH.  The  conspicuous 


428  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXV 

tendency  of  the  carboxyl  group  is  for  it  to  lose  the  H  as  a  positive  ion, 
leaving  the  remainder  of  the  molecule  as  a  real  negative  ion,  as  for  exam- 
ple (CH8CQO)~.  Then  a  metallic  ion,  for  instance  sodium,  can  attach 
itself,  forming  for  instance  sodium  acetate,  (CH3COO)~Na+.  While  we 
have  written  this  as  if  it  were  a  real  ionic  compound,  to  emphasize  this 
quality  more  than  in  most  organic  compounds,  still  such  a  substance  is 
not  as  definitely  ionic  as  in  the  inorganic  salts.  For  the  sodium  in  this 
case  furnishes  a  single  electron,  just  like  hydrogen,  so  that  we  can  consider 
it  as  forming  a  homopolar  bond  with  the  oxygen.  In  sodium  acetate, 
for  instance,  we  could  write  the  structure  just  as  in  acetic  acid  above, 
with  the  replacement  of  the  hydrogen  by  sodium.  The  distinction 
between  this  way  of  writing  it  and  the  ionic  way  is  simply  a  matter  of 
degree,  depending  on  how  much  the  shared  electrons  between  the  sodium 
and  the  oxygen  are  really  shared  (the  homopolar  interpretation)  or  how 
much  they  are  definitely  held  to  the  oxygen  (the  ionic  interpretation). 
The  compounds  formed  in  this  way,  by  replacing  the  H  in  the  carboxyl 
group  by  a  metal,  are  called  the  esters  and  are  a  very  important  group  of 
compounds. 

It  is  obvious  that  in  a  section  such  as  this,  it  is  impossible  to  do  more 
than  give  a  cursory  notice  to  a  very  few  of  tho  many  important  organic 
radicals.  All  we  have  hoped  to  do  is  to  give  some  idea  of  the  principles 
of  valence  leading  to  the  bonding.  Those  same  principles  continue  to 
be  a  guide  in  the  more  complicated  radicals  as  well.  At  least,  one  can  get 
some  idea  of  the  way  in  which,  by  attaching  any  one  of  a  great  number  of 
radicals  to  any  of  the  possible  points  of  attachment  in  a  carbon  chain 
structure,  one  can  get  an  enormous  number  of  compounds,  as  one  actually 
finds  in  organic  chemistry. 

3.  The  Double  Bond,  and  the  Aromatic  Compounds. — Ethylene,  with 

H.        .H 
structure     C::C     ,   and   acetylene,    with  structure   H:C:::C:H,  are 

H-        'H 

elementary  examples  of  the  double  and  triple  bonds  between  carbons. 
The  carbon-carbon  distances  are  1.34  A  and  1.22  A,  in  comparison  with 
1.54  A  for  single  bonds.  This  illustrates  the  general  rule  that  inter- 
nuclear  distances  are  less  with  double  bonds  than  with  single,  and  still 
less  with  triple  bonds.  There  is  one  feature  of  interest  in  ethylene. 
Unlike  the  situation  with  the  single  bond  in  ethane,  there  is  not  the 
possibility  of  free  rotation  about  the  double  bond.  That  is,  the  hydrogens 
all  tend  to  lie  in  a  plane,  as  our  structural  formula  would  indicate.  This 
tendency  is  true  in  general  with  double  bonds. 

The  best-known,  and  at  the  same  time  the  most  puzzling,  double 
bonds  occur  in  the  benzene  ring,  which  is  the  foundation  of  the  aromatic 
compounds.  Benzene,  CeHc,  is  a  plane  hexagonal  structure,  as  shown  in 


SEC.  3] 


ORGANIC  MOLECULES  AND  THEIR  CRYSTALS 


429 


Fig.  XXV-7.  The  carbon-carbon  distance  is  1.39  A,  definitely  less  than 
the  1.54  A  spacing  in  the  chain  compounds,  but  slightly  greater  than  the 
1.34  found  with  double  bonds  in  ethylene.  Each  carbon  is  bonded  to 
only  three  other  atoms,  rather  than  four, 
and  these  three  lie  in  a  plane  at  angles 
of  120°  to  each  other,  rather  than  being 
arranged  tetrahedrally  in  space.  Car- 
bon in  such  a  compound  seems  to  behave 
definitely  differently  from  its  behavior 
in  the  aliphatic  compounds.  Something 
of  a  guide  to  the  interpretation  is  fur- 
nished, however,  by  the  structure  of 
graphite,  a  form  of  carbon  having  the 
same  relation  to  the  aromatic  compounds 
that  diamond  has  to  the  aliphatic  com- 
pounds. The  graphite  structure  is 
shown  in  Fig.  XXV-8.  It  will  be  seen 
that  the  atoms  are  arranged  in  sheets, 
forming  regular  hexagons  similar  to  the 
benzene  ring  in  the  sheets.  Not  only 

are  the  hexagons  of  the  same  shape  as  the  ring  but  they  are  of 
almost  the  same  size;  the  carbon-carbon  spacing  between  neighbors  in  a 
sheet  is  1.42  A,  very  slightly  larger  than  the  value  1.39  A  in  benzene. 


FIG.  XXV-7.— The  benzene  molecule, 
C6H6. 


FIG.  XXV-8. — The  graphite  structure. 

The  sheets,  however,  are  3.4  A  apart,  a  much  greater  distance.  This  is 
as  if  the  valence  bonds  acted  wholly  within  the  separate  sheets,  and  only 
Van  der  Waals  forces,  or  similar  weak  forces,  held  the  sheets  together  and 
were  not  able  to  pull  them  very  close  together.  This  impression  is  made 
stronger  by  the  physical  properties  of  graphite.  It  is  a  very  soft  material 


430  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXV 

and  a  good  lubricant,  and  the  lubricating  properties  arise  because  one  of 
the  sheets  slides  over  another,  indicating  that  the  forces  between  sheets 
are  small  and  easily  overcome.  It  is  strongly  suggested  that  the  graphite 
structure  is  similar  to  the  benzene  ring,  the  carbon  being  in  the  same  form 
in  both  substances.  In  harmony  with  this  point  of  view,  it  is  found  that 
benzene  rings  can  join  together  the  same  way  that  the  hexagons  do  in 
graphite.  Thus  if  we  join  two  hexagons  we  have  naphthalene,  CioH8, 
shown  in  Fig.  XXV-9.  More  and  more  rings  can  continue  to  join  up  in 
this  way,  so  that  we  have  the  same  possibility  of  continuous  extension  and 
complication  with  the  aromatic  compounds  that  we  have  with  the  ali- 
phatic ones  with  their  carbon  chains. 


FIG,  XXV-9.  —  The  structure  of  naphthalene, 


Not  only  can  benzene  rings  join  each  other,  but  also  the  hydrogens 
can  be  replaced  by  various  organic  radicals.  Thus,  for  instance,  one 
hydrogen  in  benzene  can  be  replaced  by  an  OH  group,  as  in  the  alcohols 
among  the  aliphatic  compounds.  The  resulting  compound,  CeH6OH,  is 
phenol,  or  carbolic  acid.  Or  one  hydrogen  can  be  replaced  by  a  carboxyl 
group,  COOH,  resulting  in  benzoic  acid  C6H6COOH.  If  the  hydrogen 
in  the  carboxyl  is  replaced  by  a  metallic  atom,  we  have  a  benzoate  of 
the  corresponding  metal,  A  great  number  of  other  radicals  can  replace 
the  hydrogens.  And  we  observe  that  if  more  than  one  of  the  hydrogens 
of  a  ring  is  replaced  by  some  other  radical,  there  are  a  number  of  possible 
positions  for  the  radicals  on  the  ring.  Thus  if  two  of  the  hydrogens  of  a 
benzene  ring  are  replaced  by  chlorines,  we  have  dichlorbenzene,  with 
three  possible  structures,  as  indicated  and  named  in  Fig.  XXV-10.  The 
three  types  are  called  orthodichlorbenzene,  metadichlorbenzene,  and 
paradichlorbenzene,  according  to  the  position  of  the  second  chlorine  with 


SBC.  3] 


ORGANIC  MOLECULES  AND  THEIR  CRYSTALS 


431 


respect  to  the  first.  Three  such  compounds  can  often  have  quite  different 
properties.  In  particular,  it  is  plain  that  in  the  case  shown  the  para- 
compound  is  symmetrical  and  can  have  no  dipole  moment,  while  the 
ortho-  and  meta-  compounds  are  quite  unsymmetrical  and  have  a  con- 
siderable dipole  moment,  since  the  chlorine  as  usual  pulls  a  good  deal  of 
negative  charge  to  itself. 


(a)  ortho- 

(c)  para- 

Fio.  XXV-10. — The  three  forms  of  dichlorbenzene.     (a)  ortho-,  (6)  mota-, 
(c)  paradichlorbenzene. 


H 


H.   ./ 

•C* 


H 

•  • 

C 


•  c  ' 


»H 


.C. 


H'.C/C..%C-'H 


H 
(a) 


H 
(b) 


\J 

V 
(c) 

Fio.  XXV-11. — Three  alternative  valence  structures  for  benzene. 

The  valence  properties  of  the  carbon  in  the  benzene  ring  do  not  at 
first  sight  fit  in  with  our  rules  for  determining  valence.  To  explain  the 
valence  by  homopolar  bonds,  we  must  assume  that  each  carbon  is  bound 
to  one  of  its  neighboring  carbons  by  a  single  bond,  to  the  other  one  by  a 
double  bond,  and  to  the  hydrogen  or  other  radical  by  a  single  bond.  Two 
possible  ways  of  doing  this  are  shown  in  Fig.  XXV-11  (a)  and  (6).  This 
structure  would  suggest  that  the  hexagon  should  not  be  a  regular  one,  but 


432 


INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XXV 


that  instead  the  pairs  of  carbons  held  by  double  bonds  should  be  closer 
together  than  those  held  by  single  bonds.  Experimentally  this  is  not  the 
case,  however.  X-ray  and  electron  diffraction  methods  show  that  the 
hexagon  is  a  regular  one.  Furthermore,  if  there  were  real  difference 
between  the  single  and  double  bonds,  a  molecule  with  two  radicals  in  the 
ortho-position  and  a  double  bond  between  would  be  different  from  a 
molecule  with  the  same  two  radicals  in  the  same  position  but  with  a 
single  bond  between.  No  such  difference  is  observed.  For  all  these 
reasons  one  concludes  that  the  structures  of  Fig.  XXV-11  (a)  and  (6)  do 
not  correspond  exactly  with  the  facts.  It  is  generally  considered  that 
the  true  state  of  affairs  is  a  sort  of  combination  of  the  two  structures  of 
(a)  and  (6),  in  which  the  pair  of  electrons  forming  a  double  bond  is  some- 


FIG.  XXV-12. — Structure  of  tho  naphthalene  crystal  as  determined  by  x-ray  analysis. 
(From  J.  Monteath  Robertson,  Science  Progress  32,  246  (1937)  by  permission  of  the  author.) 

times  found  as  in  one  of  the  structures,  sometimes  as  in  the  other.  This 
may  perhaps  be  symbolized  as  in  (c)  of  the  figure,  where  the  extra  electrons 
are  shown  in  a  position  where  they  could  contribute  somewhat  to  both 
bonds.  This  valence  structure  of  benzene,  though  it  does  not  fit  in  with 
ordinary  ideas  of  valence  very  satisfactorily,  proves  to  be  very  well 
explained  in  terms  of  the  quantum  theory. 

4.  Organic  Crystals. — Most  organic  compounds  are  definitely  formed 
of  molecules,  arranged  in  some  fairly  closely  packed  form  in  the  crystals. 
As  with  the  simpler  homopolar  compounds,  the  molecules  are  held 
together  largely  by  Van  der  Waals  forces,  though  in  some  cases  where 
the  molecules  have  dipole  moments,  there  are  electrostatic  forces  as  well. 
The  intermolecular  forces,  as  with  the  simpler  compounds,  are  small 
compared  to  the  energy  of  dissociation  in  most  cases,  and  the  distances 
between  molecules  are  large  compared  to  the  interatomic  distances 


SEC.  4] 


ORGANIC  MOLECULES  AND  THEIR  CRYSTALS 


433 


within  a  single  molecule.  We  shall  not  attempt  any  cataloguing  of  the 
different  types  of  crystal  structure  found  in  the  crystals;  almost  every 
compound  has  its  own  form  of  structure.  Often  they  are  complicated  on 
account  of  the  complicated  shape  of  the  molecules.  As  a  single  example, 
we  show  in  Fig.  XXV-12  a  diagram  of  the  naphthalene  crystal.  This  is 
the  direct  result  of  the  x-ray  diffraction  experiments  and  shows  a  projec- 
tion of  the  crystal  along  a  certain  axis,  by  giving  contour  lines  indicating 
the  density  of  electric  charge.  The  naphthalene  molecules  stand  out 
clearly,  and  we  see  by  comparison  with  Fig.  XXV-9,  in  which  the  double- 
ring  structure  of  this  molecule  is  shown,  that  the  plane  of  the  molecule  is 
inclined  to  the  plane  of  the  paper.  The  other  organic  crystals  which  havo 
been  investigated'  all  have  the  same  general  properties  of  representing 
fairly  closely  packed  structures  of  distinct  molecules. 

The  general  remarks  regarding  energy  which  we  have  made  in  Chaps. 
IX  and  XXII  apply  here  also.  The  interatomic  forces  within  the  mole- 
cules can  be  well  represented  by  Morse  curves,  as  in  Chap.  IX,  Sec.  1. 
In  Table  XXV-2  wo  givo  values  of  D  and  re  for  the  most  important  bonds 

TABLE  XXV-2. — CONSTANTS  OF  BONDS  IN  ORGANIC  MOLECULES 


Substance 

Bond 

D,  kg.-cal. 

D,  electron 
volts 

ret  A 

Chain  hydrocarbon.  .    .    . 
Chain  hydrocarbon  
Alcohols  and  ethers 
Amines        .  .     .    . 

C—  H 
C—  C 
C—  0 
C—  N 

101 
83 
83 
67 

4.34 
3.60 
3.6 

2.88 

1.1 
1.54 
1.44 

Fluorine-subst.  hydrocarbons 
Chlorine-subst.  hydrocarbons  
Brominc-subst.  hydrocarbons     . 
lodine-subst  hydrocarbons 

C—  F 
C—  Cl 
C—  Br 
C—  I 

125 
79 
66 
51 

5.40 
3  41 
2.83 
2  2 

1  36 
1.77 
1.93 
2  28 

Carbon  double  bond     

C::C 

150 

6.46 

1.34 

Carbon  triple  bond         

C:::C 

200 

8.7 

1.22 

Bond  energies  taken  from  Pauling,  J.  Am.  Chem.  floe.,  54,  3570  (1932). 

occurring  in  organic  compounds,  similar  to  the  values  for  diatomic  mole- 
cules given  in  Table  IX-1.  In  contrast  to  these  high  values  of  heats  of 
dissociation,  the  heats  of  vaporization  of  the  simpler  compounds  are 
small.  Thus  in  Table  XXV-3  we  give  heats  of  vaporization,  again  in 
kilogram-calories  per  mole,  of  some  of  the  organic  compounds  we  have 
mentioned.  The  latent  heats,  as  we  see  from  the  table,  are  of  the  same 
order  of  magnitude  as  those  for  inorganic  compounds  shown  in  Table 
XXIV-3.  The  alcohols  stand  out  as  having  rather  high  latent  heats,  for 
the  same  reason  that  water  does:  their  OH  groups  tend  to  hold  the  mole- 
cules together  by  electrostatic  attraction.  In  addition,  there  is  a  general 
trend  toward  higher  latent  heats  as  we  go  down  the  list.  This  is  natural, 


434  INTRODUCTION  TO  CHEMICAL  PHYSICS        [CHAP.  XXV 

TABLE  XXV-3. — LATENT  HEATS  OF  VAPORIZATION  OF  ORGANIC  COMPOUNDS 


Substance 

Formula 

Latent  heat 

Methane  

CH4 

2.3 

Ethane  

C2Hfl 

3.9 

Methyl  chloride  .         .                 .... 

CH8C1 

4.7 

Methyl  alcohol 

CHaOH 

9  2 

Ethyl  alcohol  

C2H6OH 

10.4 

Propane  

C3H8 

4.5 

Chloroform  

CHC1, 

8.0 

Acetic  acid  

CHsCOOH 

5.0 

iso-Butane  

CH(CHs)3 

5.5 

T&-Butane  

\u/4jtl.io 

5.6 

Diethyl  ether  

(C2H6)20' 

7.0 

Naphthalene  

Ciolls 

9.7 

w^Decane  

^IO'H'22 

8.5 

Latent  heats  are  in  kilogram-calories  per  mole,  taken  from  Landolt's  Tables,  and  are  for  the  lowest 
available  temperature  in  each  case. 

for  the  substances  of  higher  melting  and  boiling  points  lie  toward  the 
end  of  the  list,  and  if  the  entropy  of  melting  is  approximately  constant  for 
the  various  substances,  as  it  is  found  to  be,  we  see  that  the  greater  the 
melting  or  boiling  point,  the  higher  the  corresponding  latent  heat.  Aside 
from  these  points,  the  important  thing  to  observe  in  comparing  Tables 
XXV-2  and  XXV-3  is  the  fact  that  heats  of  vaporization  are  much  less 
than  heats  of  dissociation,  so  that  molecules  in  organic  compounds  tend 
to  have  separate  and  independent  existence. 


CHAPTER  XXVI 
HOMOPOLAR  BONDS  IN  THE  SILICATES 

In  the  preceding  chapters  we  have  discussed  the  general  nature  of 
the  homopolar  bond  and  have  seen  how  it  operates  in  certain  inorganic 
molecules  and  in  the  organic  compounds.  In  the  present  chapter  we 
shall  continue  the  discussion  to  include  the  silicates,  the  substances  form- 
ing the  foundation  of  a  great  many  of  the  rocks  and  minerals.  Silicon  and 
oxygen,  like  carbon,  can  form  continuous  chains,  and  in  this  way  can  build 
up  a  skeleton  structure  for  a  great  variety  of  compounds.  First  we  shall 
describe  this  chain  structure  and  then  we  shall  go  on  to  show  how  it  is 
found  in  substances  of  many  different  sorts. 

1.  The  Silicon-Oxygen  Structure. — Silicon  has  four  outer  electrons, 
the  same  as  carbon.  Like  carbon,  it  can  form  four  homopolar  bonds, 
arranged  in  a  tetrahedral  structure.  Thus  we  have  SiH4,  mentioned  in 
the  preceding  chapter,  and  substitution  products  like  SiF4,  SiCl4,  etc. 
But  the  characteristic  compounds  are  those  in  which  oxygen  atoms  are 
held  by  the  four  bonds.  Since  oxygen  is  divalent,  each  oxygen  must  also 
be  bonded  to  something  else,  and  the  simplest  case  is  that  in  which  each 
oxygen  is  bonded  also  to  a  hydrogen,  forming  Si(OH)4,  with  the  structure 

H 

:0: 

H:0:Si:O:H. 

:9: 
H 

In  this  structure,  orthosilicic  acid,  the  oxygens  surround  the  silicons 
tetrahedrally,  and  a  hydrogen  is  somewhat  loosely  attached  to  each 
oxygen.  If  the  hydrogens  are  detached  as  positive  ions,  we  leave  behind 
the  orthosilicate  radical  (SiO4)~~4,  a  tetrahedral  structure  similar  geometri- 
cally to  the  sulphate  (804) —  and  perchlorate  (C1O4)~  ions  mentioned  in 
Chap.  XXIII.  The  orthosilicate  radical  occurs  as  a  negative  ion  in  the 
orthosilicates,  typical  ionic  crystals  in  which  the  orthosilicate  ion,  and 
certain  positive  ions,  form  an  ionic  lattice.  An  example  is  magnesium 
orthosilicate,  or  olivine,  Mg2(Si04),  a  regular  structure  of  Mg++  ions  and 
orthosilicate  radicals.  In  this,  as  in  practically  all  silicates,  the  silicon- 
oxygen  distance  is  approximately  1.60  A.  This  can  be  compared  with  the 
carbon-oxygen  distance  of  1.44  A  found  with  single  bonds  in  the  alcohols 

435 


436 


INTRODUCTION  TO  CHEMICAL  PHYSICS       [CHAP.  XX VT 


and  ethers,  and  seems  to  indicate  a  similar  bonding  in  this  case,  the  silicon 
atom  of  course  being  larger  than  the  carbon  atom. 

There  is  no  reason  why,  instead  of  losing  its  hydrogen,  one  of  the  oxy- 
gons attached  to  a  silicon  cannot  attach  itself  to  another  silicon.     This 

:0:    :0: 
leads  to  the  structure  (Si2O7)~6,  with  valence  structure  :O:Si:O:Si:O:. 


(sio4y 


In  Fig.  XXVI-1  we  show  the  (SiO4)~4  and  (8i2O7)-6  ions,  which  are  found 
in  ionic  crystals,  occurring  in  nature  as  minerals.  The  positive  ions 
found  associated  with  them  include  a  variety  of  metallic  ions,  such  as 

Ca++,  Zn++  Na+,  Mg++,  A1+++  etc.  We 
shall  not  go  into  the  crystal  structure  of 
these  substances,  for  they  are  rather  com- 
plicated for  the  most  part,  and  yet  do  not 
show  any  new  principles  of  structure  that 
we  have  not  already  met  in  our  discussion 
of  ionic  crystals,  with  one  exception.  This 
exception  is  the  fact  that  the  natural 
minerals  show  a  rather  strong  tendency 
to  substitute  one  positive  ion  for  another. 
That  is,  the  minerals  are  not  always  of  the 
same  chemical  composition.  Often  a  posi- 
tive ion  is  missing  from  the  lattice,  and  its 
place  taken  by  another  positive  ion,  with- 
out seriously  distorting  the  lattice.  Such 
substitutions  take  place  only  when  the  two 
ions  in  question  have  about  the  same  ionic 
radii.  For  instance,  from  the  table  of  ionic 
radii,  Table  XXIII-2,  we  find  that  Ca++ 
has  a  radius  of  0.95  A,  and  Na+  a  radius  of  1.05  A.  These  are  near  enough 
alike  so  that  calcium  and  sodium  can  substitute  for  each  other  in  the 
lattice.  Of  course,  these  ions  have  different  charges,  and  such  a  substitu- 
tion would  upset  the  electrostatic  equilibrium,  resulting  in  a  crystal 
having  a  net  electric  charge.  This  is  not  allowed,  and  to  balance  it  other 
substitutions  are  always  made.  There  is,  for  instance,  a  silicate  contain- 
ing Ca"1"*  or  Na+  interchangeably,  and  Mg4""*"  and  A1"H4'  interchangeably. 
By  having  suitable  relations  between  the  amounts  of  these  four  ions,  the 
crystal  can  always  be  kept  electrically  neutral  and  yet  may  have  a  variable 
composition.  Such  a  situation  is  very  often  the  case  in  minerals. 

2.  Silicon-oxygen  Chains.  —  In  Fig.  XXVI-1,  we  have  seen  how 
two  or  more  silicons  can  be  joined  together  by  oxygens.  This  process  in 
the  ideal  case  can  continue  indefinitely,  forming  endless  chains  and  in 
practice  forming  chains  of  varying  length.  The  simplest  such  chain  is 


FIG.  XXVI-1. —Structure  of  silicate 
ions. 


SBC.  2] 


HOMOPOLAR  BONDS  IN  THE  SILICATES 


437 


that  in  which  each  silicon  is  joined,  through  oxygens,  to  two  other  silicons. 
This  is  shown  in  Fig.  XXVI-2.    The  valence  structure  can  be  symbolized 


:Si:6:Si::Si::Si  etc.     This  structure,  like  the  orthosilicate  ion,  is 

:0:":0:":0:":0: 

electrically  charged  and  is  sometimes  called  the  metasilicate  ion.     A 

single  unit  of  the  structure  evidently  has  the  structure  (SiO3)     . 


FIG.  XXVI-2. — Section  of  metasilicate  chain,  (SiO8) — .     (a)  View  looking  directly  down. 

(6)  Perspective. 

Each  unit  of  the  metasilicate  ion  has  two  oxygens  carrying  negative 
charges,  which  are  able  to  form  homopolar  bonds  with  other  atoms.  The 
simplest  way  they  can  do  this  is  to  take  up  hydrogens,  forming  bonds 
that  are  partly  polar,  but  mostly  homopolar,  with  the  hydrogens,  and 
resulting  in  metasilicic  acid,  H2SiO3.  It  is  evident  that  the  single  unit 
H2SiO3  has  no  separate  existence;  the  whole  structure  arises  when  the 
silicates  are  joined  together  to  form  chains.  Similarly  the  oxygens  might 
take  up  alkali  ions,  for  example  sodium,  forming  sodium  metasilicate, 
Na2SiOa,  again  made  out  of  endless  chains.  If  the  chains  are  finite,  as  of 
course  they  really  would  be,  the  proportion  of  Na  increases,  until  finally 
if  the  chains  are  only  one  silicon  long,  we  have  sodium  orthosilicate, 
Na4SiO4,  which  we  have  mentioned  before.  There  is  the  possibility  of  a 
continual  variation  from  one  of  these  limiting  cases  to  the  other,  as  the 
average  length  of  the  chain  changes.  These  substances,  intermediate 


438  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXVI 

between  Na2SiO8  and  Na4Si04,  have  remarkable  properties.  They  are 
commonly  called  water  glass.  If  dissolved  in  water  and  left  to  set,  they 
form  a  viscous,  jellylike  mass.  The  explanation  is  the  same  that  we  gave 
in  the  preceding  chapter  for  the  viscosity  of  the  long  chain  hydrocarbons: 
the  long  stringlike  molecules,  on  account  of  the  possibility  of  rotation 
about  the  bonds  joining  neighboring  atoms,  are  flexible  and  get  tangled 
up  in  each  other.  Similarly  metasilicic  acid,  H2SiO8,  in  water,  forms  the 
remarkable  substance  known  as  silica  gel.  This  again  is  a  jellylike 
material,  getting  stiffer  and  stiffer  as  it  is  allowed  to  set.  It  is  remark- 
able, like  all  jellies,  in  the  very  small  amount  of  material  required  to 
stiffen  up  a  large  mass  of  water.  Two  per  cent  or  so  of  EUSiOs  in  water 
will  transform  the  whole  mass  into  a  gel.  Presumably  the  action  is 
somewhat  more  complicated  than  simply  the  entanglement  of  the  stringy 
molecules  with  each  other.  A  hydrogen  atom  of  one  chain,  carrying  a 
net  positive  charge,  can  happen  to  come  close  to  an  oxygen  of  another  with 
its  negative  charge.  The  resulting  electrostatic  attraction,  holding  the 
chains  rather  tightly  together,  is  analogous  to  the  attraction  of  hydrogens 
for  oxygens  found  in  water,  discussed  in  Sees.  3  and  4,  Chapter  XXIV. 
By  such  bonding  of  one  chain  to  another,  the  structure  could  be  stiffened 
greatly,  so  that  it  would  take  on  some  of  the  properties  of  a  solid.  Of 
course,  with  only  occasional  joinings  such  as  we  could  have  with  a  small 
percentage  of  metasilicic  acid  in  a  great  deal  of  water,  the  rigidity  would 
be  much  less  than  in  a  real  solid,  but  that  is  just  the  characteristic 
behavior  of  jellies,  which  act  solid  and  yet  can  be  greatly  distorted  by  a 
small  force.  These  inorganic  jellies  are  interesting  in  that  they  probably 
show  us  what  is  the  essential  structure  of  the  organic  jellies  like  gelatin,  for 
instance.  They  are  of  great  practical  importance,  in  that  they  furnish 
the  fundamental  structure  in  cement  and  concrete.  They  can  be  pre- 
pared in  water  as  fairly  fluid  materials  but  become  gradually  stiffer  as  the 
chains  join  together,  and  as  water  evaporates  out  of  them. 

Metasilicate  chains  of  varying  length  are  found  not  only  in  jellylike 
materials,  but  also  in  crystals  and  minerals.  In  such  cases,  the  chains 
are  straightened  out  and  really  become  indefinitely  long,  running  parallel 
to  each  other  through  the  whole  length  of  the  crystal.  Positive  ions  are 
arranged  in  regular  positions  between  the  chains,  so  as  to  make  the  whole 
structure  electrically  neutral.  An  example  is  diopside,  CaMg(Si()3)2. 
These  materials  are  really  ionic  crystals,  in  which  the  negative  ions, 
instead  of  being  of  atomic  dimensions,  are  lengthened  out  in  one  direction 
to  be  of  large-scale  size. 

More  important  than  the  simple  metasilicate  chain,  in  the  structure 
of  crystals,  is  the  double  chain  found  in  the  amphiboles.  This  structure 
is  shown  in  Fig.  XXVI-3,  looking  down  on  the  top.  A  unit  of  the  chain 
contains  four  silicons,  eleven  oxygens,  as  we  see  by  counting,  and  it  has 


SBC.  3]  HOMOPOLAR  BONDS  IN  THE  SILICATES  439 


six  oxygens  capable  of  forming  bonds,  so  that  its  structure  is 
This  chain,  unlike  that  of  the  metasilicates,  has  a  type  of  cross-bracing 
which  makes  it  quite  rigid.  Thus  substances  of  this  structure  do  not  form 
jellies  or  viscous  fluids,  since  these  depend  on  the  flexibility  of  the  chains. 
The  characteristic  form  of  materials  containing  the  double  chain  is  the 
crystalline  form,  the  chains  extending  parallel  to  each  other  through  the 
crystal,  and  the  interstices  filled  in  with  positive  ions  in  such  a  way  as  to 
make  the  structure  neutral  and  hold  it  together.  The  ionic  forces  holding 
the  crystal  together,  however,  are  not  so  strong  as  the  homopolar  forces 
acting  between  atoms  within  the  chain.  Thus  mechanical  forces  insuffi- 
cient to  break  the  chains  can  pull  them  apart  and  separate  them  into 
fibers.  Such  a  fibrous  structure  is  characteristic  of  the  amphiboles  and 
the  double  chain  compounds.  The  most  familiar  example  is  asbestos, 


FIG.  XXVI-3. — Double  silicate  chain,  view  looking  directly  down. 

a  very  strongly  bound  material  with  high  melting  point,  and  yet  one  in 
which  the  fibers  can  be  separated  from  each  other  with  ease.  These 
simple  fibrous  materials  are  interesting  in  that  they  suggest  the  way  in 
which  the  more  complicated  fibers  of  organic  chemistry  are  built  up. 
They  also  contain  chains,  generally  double  or  more  complicated  chains 
to  get  the  necessary  mechanical  rigidity,  the  atoms  and  bonds  concerned 
being  typical  of  organic  compounds.  Since  even  the  simplest  of  naturally 
fibrous  materials  are  quite  complicated,  we  shall  give  no  examples  of 
organic  fibres. 

3.  Silicon-Oxygen  Sheets. — The  chain  structure  of  Fig.  XXVI-3  can 
be  extended  to  form  a  two-dimensional  sheet,  as  shown  in  Fig.  XXVI-4. 
A  unit  of  this  structure  has  the  composition  (Si2OB)  ,  but  as  with  the 
chains  this  individual  unit  has  no  existence  by  itself.  Sheets  of  this  kind 
are  found  in  the  structure  of  mica.  This  mineral  consists  of  silicon- 
oxygen  sheets  piled  parallel  to  each  other,  alternating  with  sheets  contain- 
ing metallic  ions,  as  Al,  Mg,  etc.,  along  with  oxygen  and  hydrogen.  The 
sheets  are  stronger  than  the  ionic  bonds  holding  them  together,  so  that 
mechanical  force  easily  cleaves  the  crystal  into  thin  sheets.  This  very 
characteristic  property  of  mica  thus  follows  directly  from  its  valence 


440  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXVI 

structure.  Of  course,  it  should  not  be  thought  that  one  can  really  cleave 
the  crystal  down  to  single  silicon-oxygen  sheets;  this  would  demand  much 
more  delicate  technique  than  can  be  used.  But  the  tendency  to  split  one 
sheet  from  another  is  great  enough  so  that  cleavage  easily  occurs  along 
these  planes. 

Another  very  important  group  of  materials,  the  clays,  are  formed 
from  silicon-oxygen  sheets.  In  this  case,  however,  the  sheet  is  not  the 
simple  one  shown  in  Fig.  XXVI-4,  but  each  sheet  is  bound  by  homopolar 
valences  to  another  sheet  of  different  composition.  For  instance,  in 
kaolinite,  one  form  of  clay,  one  starts  with  a  silicon-oxygen  sheet.  Above 
that,  there  is  an  aluminum  layer,  with  as  many  aluminums  as  silicons, 
bonded  to  the  oxygens.  The  other  bonds  of  the  aluminums  extend  to  a 


FIG.  XXVI-4. — Silicon-oxygen  sheet. 

still  higher  sheet  of  oxygens,  and  finally  these  oxygens  hold  hydrogens  by 
their  unused  valences.  This  makes  an  electrically  neutral  sheet  with  no 
unshared  homopolar  bonds,  so  that  successive  sheets  are  held  to  each 
other  only  by  polarization  and  Van  der  Waals  forces.  In  this  they 
resemble  the  sheets  of  graphite,  described  in  Sec.  3  of  the  preceding  chap- 
ter. Like  graphite,  the  clays  have  a  soft  and  plastic  structure,  resulting 
from  the  ease  with  which  one  sheet  can  slide  over  another.  They  are 
quite  different  from  mica,  in  which  the  bonds  between  sheets,  being 
ionic,  hold  the  sheets  comparatively  firmly  together.  The  clays  are  also 
different  from  mica  in  that  large  crystalline  sheets  cannot  be  obtained. 
It  has  been  suggested  that  this  may  be  because  the  two  sides  of  the  sheet, 
the  silicon  side  and  the  aluminum  side,  may  have  slightly  different  natural 
sizes,  so  that  one  sheet  may  be  under  compression,  the  other  under  ten- 
sion, in  the  material.  This  would  mean  that  thin  sheets  would  have  a 
tendency  to  pucker  and  break,  preventing  the  formation  of  large  sheets. 


SBC.  4)  HOMOPOLAR  BONDS  IN  THE  SILICATES  441 

4.  Three-dimensional  Silicon-Oxygen  Structures. — There  are  a 
number  of  ways  in  which  silicons  and  oxygens  can  be  joined  together  into 
three-dimensional  structures,  filling  all  space.  These  structures  are  all 
alike  in  that  each  silicon  is  surrounded  by  four  oxygens  and  each  oxygen 
by  two  silicons,  so  that  the  formula  is  Si02.  The  valence  bonds  are 
completely  used  up  in  holding  the  silicons  and  oxygons  together,  so  that 
there  is  no  possibility  of  adding  other  atoms  to  the  structure.  These 
materials,  then,  being  bound  in  all  directions  by  valence  bonds,  are  very 
strong  and  rigid,  something  like  diamond,  though  not  quite  so  hard 
Perhaps  the  simplest  of  these  struc- 
tures to  visualize  is  cristobalite, 
shown  in  Fig.  XXVI-5.  This 
structure  is  very  similar  to  the  dia- 
mond structure,  as  shown  in  Fig. 
XXV-6.  One  can  start  with  the 
diamond  structure,  place  a  silicon 
in  the  position  occupied  by  each 
carbon,  and  an  oxygen  midway 
between  each  silicon  and  each  of  its 
four  neighbors,  and  one  has  the 
cristobalite  structure.  This  is  a 
slightly  undesirable  structure  for 
the  atoms  to  form,  however,  from  Fla'  XXVI'5-Tho  -istobaiite  structure. 

the  standpoint  of  the  oxygens.  We  note  that  it  requires  the  two 
valences  of  the  oxygen  to  be  diametrically  opposite  each  other.  On 
the  other  hand,  as  we  have  seen  earlier,  oxygen  prefers  to  have  its 
two  valence  bonds  make  roughly  the  tetrahedral  angle  109°  with  each 
other.  It  is  likely,  as  a  matter  of  fact,  that  the  oxygens  in  cristobalite 
are  really  somewhat  out  of  the  line  joining  neighboring  silicons, 
just  in  order  to  have  their  two  valences  at  an  angle  with  each  other. 
They  oscillate  around  the  midpoints  of  the  lines  joining  silicon  atoms, 
however,  rather  than  being  found  always  on  one  side  or  the  other, 
as  one  tells  from  the  crystal  symmetry,  which  is  the  same  as  for  the 
diamond  lattice.  In  the  other  crystalline  forms  of  SiO2,  in  contrast  to 
cristobalite,  the  two  bonds  formed  by  each  oxygen  make  much  more 
nearly  the  angle  that  is  preferred.  The  best-known  form,  of  course,  is 
quartz.  This  is  a  decidedly  complicated  crystal,  atoms  being  arranged 
with  a  screwlike  symmetry,  as  is  shown  by  the  fact  that  quartz  rotates 
the  plane  of  polarization  of  polarized  light.  We  shall  not  give  its 
structure,  but  merely  point  out  that  as  in  other  cases  each  silicon  is 
surrounded  tetrahedrally  by  four  oxygens,  each  oxygen  by  two  silicons, 
and  that  the  angle  between  the  two  bonds  from  each  oxygen  is  about 
145°. 


442 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXVI 


From  the  fact  that  SiOa  exists  in  several  forms,  we  see  that  the  whole 
structure  of  this  material  is  not  determined  by  the  nearest  neighbors  of  a 
given  atom.  That  is,  one  can  have  each  silicon  surrounded  by  four 
oxygens,  each  oxygen  by  two  silicons,  in  many  different  ways.  This  fact 
is  responsible  for  the  amorphous  nature  of  fused  quartz,  the  simplest  form 
of  glass.  This  is  an  irregular  structure,  with  no  indication  of  a  repeating 
regularity.  And  yet  the  immediate  neighbors  of  any  given  atom  are 
arranged  as  they  are  in  cristobalite.  It  is  rather  hard  without  a  model 

to  see  how  this  can  be  done,  but  we  show  in 
Fig.  XXVI-6  a  two-dimensional  schematic 
representation  of  the  same  situation.  In 
this  case,  the  black  dots  represent  the  silicons, 
and  each  is  surrounded  by  three  circles  repre- 
senting oxygens  (instead  of  four  as  in  the 
three-dimensional  case),  while  each  oxygen 
is  between  two  silicons.  The  diagram  (6), 
representing  the  glass,  shows  that  by  only 
slight  distortions  of  the  bond  angles  of  the 
oxygens,  the  molecules  can  be  joined  in  a 
quite  irregular  structure.  Some  of  the  bonds 
in  the  glass  are  under  more  strain  than 
others,  and  are  fairly  easy  to  break.  Thus 
as  the  temperature  is  raised,  some  bonds  will 
break  and  the  material  will  lose  a  little 
strength.  If  bonds  break,  it  may  be  that 
they  can  join  again  with  a  different  arrange- 
FIO.  xxvi-6.— Schematic  repre-  ment,  resulting  in  permanent  distortion  of 
sentation  of  crystal  and  glass.  t}ie  material.  It  is  in  this  way  that  the  glass 
flows  at  high  temperatures.  There  is  no  sharp  melting  point  as  there  is 
with  a  crystal,  for  this  demands  that  in  all  parts  of  the  material  the  bonds 
weaken  at  the  same  temperature.  Rather  we  might  say  that  on  account 
of  strain  some  parts  of  the  glass  have  higher  melting  points,  some  lower, 
so  that  the  softening  is  gradual. 

The  actual  glasses  generally  contain  Na20,  or  other  constituents,  in 
addition  to  SiO2.  Now  if  we  look  back,  we  notice  that  the  proportion  of 
oxygen  to  silicon  is  higher  in  the  compounds  where  the  atoms  form  chains 
or  sheets  than  in  the  three-dimensional  lattice  of  SiO2.  The  reason  is 
that  the  chains  or  sheets  contain  some  oxygens  bonded  to  only  one  silicon 
rather  than  two,  the  other  bond  being  attached  to  some  other  atom.  By 
analogy,  we  see  that  in  a  soda  glass,  the  extra  oxygens  coming  from  the 
Na20  will  join  onto  silicons  and  will  introduce  some  oxygens  bonded  to 
only  one  silicon.  Obviously  the  other  bond  will  hold  the  oxygen  to 
sodium  atoms.  Such  a  glass,  then,  will  have  fewer  cross  links  or  bonds 


SBC.  4]  HOMOPOLAR  BONDS  IN  THE  SILICATES  443 

between  atoms  than  pure  8102  and  consequently  will  be  softer,  and  will 
begin  to  soften  at  a  lower  temperature.  These  properties  of  course  are 
characteristic  of  the  soft  glasses.  In  any  case,  however,  the  glass  struc- 
tures, like  the  SiC>2  structures,  have  a  very  strong  and  rigid  bracing,  so 
that  it  is  natural  that  they,  like  diamond,  which  also  has  valenco  binding 
in  three  dimensions,  should  form  very  hard  substances. 


CHAPTER  XXVII 
METALS 

Most  of  the  elements  are  metals.  There  are  six  inert  gases  and 
twelve  or  fourteen  definitely  electronegative  elements,  leaving  approxi- 
mately seventy  metallic  elements.  We  have  already  seen  how  they 
combine  with  negative  radicals  to  form  ionic  substances,  and  how  on 
occasion  they  become  bound  in  organic  compounds,  by  a  bond  partly 
ionic  and  partly  homopolar.  But  the  elements  themselves  take  the 
familiar  metallic  form,  and  mixtures  or  compounds  of  one  metal  with 
another  form  alloys,  similar  in  most  respects  to  pure  metals.  As  we  know, 
the  metals  are  fairly  strongly  bound  substances.  Their  melting  points 
are  rather  high,  mercury  and  gallium  being  the  only  ones  that  are  liquid 
near  room  temperature.  Thus  we  must  assume  that  there  are  fairly 
strong  bonds  between  the  atoms  of  the  metal.  As  we  saw  in  Chap. 
XXII,  these  bonds  resemble  the  homopolar  type,  but  it  is  generally 
considered  best  to  treat  them  as  in  a  class  by  themselves,  the  metallic 
bonds. 

1.  Crystal  Structures  of  Metals. — Metals  are  divided  fairly  definitely, 
though  not  perfectly  sharply,  into  two  groups:  the  ordinary  metals  and 
a  group  of  peculiar  metals  verging  on  the  nonmetals.  These  peculiar 
metals  come  at  the  ends  of  the  rows  of  the  periodic  table  and  include 
approximately  the  following:  C,  Si  (both  of  which  have  some  metallic 
properties,  though  we  have  previously  treated  them  as  nonmetals),  Ga, 
Ge,  As,  Se  (which  is  partly  metallic),  In,  Sn,  Sb,  Te,  Tl,  Bi.  The  ones 
we  have  just  named  are  peculiar  in  that  their  bonds  are  very  similar  to 
homopolar  bonds.  Thus  Si,  Ge,  Sn,  all  crystallize  in  the  diamond  struc- 
ture, like  carbon,  so  that  we  can  consider  that  they  have  four  homopolar 
valences  binding  them  to  their  four  nearest  neighbors.  Arsenic,  anti- 
mony, and  bismuth  have  peculiar  structures  in  which  each  atom  is  bound 
to  three  nearest  neighbors,  corresponding  to  the  three  homopolar  valences 
which  we  should  expect  these  elements,  like  nitrogen  and  phosphorus,  to 
form.  The  binding  joins  atoms  in  layers,  resulting  in  brittleness  and  a 
tendency  to  cleave  easily  along  the  layers.  And  selenium  and  tellurium, 
with  two  valences  like  oxygen  and  sulphur,  tend  to  form  chains,  each  atom 
being  bound  to  two  nearest  neighbors,  a  formation  found  alsx>  in  sulphur, 
though  not  in  oxygen. 

The  ordinary  metalsT  in  contrast  to  these  peculiar  elements,  do  na 
form  homopolar  valence  bonds  in  any  ordinary  sense.    Each  atom  has  to< 

444 


SBC.  1]  METALS  445 

many  nearest  neighbors  to  share  electron  pairs  with  each;  they  try^to 
share  electrons^with  their  neighbors  but  there  are  not  enough  to  go  around. 
This  has  9flY*!Tftl  *p«"1+-g-  In  the  first  place,  there  is  no  such  directional 
property  to  the  bonds  as  we  have  with  real  homopolar~compounds. 
Instead  of  tending  to  come  off  in  tetrahedral  directions  or  something  of 
that  sort,  the  metallic  bonds  can  attract  other  atoms  no  matter  in  what 
directions  they  may  lie.  Thus  the  atoms  act  as  if  they  were  spheres 
without  preferred  directions,  as  far  as  the  crystal  structure  is  concerned, 
and  the  ordinary  metals  crystallize  in  forms  that  are  characteristic  of 
close  packed  spheres.  Secondly,  since  the  valence  bonds  arc  shared,j*p 
to  speak,  between  several  pairs  of  atomsT  they  are  not  so  strong  as  real 
homopolar  valences.  Thiy?  a  metallic  crystal,  though  much  stronger 
than  one  held  togcther_hy  Van  der  ^Vqalfi  fniwn,  is  wnn-lror  t.lmn  nno,  Ifrn 
diamond  or  quartz,  held  fry  p^re  frQnfiopolar  y***™*™  Furthermore,  the 
fewer  electrons  an  atom  has  to  share  with  its  neighbors,  the  weaker  its 
bonds,  so  that  the  alkali  metals,  with  only  one  electron  each  outside 
RlQaed-flhclls.  are  the  weakest  metals  mechanically  and  the  lowest  molting 
pnes,  the  alkaline  earths  are  next,  and  sp  HP,  white  thn  iron,  puling  jnm,  anH 
platinum  groups  have  thojrtrongest  bonds.  Finally,  the  electrons  hold^ 
ing  the  metal  together  aTenot  localized  between  definite  pairs  of  atoms,  as 
they  are  in  homopolar  bondsT  but  can  ^ove  fl.rnimHJ 


Found  between  one  pain  of  atoms,  sometimes  between  another. 
result  of  this  freedom  of  motion  of  the  electrons  is  the  property  of 
conductivity,  the  most  charactgngtic  distinguishing  feature  of  f.hp. 
With  this  preliminary  view  of  metals,  we  shall  now  consider  their  crystal 
structure  and  later  go  on  to  some  of  their  other  properties. 

The  ordinary  metals  crystallize  in  one  of  three  structures:  the  facc- 
centered  cubic,  hexagonal,  or  body-centered  cubic.  Of  these,  the  face- 
centered  cubic  and  hexagonal  close-packed  structures  have  been  described 
in  Chap.  XXIV,  Sec.  4.  The  face-centered  cubic  structure  is  shown  in 
Fig.  XXIV-1  and  the  hexagonal  close-packed  structure  in  Fig.  XXIV-2. 
Sometimes  in  the  metals  the  latter  structure  is  slightly  distorted,  being 
elongated  along  the  vertical  axis  of  Fig.  XXIV-2,  so  that  it  is  no  longer 
a  close-packed  structure.  The  body-centered  cubic  structure  consists  of  a 
simple  cubic  lattice,  with  atoms  at  the  corners  of  the  cubes  and  also  at  the 
centers.  Thus  it  resembles  the  caesium  chloride  structure,  shown  in 
Fig.  XXIII-2,  except  that  its  atoms  are  all  alike,  instead  of  their  being 
two  sorts  of  atoms  as  in  the  caesium  chloride  structure.  In  the  face- 
centered  cubic  structure  each  atom  has  twelve  equidistant  neighbors.  In 
the  hexagonal  close-packed  structure  each  atom  also  has  twelve  equi- 
distant neighbors,  but  if  the  structure  is  distorted  so  that  it  is  no  longer 
close  packed,  the  six  atoms  in  the  basal  plane  will  be  at  a  different  dis- 
tance from  the  six  neighbors  in  adjoining  planes.  In  the  body-centered 


446  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

cubic  structure  each  atom  has  eight  equidistant  neighbors.  In  addition 
to  these  structures,  several  elements  crystallize  in  the  diamond  structure, 
with  four  equidistant  neighbors.  The  other  structures,  in  which  the 
unusual  metals  crystallize,  will  be  described  later. 

We  now  give,  in  Table  XXVII-1,  the  crystal  structures  and  lattice 
spacings  of  the  metals.  For  each  metal  we  give  the  structure,  abbreviat- 
ing face-centered  cubic,  hexagonal,  body-centered  cubic,  and  diamond 
structures  by  f.c.,  hex.,  b.c.,  di.,  respectively.  Those  not  crystallizing 
in  these  structures  are  indicated  by  an  asterisk  and  will  be  taken  up  later. 
In  addition,  we  give  the  distance  to  the  nearest  neighbor  (or  to  the  two 
sets  of  nearest  neighbors,  in  the  case  of  hexagonal  structures),  in  ang- 
stroms. It  will  be  observed  that  some  of  the  metals  crystallize  in  more 
than  one  form;  this  is  the  phenomenon  of  polymorphism,  mentioned  in 
Chap.  XI,  Sec.  7.  The  rare  earths  are  omitted  from  our  list,  since 
practically  none  of  them  can  be  obtained  in  the  metallic  form. 

In  examining  Table  XXVII-1,  the  first  thing  to  notice  is  the  regu- 
larity of  the  interatomic  distances  from  metal  to  metal.  Going  through  a 
row  of  the  table,  as  from  potassium  to  selenium,  we  observe  that  the 
largest  distance  is  for  the  alkali,  the  next  largest  for  the  alkaline  earth, 
with  continuing  decrease  for  one  or  two  more  elements.  Then  there  is 
surprising  constancy  for  the  rest  of  the  elements  of  the  series.  We  have 
already  mentioned  that  the  alkalies,  having  the  fewest  electrons  for 
sharing,  are  the  most  loosely  bound  metals,  and  the  alkaline  earths  the 
next  most  loosely  bound;  the  interatomic  distances  bear  this  out,  since 
large  interatomic  distance  corresponds  to  loose  binding.  Of  course,  in 
addition  to  this  trend,  there  is  the  natural  increase  in  size  as  we  go  to 
heavier  atoms  of  the  same  type,  as  from  lithium  to  caesium.  This 
tendency  is  much  less  marked  for  the  more  tightly  bound  metals,  however. 
For  instance,  cobalt,  rhenium,  and  iridium,  similar  elements  in  the  three 
long  periods,  have  almost  exactly  the  same  interatomic  distances. 

It  is  interesting  to  observe  that  this  constancy  of  interatomic  distances 
seems  to  hold  irrespective  of  the  particular  crystal  structure  which  the 
element  may  have.  Thus  cobalt  has  exactly  the  same  interatomic  dis- 
tance in  its  hexagonal  and  face-centered  modifications,  lanthanum  and 
thallium  have  almost  exactly  the  same,  and  iron  changes  only  slightly 
from  one  structure  to  another.  As  we  look  over  the  table,  while  the 
structures  appear  to  be  arranged  almost  at  random,  there  seems  to  be  no 
correlation  between  spacing  and  crystal  structure.  For  instance,  in  the 
series  chromium,  manganese,  iron,  cobalt,  we  have  a  body-centered  cubic 
structure,  a  complex  one  (manganese),  a  face-centered  and  a  body- 
centered  cubic  form  of  iron,  and  a  hexagonal  and  a  face-centered  form  of 
cobalt,  and  yet  the  spacings,  2.49,  2.50,  2.48,  and  2.57,  2.71,  show  only  a 
general  trend,  rather  than  erratic  variation.  This  fits  in  with  the  idea 


SBC.  1]  METALS  447 

TABLE  XXVII-1. — CRYSTAL  STRUCTURE  AND  INTERATOMIC  DISTANCES  IN  METALS 

(Angstroms) 
Abbreviations:  b.c.  body-centered  cubic;  f.c.  face-centered  cubic;  hex.  hexagonal; 

di.  diamond;*  other  structures 

Li  b.c.  Na  b.c.          K  b.c.  Rb  b.c.         Cs  b.c. 

3.03  3.72  4.50  4.86  5.25 


Be  hex.         Mg  hex.         Ca  f.c. 
2.28              3  20               3  93 
2  24              3  19 

Sr  f.c. 
4.29 

Ba  b.c. 
4.35 

B                  Al  f.c.            Sc 
2.85 

Y 

3.58 

La  hex.,  f.c. 
3.72,3  73 

Ti  hex. 
2.95 
2.90 

Zr  hex. 
3.23 

3.18 

Hf  hex. 
3  32 
3  33 

Vb.c. 
2  63 

Nb 

Ta  b.c. 
2.88 

Cr  b.c. 
2  49 

Mo  b.c. 
2.72 

Wb.c. 
2  73 

Mn  * 
2  50 

Fe  f.c.,  b.c. 
2  57,  2  48 

Ru  hex. 
2  69 
2  65 

Os  hex. 
2.71 
2  67 

Co  hex.,  f  <•. 
2  71 

Rh  f.c. 
2  69 

Ir  f.c. 
2  70 

Ni  f.c. 
2.49 

Pd  f.c. 
2  74 

Pt  f.c. 
2.76 

Cu  f.c. 
2.55 

Ag  f.c. 
2  88 

Au  f.c. 
2.87 

Zn  hex. 
2.65 
2.94 

Cd  hex. 
2.97 
3  30 

Hg* 
2  99 

Ga  * 
2  56 

In  * 
3.24, 
3.33 

Tl  hex.,  f.c. 
3  45,  3  43 

Si  di.             Ge  di. 
2.35              2.43 

Sndi. 
2.80 

Pb  f.c. 
3.49 

As  * 
2.50 

Sb  * 
2.88 

Bi  * 
3.10 

Se  * 
2.32 

Te  * 
2.88 

448  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

that  the  metallic  atoms  act  very  much  like  spheres,  as  far  as  their  packing 
is  concerned,  so  that  the  interatomic  distances  of  Table  XXVII-1  measure 
the  diameter  of  these  spheres,  irrespective  of  the  particular  way  they 
happen  to  be  packed  together.  There  is  another  conclusion  to  be  drawn 
from  the  fact  that  the  lattice  spacing  seems  to  be  independent  of  crystal 
structure.  Surely  the  energy  of  the  crystal  will  depend  principally  on  the 
lattice  spacing.  Then  we  must  expect  that  the  same  metal  can  exist  in 
different  structures  with  almost  exactly  the  same  energy.  This  moans 
that  it  should  be  possible  to  change  from  one  structure  to  another  very 
easily,  and  that  very  slight  and  apparently  trivial  circumstances  can 
determine  which  structure  really  has  the  lowest  energy  and  is  stable. 
This  makes  the  polymorphism  of  the  metals  seem  reasonable,  and  it 
makes  it  comprehensible  that  there  should  be  such  apparent  lack  of  order 
in  the  crystal  structure  of  the  elements.  Looking  at  Table  XXVII-1,  it- 
is  very  difficult  to  draw  any  general  conclusions  as  to  why  the  various 
elements  should  have  the  structures  they  do.  There  seems  to  be  an 
almost  random  distribution  of  the  various  structures  among  the  elements. 
But  this  is  not  unnatural  if  the  structure  depends,  not  on  very  fundamen- 
tal properties  like  valence  but  on  relatively  minor  details  of  the  atomic 
structure. 

We  must  still  say  something  about  the  structures  indicated  with  a 
star  in  Table  XXVII-1.  Manganese  is  a  very  peculiar  and  anomalous 
exception  to  the  general  order  of  the  elements.  It  is  the  only  definite 
metal,  far  from  the  nonmetals  in  the  table,  which  has  a  complicated 
structure.  The  structure  is  really  complicated :  there  are  58  atoms  in  the 
unit  cell.  No  one  has  suggested  a  very  convincing  reason  why  this 
should  be  so.  It  becomes  a  little  less  remarkable,  however,  when  one 
studies  the  crystal  structure  of  alloys.  It  is  found  that  some  alloys  of 
quite  ordinary  metals  have  very  complicated  structures,  as  for  instance 
one  of  the  forms  of  the  copper-zinc  alloy  found  in  brass.  These  com- 
plicated structures  come  from  the  mixture  of  two  different  kinds  of 
atoms.  It  has  been  suggested  that  possibly  manganese  atoms  exist  in  two 
different  forms  in  its  crystal,  perhaps  corresponding  in  some  indirect 
way  to  the  two  valences  of  manganese,  and  that  its  structure  is  really 
more  like  that  of  an  alloy  than  of  a  pure  metal.  However  that  may 
be,  it  remains  a  peculiar  and  unexplained  exception. 

The  other  complicated  structures  come  at  the  ends  of  the  groups  in 
the  periodic  table,  and  as  we  have  said  they  correspond  to  something 
more  like  homopolar  bonds  than  metallic  bonds.  We  have  already 
commented  on  germanium  and  tin  (the  so-called  gray  modification  of 
tin),  which  crystallize  in  the  diamond  structure,  corresponding  to  the 
four  homopolar  bonds  which  they  could  form.  They  are  of  course  very 
different  from  diamond  in  their  properties,  though  silicon  is  between  a 


SBC.  1J  METALS  449 

metal  and  a  nonmetal  in  its  behavior,  forming  a  sort  of  bridge  between 
diamond  and  the  others.  Thus  the  melting  points  are  3500°C.  (dia- 
mond), 1420°C.  (silicon),  959°C.  (germanium),  232°C.  (tin),  indicating  a 
rapid  but  not  discontinuous  decrease  of  tightness  of  binding  as  we  go 
from  diamond,  with  pure  homopolar  bonds,  to  germanium  and  tin,  which 
are  much  more  metallic.  In  electrical  properties,  diamond  is  a  very 
good  insulator,  silicon  is  a  substance  of  fairly  high  resistance  but  no 
higher  than  some  alloys,  and  germanium  and  tin  are  fairly  good  conduc- 
tors. In  appearance,  diamond  of  course  is  transparent,  while  silicon, 
germanium,  and  tin  are  all  grayish  materials,  of  fairly  metallic  appear- 
ance. In  mechanical  properties,  silicon  possesses  some  of  the  hardness 
of  diamond  but  germanium  and  tin  are  soft. 


FIG.  XXVII-1. — Layer  of  atoms  from  tho  bismuth  crystal. 

The  next  substances  that  we  shall  consider  are  arsenic,  antimony,  and 
bismuth.  Their  crystals  consist  essentially  of  layers  similar  to  that  shown 
in  Fig.  XXVII-1.  In  such  a  layer,  each  atom  is  bound  to  three  neighbors, 
by  bonds  making  almost  a  right  angle  with  each  other,  but  slightly  opened 
out,  so  as  to  approximate  the  tetrahedral  angles.  If  we  now  stack  such 
layers  on  top  of  each  other,  wo  can  arrange  things  so  that  each  atom  will 
also  have  three  neighbors  in  the  layer  above,  which  are  only  slightly 
further  away  than  the  three  neighbors  in  its  own  layer.  These  six  neigh- 
boring atoms,  three  in  its  own  layer  and  three  in  the  next,  will  surround 
the  atom  very  much  like  the  six  nearest  neighbors  in  a  simple  cubic 
lattice.  In  other  words,  the  bismuth  structure  can  be  regarded  as  a 
slightly  distorted  simple  cubic  lattice,  distorted  in  such  a  way  that  each 
atom  has  three  nearest  neighbors.  As  we  have  already  mentioned,  the 
bonds  to  nearest  neighbors  have  something  of  the  character  of  homopolar 
bonds,  and  each  atom  has  just  enough  electrons  to  form  three  bonds. 
But  the  homopolar  character  is  not  strong,  the  substance  resembling 
a  metal  more  than  a  homopolar  compound.  Still,  the  layer  structure  has 
an  important  effect  on  the  properties  of  the  substances.  Single  crystals 
of  arsenic,  antimony,  and  bismuth  cleave  very  easily,  splitting  between 
layers,  so  that  the  metals  are  very  brittle.  And  while  the  electrical 
conductivity  is  fairly  good  in  directions  parallel  to  the  layers,  it  is  very 


450  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

poor  in  the  direction  normal  to  the  layer,  as  if  the  electrons  had  trouble 
jumping  from  one  layer  to  another. 

Selenium  and  tellurium  have  structures  in  which,  as  we  have  said 
before,  the  atoms  form  chains,  each  atom  being  joined  to  its  two  neighbors 
in  the  chain.  The  chains  are  not  straight,  but  helical,  would  up  like 
springs.  The  electrical  conductivity  is  much  poorer  at  right  angles  to  the 
directions  of  the  chains  than  along  them,  as  we  should  expect.  There  is 
one  remarkable  mechanical  property  of  tellurium:  when  it  is  put  under 
hydrostatic  pressure,  although  of  course  the  volume  decreases,  the  length 
along  the  direction  of  the  chains  increases.  It  is  practically  the  only 
substance  known  that  expands  in  any  direction  under  hydrostatic  pres- 
sure. The  interpretation  is  that  the  sidewise  pressure  tends  to  straighten 
out  the  springs,  thus  lengthening  them,  though  they  contract  laterally 
enough  to  decrease  the  volume  as  a  whole. 

2.  Energy  Relations  in  Metals. — In  Chap.  XXIII  we  have  seen  that 
we  can  understand  the  equation  of  state  of  the  ionic  crystals  in  a  good 
deal  of  detail.  We  shall  now  show  that  similar  progress  can  be  made  in 
considering  the  energy  relations  in  metals.  As  in  that  chapter,  we  shall 
begin  by  considering  the  empirical  values  of  compressibility  and  its 
change  with  pressure,  and  shall  see  whether  the  value  of  the  constant  7 
computed  from  them  agrees  with  the  value  found  from  the  thermal  expan- 
sion, as  Grtineisen's  theory  would  indicate  that  it  should.  We  give  the 
necessary  information  in  Table  XXVII-2.  First  we  give  the  quantities 
Pi  and  P2,  of  Chap.  XIII,  determined  from  the  experimental  values  of 
compressibility  and  its  change  with  pressure.  Then  we  give  values  of  7 

2       P 

found  from  Eq.  (4.6),  Chapter  XIV:  7  =  —  ^  +  -™     For  comparison, 

o      /i 

we  give  values  of  7  found  by  Griineisen's  relation  (4.16)  of  Chap.  XIII, 
thermal  expansion  =  7xCV/T0.  The  agreement  is  definitely  poorer  than 
with  ionic  crystals,  though  still  correct  as  to  order  of  magnitude.1 

The  relations  shown  in  Table  XXVII-2  tell  nothing  about  the  theory 
giving  the  interatomic  energy  as  a  function  of  volume.  We  cannot  get 
nearly  so  far  with  this  as  we  could  with  ionic  crystals.  There  the  attrac- 
tive forces  between  ions  were  simple  Coulomb  forces, .  which  we  could 
calculate  exactly,  though  we  had  to  approximate  the  repulsive  forces  by 
an  empirical  formula,  which  we  took  to  be  an  inverse  power  term.  For 
metals,  the  theory  of  the  metallic  bond  is  so  complicated  that  the  forces 

1  Even  the  order  of  magnitude  is  incorrect  for  the  less  compressible  metals,  most 
conspicuously  for  W  and  Pt,  according  to  Bridgman's  original  measurements  of  P\ 
and  Pa.  A  revision  of  Bridgman's  value  of  Fa  for  iron,  however  [see  Bridgman,  Phys. 
Rev.,  67,  235  (1940),  Slater,  Phys.  Rev.,  (1940)],  makes  possible  a  revision  of  the  P2's 
for  other  incompressible  metals,  resulting  in  the  figures  tabulated  in  Table  XXVII-2, 
and  bringing  about  agreement  with  Gruneisen's  value  of  -y  in  almost  all  cases. 


SBC.  2] 


METALS 


451 


have  been  calculated  for  only  a  few  of  the  simplest  metals,  the  alkalies, 
and  even  there  the  results  are  available  only  in  the  form  of  numerical 
calculations,  though  there  are  analytic  approximations  that  work  fairly 
well  for  sodium,  which  turns  out  to  be  the  simplest  of  all  metals  theoret- 
ically. We  have  pointed  out  in  Chap.  XXII,  however,  that  the  bonds 
between  atoms  in  a  metal  are  not  entirely  different  in  their  properties 
from  homopolar  bonds.  Thus  it  is  not  unreasonable  to  approximate  the 
TABLE  XXVII-2. — QUANTITIES  CONCERNED  IN  EQUATION  OF  STATE  OF  METALS 


P, 

_2      P* 

7 

3      /*, 

(Gruneisen) 

Li  

0.115  X  10ia 

0.149  X  101* 

0  63 

1  17 

Be          .   .            

1.17 

6.2 

4  63 

Na  

0.064 

0.160 

1.83 

1.25 

MK 

0  338 

0  77 

1  62 

Al 

0  757 

1  46 

1  27 

2  17 

K... 

0  028 

0.090 

2.55 

1.34 

Ca.     . 

0  168 

0.23 

0.71 

V  

1  04 

5.6 

2.7 

Cr.  .  .  . 

1  93 

6.4 

2.6 

Mn..                       . 

1  24 

8.1 

5.8 

2.42 

Fe..   . 

1  73 

4.0 

1.7 

1.60 

Co    . 

1.85 

5.1 

2.1 

1.87 

NL... 

1.90 

5  4 

2  2 

1.88 

Cu... 

1.39 

3  6 

1  9 

1  96 

Rb  . 

1  48 

Sr. 

0  122 

0.13 

0  40 

Zr. 

0  91 

5  6 

5  5 

Mo        

2.88 

1.57 

Pd               .           

1.93 

5  9 

2  4 

2.23 

Ag  

1.01 

3  2 

2  5 

2.40 

Cd.. 

0.70 

2  6 

3  04 

Cs..               

1.29 

Ba  

0.098 

0  123 

0.60 

La  

0.284 

0  34 

0.53 

Hf  

1.11 

3  2 

2.22 

Ta  

2.09 

1.75 

w  

3.40 

8 

1  7 

1.62 

Pt  

2.78 

11. 

3  3 

2.54 

Au.         .       . 

1.73 

3  03 

Pb  *     . 

0.42 

1  30 

2  42 

2.73 

Measurements  of  compressibility,  from  which  Pi  and  Pi  are  computed,  are  from  Bridgman,  Proc. 
Am.  Acad.,  58,  165  (1923),  62,  207  (1927),  63,  347  (1928),  64,  51  (1929),  68,  27  (1933),  70,  71  (1935), 
72,  207  (1938).  Values  of  y  by  Grtineisen'e  method  are  taken  from  the  article  by  Grtlneisen,  "  Zustand 
dee  fasten  Kcrpera,"  in  "Handbuch  der  Physik."  Vol.  X,  Springer,  1926. 


452  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

internal  energy  of  a  metal  by  a  Morse  curve,  as  in  Eq.  (1.2)  of  Chap.  IX. 
That  is,  we  write 

t/o  =  I/(e-2o(r-ro)  -  2e-a(r~ro)).  (2.1) 

In  Eq.  (2.1),  C/o  represents  the  energy  of  the  crystal  at  the  absolute  zero 
as  a  function  of  the  distance  r  between  nearest  neighboring  atoms,  where 
ro  represents  the  value  of  this  distance  in  equilibrium,  when  f/o  has  its 
minimum  value.  The  quantity  L  is  the  energy  required  to  break  up  the 
metal  into  atoms  at  the  absolute  zero,  or  the  latent  heat  of  vaporization 
at  the  absolute  zero,  a  quantity  that  can  be  found  from  experimental 
measurements  of  vapor  pressure,  as  we  saw  in  Chap.  XI.  The  quantity 
a  is  an  empirical  constant.  We  shall  determine  L  from  the  experimental 
value  of  the  vapor  pressure,  ro  from  the  experimental  density  at  the 
absolute  zero,  and  a  from  the  compressibility.  To  compare  with  our 
formulation  of  Chap.  XIII,  we  expand  Eq.  (2.1)  in  Taylor's  series,  finding 

[fr   _  A2               /r    _  A3  1 

(ar0)2(~ )    +  (ar0)3f-- )     •  •  •     .       (2.2) 

Then,  comparing  with  Eq.  (3.4)  of  Chap.  XIII,  which  is 

f/o  =  t/oo  +  ATcrgjJ  PI^—-^  ~  9(F?  -  PJ) 
we  have 


+  2].  (2.4) 

'6 

In  terms  of  these  expressions,  we  have 

2       ,      JT2  CWO       ,1  te*    ~\ 

y  == r  J5~  ==   i    o*  \2-O) 

To  use  these  formulas,  we  must  know  the  geometrical  constant  c, 
given  by  the  relation  that  the  volume  per  atom  is  erg.  In  the  body- 
centered  cubic  structure,  we  have  a  cube  with  an  atom  at  each  of  the 
eight  corners  and  one  in  the  center.  Each  of  the  eight  corners  is  shared 
equally  by  eight  cubes,  so  that  each  of  these  atoms  counts  one-eighth 
in  the  cube  in  question.  Thus  the  cube  contains  f  +  1  =  2  atoms.  The 
semi-body  diagonal  is  ro,  so  that  the  body  diagonal  is  2rp,  and  the  side  of 
the  cube  is  2r0/\/3.  Thus  a  cube  of  volume  (2r0/-\/3)3  contains  two 
atoms,  or  a  cube  of  volume  4r8/(3)**  contains  one  atom,  so  that 

c  =  —  for  the  body-centered  cubic  structure.  (2.6) 


SEC.  2]  MKTALS  453 

In  the  face-centered  cubic  structure,  a  cube  contains  eight  atoms  at  the 
eight  corners  (counting  as  f  =  1  atom  within  the  cube),  and  six  at  the 
centers  of  the  six  faces  (counting  as  f  =  3  within  the  cube).  Thus  each 
cube  contains  four  atoms  in  all.  The  semi-face  diagonal  is  r0,  the  face 
diagonal  2r0,  the  side  of  the  cube  \/2ro,  the  volume  2*VJ.  Thus,  since 
this  volume  contains  four  atoms,  the  volume  per  atom  is  (l/\/2)rj!,  and 

c  =  — 7=  for  the  face-centered  cubic  structure.  (2.7) 

V2 

By  similar  methods  we  can  show 

o 

c  =  •$&  for  the  diamond  structure,  (2.8) 

-  for  the  hexagonal  close-packed  structure.  (2.9) 


It  is  natural  that  the  face-centered  and  hexagonal  close-packed  structures, 
both  corresponding  to  the  close  packing  of  spheres,  should  have  the 
same  value  of  c.  In  case  the  hexagonal  lattices  are  not  close  packed,  there 
is  an  additional  correction  factor  in  c,  which  we  shall  not  evaluate. 

We  can  now  use  the  observed  heat  of  evaporation,  the  observed  r*o, 
from  Table  XXVII-1,  and  the  observed  PI,  or  reciprocal  of  the  compressi- 
bility, from  Table  XXVII-2,  to  determine  the  values  of  L  and  of  a.  In 
doing  so,  we  neglect  the  difference  between  PI  and  PJ,  or  between  the 
compressibility  at  room  temperature  and  its  value  at  the  absolute  zero. 
Then,  using  Eqs.  (2.4)  and  (2.5),  we  can  compute  7,  Griineisen's  constant, 
in  terms  of  these  quantities.  In  Table  XXVII-3  we  give  the  necessary 
information.  It  is  seen  from  the  table  that  the  computed  values  of  7 
agree  fairly  well  with  the  observed  values.  This  is  an  indication,  there- 
fore, that  the  Morse  curve  is  not  an  entirely  unsuitable  potential  energy 
function  for  a  metal.  The  agreement  between  computed  and  observed  7 
is,  in  fact,  rather  surprisingly  good  and  indicates  the  degree  of  accuracy 
obtainable  from  a  theoretical  calculation  of  thermal  expansion  from  the 
compressibility  and  specific  heat. 

We  have  already  mentioned,  in  connection  with  the  lattice  spacings, 
that  the  alkali  and  alkaline  earth  metals  are  the  most  loosely  bound,  the 
transition  group  metals  like  the  iron  group  the  most  tightly  bound. 
This  can  be  easily  seen  from  Table  XXVII-3.  A  tightly  bound  crystal 
has  a  large  latent  heat  and  a  low  compressibility,  or  large  value  of  PI.  It 
is  striking  to  see  how  these  quantities  change,  for  instance  from  caesium 
to  tungsten.  The  latent  heat  of  caesium  is  18.7  kg.-cal.  per  mole,  of 
tungsten  203,  almost  the  lowest  and  highest  respectively  for  any  two 
metals.  And  the  value  of  Pi  is  .014  X  1012  dynes  per  square  centimeter 
(or  about  14,000  atm.)  for  caesium,  and  3.40  X  1012  for  tungsten.  The 


454  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

TABLE  XXVII-3. — QUANTITIES  CONCERNED  IN  ENERGY  RELATIONS  OP  METALS 


y  (Eq.  2.5) 


Li 36.0 

Na 26.2 

Mg 36.6 

Al 67.6 

K 21.9 

Ca 42.8 

Cr 89.4 

Fe 96.5 

Co 74.0 

Ni 98.1 

Cu 81.7 

Zn 31.4 

Rb 20.6 

Mo 156. 

Ag 69.4 

Cd 27.0 

Cs 18.7 

W 203. 

Pt 125. 

Au 90.7 


0.80 

0.67 
1.14 
1.21 

0.53 
0.84 
1.64 
1.45 
1.76 
1.49 
1.41 
1.70 

0.47 
1.58 
1.39 
1.93 

0.44 
1.51 
1.68 
1.58 


1.55 

1.58 
2.15 
2.05 


53 

.98 


2.37 
2.20 
2.72 
2.19 
2.13 
2.68 


1.48 
2.48 
2.33 
3.35 

,1.48 
2.39 
2.65 
2.60 


Values  of  latent  heat  of  vaporization  L  are  taken  from  Landolt's  Tables.  They  are  in  kilogram- 
calories  per  gram  mole.  Values  of  a,  in  reciprocal  angstroms,  are  computed  by  Eq.  (2.4),  using  data 
from  Tables  XXVII-1  and  XXVII-2.  Values  of  y,  computed  from  Eq.  (2.5),  are  to  be  compared  with 
values  computed  by  GrUneisen's  method,  tabulated  in  Table  XXVII-2. 

compressibility  of  caesium  is  the  highest  not  only  for  any  metal  but  for 
any  known  solid  at  ordinary  temperatures,  while  that  of  tungsten  is 
among  the  lowest  known.  The  values  of  PI  show  a  continuous  increase 
from  caesium  to  tungsten,  having  the  values  .014  X  1012,  .098,  288, 1.11, 
2.09,  3.40;  it  is  to  be  presumed  that  if  the  latent  heats  of  the  intermediate 
elements  were  known,  they  would  show  a  continuous  increase  in  a  similar 
way.  It  is  interesting  to  note  that  toward  the  ends  of  the  periods  in  the 
table  the  binding  again  becomes  somewhat  less  tight.  Thus  the  com- 
pressibilities increase  and  the  latent  heats  decrease  quite  strikingly,  in 
the  series  Ni-Cu-Zn,  Pd-Ag-Cd,  and  even  more  in  Pt-Au-Hg,  the  latent 
heat  of  mercury,  like  its  melting  point,  being  the  lowest  for  any  metal,  and 
that  of  platinum  being  among  the  highest.  Thus  it  is  quite  clear  that  the 
transition  groups  of  metals  are  much  more  strongly  bound  than  their 
neighbors  either  before  or  after  them  in  the  table. 

While  we  are  not  prepared  to  say  much  about  the  quantitative  details 
of  the  forces  holding  a  metal  together,  still  it  is  not  hard  to  understand  in  a 


SEC.  2] 


METALS 


455 


qualitative  way  the  specially  tight  binding  of  the  transition  elements. 
The  alkali  metals  have  only  one  outer  electron  per  atom.  This  forms  a 
bond  which  must  be  shared  with  all  eight  neighbors  of  an  atom.  The 
alkaline  earths  on  the  other  hand  have  two  outer  electrons  to  be  shared, 
and  the  succeeding  elements  have  three,  four,  etc.  outer  electrons.  Thus 
the  number  of  electrons  available  to  form  metallic  bonds  increases  rapidly 
as  we  go  through  a  period  of  the  table.  It  is  only  natural  that  this 
increases  the  tightness  of  binding.  It  is  a  phenomenon  not  entirely  differ- 
ent from  that  met  in  homopolar  binding,  where  double  and  triple  bonds 
give  considerably  tighter  binding  than  single  bonds;  only  in  the  metallic 
case,  each  bond  is  even  less  than  a  single  bond  in  strength.  As  we  go 
beyond  the  transition  elements,  there  is  a  reversal  of  this  tendency.  The 
electrons  added  in  tho  transition  elements  tend  to  form  completed  shells 
and  to  be  no  longer  available  for  bonding,  so  that  only  a  few  electrons  per 
atom  operate  to  hold  the  crystal  together.  This  tendency  has  not  pro- 
gressed very  far  with  the  elements  copper,  silver,  and  gold.  Though 
these  behave  chemically  to  some  extent  similarly  to  the  alkalies,  they  arc 
obviously  much  more  tightly  bound,  copper  for  instance  having  a  latent 
heat  of  81.7  against  21.9  for  potassium,  and  a  value  of  Pl  of  1.39  X  1012 
against  .028  for  potassium.  Plainly  more  than  one  electron  per  atom  is 
operative  in  the  binding  of  copper,  though  the  next  elements,  Zn,  Cd,  Hg, 
are  more  nearly  comparable  to  Ca,  Sr,  Ba. 

The  values  of  the  constants  a  in  Table  XXVII-3  are  of  the  same  order 
of  magnitude  as  the  values  for  diatomic  molecules,  given  in  Table  IX-1, 
indicating  therefore  that  the  binding  is  not  entirely  different  from  homo- 
polar  binding.  It  is  particularly  interesting  to  compare  these  constants, 
and  in  fact  all  the  properties,  of  the  metals  with  the  corresponding 
properties  of  the  diatomic  molecules  Li2,  Na2,  K2.  These  values  are 
given  in  Table  XXVII-4.  We  give  also  the  values  for  C2  for  comparison 

TABLE  XXVII-4. — COMPARISON  or  CONSTANTS  FOR  MOLECULES  AND  SOLIDS 


TO,  A 

L,  kg.-cal. 

a,A~i 

Li2  

2.67 

26.4 

0.83 

Li  metal  

3.03 

36.0 

0.80 

Na2    

3.07 

17.6 

0  84 

Na  metal  

3.72 

26.2 

0  67 

Ka        

3.91 

11.8 

0  78 

K  metal  

4.50 

21.9 

0.53 

C2  

1.31 

128. 

2.32 

Diamond  

1.54 

199. 

2.17 

456  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

with  the  properties  of  diamond,  though  of  course  it  is  not  a  metal.  While 
there  is  no  exact  parallelism  between  the  molecules  and  the  solids,  still 
there  are  strong  resemblances.  The  interatomic  spacing  in  the  molecule 
is  in  each  case  between  80  and  90  per  cent  of  the  spacing  in  the  crystal. 
This  is  to  be  explained  by  the  fact  that  the  bond  is  concentrated  between 
the  two  atoms  of  the  molecule,  pulling  them  together,  while  in  the  lattice 
it  is  shared  among  the  neighbors.  The  latent  heats  of  evaporation  are 
in  every  case  greater  than  the  heats  of  dissociation  of  the  molecule  but 
less  than  twice  the  heat  of  dissociation.  Finally,  the  values  of  the  con- 
stants a  for  the  metal,  while  they  do  not  agree  exactly  with  those  for  the 
molecules,  are  of  the  same  order  of  magnitude,  indicating  rather  close 
similarity  between  the  binding  in  the  two  cases. 

3.  General  Properties  of  Metals. — In  many  ways  the  most  con- 
spicuous property  of  metals  is  the  electrical  conductivity.  We  have 
mentioned  that  this  results  from  free  electrons,  electrons  not  definitely 
tied  up  in  any  definite  atom  or  homopolar  bond  but  free  to  wander  from 
one  bond  to  another.  We  shall  treat  conductivity  in  detail  later  on.  We 
may  mention  now  only  one  general  fact,  bearing  on  our  picture  of  conduc- 
tion. Surely,  we  should  expect  at  first  sight  that  the  more  electrons  there 
were  to  carry  the  current,  the  bigger  the  conductivity  would  be.  This 
being  so,  we  might  suppose  that  metals  with  two,  three,  or  more  valence 
electrons  per  atom  would  conduct  better  than  those  with  only  one.  The 
opposite  is  the  case,  however.  When  reduced  to  proper  units,  the  con- 
ductivities of  the  alkali  metals,  and  of  copper,  silver,  and  gold,  the 
elements  with  one  electron  per  atom,  are  greater  than  for  any  other 
substances.  The  interpretation  of  this  is  that,  though  the  other  elements 
have  more  electrons,  still  these  electrons  form  bonds  which  are  more  like 
valence  bonds,  are  localized  more  definitely  between  pairs  of  atoms,  and 
consequently  are  less  free  to  travel  around.  The  limiting  case  is  diamond, 
where  there  are  just  enough  electrons  to  form  bonds  and  there  is  no 
conductivity.  The  alkalies,  with  the  fewest  electrons,  have  at  the  same 
time  the  freest  electrons,  for  they  must  continually  circulate  about  to 
produce  the  binding  between  different  pairs  of  atoms. 

Another  characteristic  property  of  metals  is  ductility.  A  metal  can 
be  bent  and  deformed  without  breaking,  much  more  than  most  other  sub- 
stances, as  for  instance  ionic  crystals.  This  is  particularly  striking  with 
the  close-packed  metals.  In  a  metal,  the  bonds  act  quite  indiscriminately 
between  any  closely  neighboring  atoms.  They  do  not  depend  greatly 
on  the  exact  orientation  of  the  atoms,  as  the  real  homopolar  valences  do. 
Thus  a  distortion  of  the  lattice,  so  long  as  it  docs  not  involve  much  net 
change  of  the  interatomic  distances,  will  not  greatly  change  the  energy 
and  will  not  be  opposed  by  a  large  force.  And  even  a  large  distortion 
does  not  weaken  the  lattice  and  may  in  some  cases  even  strengthen 


SEC.  3]  METALS  457 

it.  This  is  shown  in  the  phenomenon  called  cold-working.  A  single 
crystal  of  a  metal  can  often  be  deformed  by  a  process  called  gliding.  As 
indicated  in  Fig.  XXVII-2,  showing  a  projection  of  a  schematic  lattice, 
there  are  planes  through  a  crystal  which  contain  unusually  many  atoms, 
and  so  are  unusually  smooth.  It  is  easy  for  part  of  the  crystal  to  slide 
over  the  rest,  slipping  or  gliding  along  one  r>  O  O  O  O  O 
of  these  planes.  The  possibility  of  gliding,  v  Q  O  O  O  O 
however,  obviously  depends  on  the  per-  o\  O  O  O  O  O 
fection  of  the  crystal.  If  there  are  local  O^v  O  O  O  O 

irregularities  in  the  glide  planes,  these  will      Q       O  \  O       O       O      O 
act  like  roughnesses  on  two  surfaces  of  a  Q       O\O       O      O 

bearing  and  will  increase  the  friction,  pro-      Q       O       O\  O       O      O 


x" 

venting  gliding.     Now  if  a  crystal  glides,  o       O       O    \O       O 

while  the  sliding  over  the  planes  will  be  O       O      O       O\O      O 

fairly  smooth  and  will  not  result  in  much  O       O       O       O  'V  O 

distortion  of  the  lattice,  still  some  atoms  O       O       O       O       O  \  O 

are  bound  to  be  pulled  out  of  place.     These  FIG.    xxvii-2.— Schematic   repre- 

.,i  ,,  ,   ri  i      ...       .  ,  sentation  of  a  glide  plane. 

will  then  act  like  irregularities  in  prevent- 
ing further  gliding,  so  that  the  crystal  will  have  become  hardened  by  its 
distortion.  If  this  process  is  long  continued,  the  metal  can  become  very 
much  harder  than  in  its  crystalline  form.  This  is  very  well  known  in 
metallurgy,  where  metals  are  hardened  by  being  hammered,  drawn,  and 
otherwise  distorted. 

The  most  conspicuous  examples  of  gliding,  and  hardening  by  cold 
work,  are  found  in  those  noncubic  crystals  that  have  only  one  set  of 

planes  over  which  gliding  is  possible. 
The  best-known  case  is  zinc.  This  is  a 
hexagonal  crystal,  but  so  far  from  close 
packed  that  the  separations  between  an 
atom  and  its  neighbors  out  of  the  basal 

/\  Y  A        plane  are  about  ten  per  cent  greater  than 

between  the  atom  and  its  neighbors  in  its 
own  plane.     That  is,  it  is  almost  a  layer 
structure,  the  binding  perpendicular  to 
Unstrefched  ^"*        the   layers   being   considerably   weaker 

S+ retched  than  the  binding  in  the  layers.     Thus 

Fio.    XXVII-3.— Gliding    of    a    zinc    ,,  .    .    f.  ..  ,  .. 

crystai.  the  zinc  crystal  slips  or  glides  very  easily 

parallel  to  the  planes  or  at  right  angles 

to  the  hexagonal  axis.  A  single  unstrained  zinc  crystal  can  be  greatly 
distorted  by  the  application  of  very  small  forces,  by  such  gliding.  For 
instance,  a  crystal  in  the  form  of  a  rod,  with  the  layers  inclined  to  the 
axis  of  the  rod,  as  in  Fig.  XXVII-3,  can  be  stretched  out,  as  shown  in  the 
figure,  by  the  application  of  a  force  that  seems  unbelievably  small  when 


458  INTRODUCTION  TO  CHEMICAL  PHYSICS     [CHAP.  XXVII 

judged  by  our  ordinary  ideas  of  the  strength  of  metals.  The  stretching 
is  plainly  produced  by  gliding,  for  in  the  actual  stretched  material  one 
can  see  the  surfaces  along  which  gliding  has  occurred,  in  a  way  shown  in  an 
exaggerated  form  in  the  figure.  But  once  the  gliding  has  occurred,  the 
rod  is  strengthened  so  much  that  it  cannot  be  pushed  back  into  its  original 
form  by  any  small  force.  The  gliding  and  hardening  are  easy  to  see  and 
understand  in  this  case,  because  there  is  only  one  set  of  slip  planes.  In  a 
cubic  crystal,  with  many  sets  of  planes  on  account  of  symmetry,  some 
parts  of  the  crystal  will  glide  along  one  plane,  other  parts  along  others, 
and  the  phenomenon  is  much  harder  to  visualize  and  interpret.  But  the 
essential  feature  is  the  same,  that  the  undistorted  crystal  glides  easily 
but  is  distorted  enough  by  gliding  so  that  it  is  very  much  hardened.  The 
ordinary  materials  that  we  meet  in  practice  are  made  up  of  many  small 
crystal  grains  and  are  ordinarily  much  distorted  in  the  process  of  manu- 
facture. Thus  they  are  very  hard  compared  to  an  undistorted  single 
crystal,  which  has  properties  that  at  first  sight  make  it  appear  very 
peculiar  and  unfamiliar. 

Still  another  characteristic  property  of  metals  is  their  ability  to  form 
alloys  or  compounds  of  different  metals  with  variable  composition.  We 
have  already  mentioned  alloys  to  some  extent  in  Chap.  XVII,  when  we 
were  talking  about  equilibrium  between  phases.  There  we  saw  that  some 
alloys  consist  of  a  single  phase,  while  others  are  mixtures  of  two  phases, 
each  phase  having  characteristic  crystal  structure  and  other  properties. 
Generally  in  these  cases  each  phase  exists  in  small  crystal  grains,  the 
whole  alloy  consisting  of  a  mixture  of  these  two  kinds  of  crystals,  in  close 
juxtaposition  to  each  other.  We  now  consider  the  nature  of  one  of  the 
pure  phases,  which  itself  generally  contains  two  or  more  elements  in 
variable  proportions.  These  phases  are  of  two  sorts,  substitutional  and 
interstitial  compounds.  In  a  substitutional  alloy,  atoms  of  one  type  in 
the  lattice  are  simply  absent  and  are  replaced  by  atoms  of  the  other 
type,  so  that  the  final  lattice  is  not  much  affected  by  the  substitution. 
Such  alloys  occur  when  the  two  (or  more)  types  of  atom  in  question  are 
about  the  same  size  and  form  similar  crystals.  Thus,  from  Table 
XXVII-1,  we  see  that  the  interatomic  distance  in  nickel  is  2.49  A  and  in 
copper  2.55  A,  quite  similar  distances,  indicating  that  the  atoms  are 
about  the  same  size.  Furthermore,  nickel  and  copper  both  form  face- 
centered  crystals.  It  is  then  not  surprising  that  any  fraction,  from  zero 
to  100  per  cent,  of  the  atoms  in  a  nickel  lattice  can  be  replaced  by  copper 
atoms.  There  is,  in  other  words,  a  complete  series  of  substitutional  alloys 
of  copper  and  nickel  for  any  composition.  All  these  alloys  have  the  same 
face-centered  structure,  with  a  lattice  spacing  that  varies  smoothly  from 
one  limit  to  the  other.  The  other  extreme  is  the  interstitial  alloy.  This 
is  found  where  an  atom  much  smaller  than  the  other  atoms  of  the  crystal 


SBC.  3)  METALS  459 

enters  into  combination.  In  this  case,  the  small  atom  does  not  substitute 
for  another  atom  but  goes  into  a  hole  between  other  atoms,  filling  one  of 
the  interstices  between  atoms,  whence  the  name.  A  characteristic  exam- 
ple is  the  alloying  of  carbon  with  iron  to  form  steel.  The  interatomic 
distance  in  diamond  is  1.54  A  and  in  iron  2.57  A,  showing  that  the  carbon 
atom  is  much  smaller  than  that  of  iron.  The  carbon  atoms  fit  between 
iron  atoms,  distorting  the  lattice  a  good  deal.  This  has  one  obvious 
effect:  by  distorting  the  lattice,  the  interstitial  atoms  prevent  gliding 
and  harden  the  metal.  It  is  for  this  reason  that  steel  is  so  much  harder 
than  iron.  The  alloying  atoms  in  an  interstitial  compound  interfere 
with  the  lattice  much  more  than  in  a  substitutional  compound,  with  the 
result  that  far  fewer  can  be  introduced  into  the  lattice.  Thus  only  a  few 
per  cent  of  carbon  can  be  introduced  into  iron,  in  contrast  to  the  case  of 
nickel  and  copper  where  any  amount  of  copper  can  be  introduced  into 
nickel. 


CHAPTER  XXVIII 
THERMIONIC  EMISSION  AND  THE  VOLTA  EFFECT 

In  the  preceding  chapter  we  took  up  the  properties  of  metals,  but  we 
have  said  very  little  about  their  most  characteristic  feature,  their  electrical 
behavior.  An  understanding  of  the  electrical  conductivity  of  metals,  and 
its  bearing  on  the  free  electrons  in  the  metal,  is  essential  to  a  proper  treat- 
ment of  the  metallic  bond  and  of  the  forces  holding  the  metal  together. 
We  shall  take  these  questions  up  in  the  next  chapter.  Before  doing 
so,  however,  we  shall  take  up  a  related  problem,  the  thermionic  emission 
of  electrons  from  hot  metals.  By  studying  the  interaction  between 
electrons  and  metals  we  can  get  some  information  about  electrons  inside 
metals,  more  or  less  in  the  same  way  that  by  studying  the  interaction 
between  electrons  and  atoms,  in  such  problems  as  resonance  and  ioniza- 
tion  potentials,  we  can  get  information  about  atomic  structure.  The 
information  is  not  so  detailed  as  with  atoms,  but  nevertheless  both  the 
practical  importance  of  the  problem  itself  and  its  bearing  on  the  structure 
of  metals  furnish  justification  for  studying  it  at  this  point. 

In  Chap.  XX,  Sec.  3,  we  spoke  about  the  detachment  of  electrons 
from  atoms,  and  in  Sec.  4  of  that  chapter  we  took  up  the  resulting  chem- 
ical equilibrium,  similar  to  chemical  equilibrium  in  gases.  But  electrons 
can  be  detached  not  only  from  atoms  but  from  matter  in  bulk,  and  par- 
ticularly from  metals.  If  the  detachment  is  produced  by  heat,  we  have 
thermionic  emission,  a  process  very  similar  to  the  vaporization  of  a  solid 
to  form  a  gas.  The  equilibrium  concerned  is  very  similar  to  the  equilib- 
rium in  problems  of  vapor  pressure,  and  the  equilibrium  relations  can  be 
used,  along  with  a  direct  calculation  of  the  rate  of  condensation,  to  find 
the  rate  of  thermionic  emission.  In  connection  with  the  equilibrium  of  a 
metal  and  its  electron  gas,  we  can  find  relations  between  the  electrical 
potentials  near  two  metals  in  an  electron  gas  and  derive  information  about 
the  so-called  Volta  difference  of  potential,  or  contact  potential  difference, 
between  the  metals.  We  begin  by  a  kinetic  discussion  of  the  collisions  of 
electrons  with  metallic  surfaces. 

1.  The  Collisions  of  Electrons  and  Metals. — When  a  slow  electron, 
with  a  few  volts'  energy,  strikes  a  clean  metallic  surface,  there  is  a  very 
large  probability  that  it  will  be  captured  by  the  metal,  and  a  very  small 
probability,  depending  on  its  energy  and  direction,  that  it  will  be  reflected. 
In  other  words,  the  most  likely  collisions  are  inelastic  ones,  in  which  the 
electron  loses  all  its  energy  and  never  gets  out  again.  The  mechanism  is 

460 


SBC.  1]       THERMIONIC  EMISSION  AND  THE  VOLTA  EFFECT  461 

simple.  The  electron  can  penetrate  a  few  atoms  deep,  but  it  is  likely  to 
have  a  collision  without  electronic  excitation  with  each  atom  it  strikes, 
losing  energy  to  produce  thermal  vibration  of  the  atoms.  A  few  such 
collisions  reduce  its  energy  far  enough  so  that  it  does  not  escape  from 
the  metal.  For  it  requires  considerable  kinetic  energy  for  an  electron  to 
leave  a  metal.  It  is  attracted  to  the  metal  by  the  so-called  image  force, 
which  we  shall  discuss  later,  and  it  requires  several  volts'  energy  to  escape. 
This  can  be  indicated  graphically  as  in  Fig.  XXVIII-1.  Here  we  show 
schematically  the  potential  energy  of  an  electron,  inside  and  outside  a 
metal.  If  an  electron  with  one  volt  kinetic  energy  outside  the  metal 
enters  it,  it  will  have  a  kinetic  energy  of  a  number  of  volts  inside  the 
metal,  since  its  total  energy,  kinetic  plus  potential,  must  remain  constant, 
and  the  potential  energy  is  lower  inside  the  metal.  If  the  electron's 
energy  is  lowered  by  inelastic  collision  from  its  original  value  E  to  a  value 


E1 


— Outside  metah- — •* > 

Fio.  XXVIII-1. — Potential  energy  of  an  electron  at  a  metallic  surface. 

E'  below  the  potential  at  infinity,  it  will  be  unable  to  escape.  This  is 
what  usually  happens.  The  exceptional  case,  reflection,  comes  when  the 
direction  of  the  electron  happens  to  be  reversed  at  its  first,  or  practically 
its  first,  collision  with  an  atom  within  the  metal,  so  that  it  comes  out  again 
without  chances  of  further  collisions.  In  such  a  case  it  will  have  lost  only 
a  small  amount  of  energy,  and  the  collision  will  be  almost  elastic. 

When  faster  electrons  strike  a  metal,  the  situation  is  more  complicated. 
In  the  first  place,  secondary  electrons  can  be  emitted,  electrons  liberated 
from  the  metal  by  the  impact  of  the  primary  electron.  In  some  cases  it 
is  not  possible  to  know  whether  the  electron  coming  out  of  the  metal  is  a 
secondary  or  the  primary,  except  in  cases  where  both  come  out,  so  that 
more  electrons  leave  the  metal  than  enter  it,  an  important  case  prac- 
tically. But  generally  from  the  velocities  it  is  possible  to  draw  conclu- 
sions as  to  whether  the  emitted  electron  is  primary  or  secondary.  The 
secondary  electrons  mostly  have  fairly  low  energies,  in  the  neighborhood 
of  ten  or  twenty  volts;  of  course,  electrons  of  such  energy  cannot  be 
emitted  unless  the  primary  electrons  had  a  suitably  larger  amount  of 
energy.  Among  the  few  primary  electrons  that  are  reflected,  some  are 
elastically  reflected,  as  with  slow  electrons,  and  have  lost  almost  no 
energy.  But  some  have  lost  just  about  as  much  energy  as  the  secondaries 


462  INTRODUCTION  TO  CHEMICAL  PHYSICS    [CHAP.  XXVIII 

acquire,  and  this  is  easily  interpreted.  It  is  assumed  in  such  cases  that 
the  primaries  have  ionized  atoms  within  the  metal,  losing  energy  in  the 
process,  and  have  then  happened  to  be  reversed  in  direction  and  to  escape 
from  the  metal.  The  secondaries  are  supposed  to  be  the  electrons  ionized 
from  the  atoms.  Since  the  probability  of  ionization,  or  the  collision  cross 
section  for  ionization,  has  its  maximum  when  the  energy  of  the  impinging 
electron  is  something  like  twice  the  ionization  potential,  it  is  reasonable 
that  these  secondaries  should  have  energies  of  a  number  of  volts. 

In  addition  to  these  reflection  phenomena,  which  are  observed  with 
all  types  of  surfaces,  there  are  some  very  special  effects  observed  when 
the  surface  is  a  crystal  face  of  a  single  crystal  of  metal.  Electrons  of 
certain  definite  energies  and  angles  of  incidence  have  an  abnormally 
large  reflection  coefficient  for  elastic  reflection,  sometimes  several  times  as 
great  as  that  found  with  neighboring  angles  and  energies.  This  phe- 
nomenon is  called  electron  diffraction  and  has  been  of  great  importance  in 
developing  wave  mechanics,  for  in  it  the  beam  of  electrons  acts  like  a 
wave,  and  the  directions  of  abnormal  reflection  are  determined  by  inter- 
ference conditions,  exactly  analogous  to  the  interference  conditions  in 
x-ray  diffraction.  While  electron  diffraction  is  of  great  importance  in 
theory  and  is  useful  in  determining  the  crystal  structure  of  surface  layers, 
it  is  not  very  important  for  our  present  purpose. 

Now  let  us  ask  what  are  the  inverse  processes  to  the  collisions  we  have 
considered.  Of  course,  the  inverse  to  an  elastic  collision  is  also  an  elastic 
collision,  so  we  do  not  need  to  consider  this  case  further.  The  inverse 
to  the  inelastic  capture  of  a  slow  electron  by  a  metal  is  obviously  a  process 
by  which  an  electron  in  the  metal  happens  to  receive  a  number  of  collisions 
in  succession  by  atoms  of  the  metal,  in  which  its  energy  increases  instead 
of  decreasing  in  the  normal  way,  thus  giving  it  enough  energy,  as  E  in 
Fig.  XXVIII-1,  so  that  it  can  escape  from  the  metal.  Speaking  in  a  less 
detailed  way,  it  is  a  process  in  which  the  electron  inside  the  metal,  by 
thermal  agitation,  happens  to  get  an  abnormal  energy  and  escapes.  The 
inverse  to  the  emission  o'f  a  secondary  is  a  complex  process  in  which  two 
electrons,  a  fast  and  a  slow  one,  strike  the  metal,  the  slow  one  recombines 
with  an  atom  of  the  metal,  which  gives  up  the  extra  energy  to  the  fast 
electron,  throwing  it  out  of  the  metal.  Of  these  processes,  the  only 
important  one  is  the  ejection  of  a  fast  electron  by  a  metal,  and  this  is 
what  is  known  as  thermionic  emission.  The  higher  the  temperature  the 
more  likely  one  is  to  find  electrons  fast  enough  to  escape,  and  the  greater 
the  thermionic  emission.  Our  task  in  the  next  sections  will  be  to  calculate 
the  rate  of  thermionic  emission,  the  number  of  electrons  per  second 
3mitted  by  a  square  centimeter  of  surface,  for  this  is  a  very  important 
quantity  in  practice.  But  it  is  a  very  difficult  thing  to  find  directly,  as  so 
many  rates  of  reaction  are  difficult  to  compute.  To  find  it  directly,  we 


SBC.  2]        THERMIONIC  EMISSION  AND  THE  VOLT  A  EFFECT  463 

have  to  know  a  great  deal  about  the  structure  of  the  metal.  In  the  next 
chapter  we  shall  investigate  this  and  shall  be  able  to  make  a  direct  calcula- 
tion. But  for  the  present  we  can  proceed  otherwise.  First  we  investigate 
the  equilibrium  of  a  metal  and  a  gas  of  electrons  outside  it.  Then  we 
find  the  rate  of  the  process  by  which  an  electron  strikes  the  surface  and 
sticks,  the  inverse  to  thermionic  emission.  By  combining  those  two 
pieces  of  information,  we  can  derive  the  rate  of  thermionic  emission,  with- 
out knowing  any  details  about  the  metal  at  all. 

2.  The  Equilibrium  of  a  Metal  and  an  Electron  Gas.  —  The  problem  of 
equilibrium  between  a  metal  and  the  electron  gas  surrounding  it  is  very 
much  like  that  of  the  vapor  pressure  of  a  solid,  which  we  have  considered 
in  Chap.  XI,  Sec.  5,  except  that  it  is  not  the  solid  itself  which  is  evaporat- 
ing, but  only  the  electrons  from  it.  Nevertheless,  with  proper  interpreta- 
tion, we  can  use  the  same  formulas  that  were  developed  in  that  section. 
For  equilibrium,  we  must  equate  the  Gibbs  free  energy  first  of  a  piece  of 
metal  and  the  electron  gas  outside  it,  then  of  the  same  piece  of  metal 
lacking  a  certain  number  of  electrons,  plus  the  gas  formed  outside  the 
metal  by  the  original  electrons  plus  those  just  separated  from  the  metal. 
The  resulting  formula  for  the  pressure  is  just  like  Eq.  (6.4),  Chap.  XI: 

T  T 


^-Lo        CTdT  (T        _ 

e  RTe    *  °  RT*J  °  (2.2) 


Here  the  quantity  i  is  the  chemical  constant  of  the  electron  gas,  defined 
in  Eq.  (3.16),  Chap.  VIII.  The  interpretation  of  the  quantities  L0  and 
Cp  must  be  examined  in  detail.  The  quantity  Lo  is  clearly  the  work 
required  to  remove  a  mole  of  electrons  reversibly  from  the  metal  at  the 
absolute  zero,  leaving  the  metal  in  its  lowest  possible  electronic  energy 
level.  Thus  Lo  is  the  latent  heat  of  vaporization  of  electrons  from  the 
metal  at  the  absolute  zero,  or  the  thermionic  work  function.  For  most 
metals  it  is  of  the  order  of  magnitude  of  four  or  five  electron  volts  (or 
around  100  kg.-cal.  per  gram  mole).  We  shall  find  its  interpretation 
in  terms  of  a  model  in  the  next  chapter.  The  quantity  Cp  is  the  difference 
between  the  heat  capacity  of  the  metal  and  its  heat  capacity  when  a  mole 
of  electrons  is  removed.  Thus  it  is  really  more  exact  to  write  the  expres- 
sion No(dCp/dN),  where  dCp/dN  is  the  change  in  heat  capacity  per  unit 
change  in  the  number  of  electrons.  Of  course,  one  cannot  remove  a  mole 
of  electrons  from  a  mole  of  metal,  for  that  would  give  it  an  enormous 
electric  charge,  but  the  quantity  NQ(dCp/dN)  is  really  merely  the  change 
in  specific  heat  per  mole  of  electrons  removed,  and  can  be  found  theo- 
retically by  removing  only  a  few  electrons,  resulting  in  a  negligible  charge. 
Since  changing  the  number  of  electrons  in  the  metal  would  result  in  a 


484  INTRODUCTION  TO  CHEMICAL  PHYSICS    [CHAP.  XXVIII 

corresponding  surface  charge  on  the  surface  of  the  metal,  we  may  regard 
our  quantity  as  the  heat  capacity  of  the  surface  charge,  per  mole.  Now 
this  is  a  very  difficult  quantity  to  find,  either  theoretically  or  experi- 
mentally. Perhaps  it  is  better  to  interpret  it  in  terms  of  Eqs.  (5.6),  (5.7), 
and  (5.8),  Chap.  XI,  which  become 

dL      5p       Ar  aCp 
df  =  2R  "  N'dN' 

L  =  Lo  +  \RT  -  JQ  AT0^  dT.  (2.3) 

The  quantity  L  represents,  here,  the  heat  absorbed  when  a  mole  of  elec- 
trons is  evaporated  at  temperature  T.  That  is,  it  is  the  change  in 
enthalpy,  or  T  times  the  change  of  entropy,  on  evaporation,  or  is  tho 
change  of  internal  energy  plus  PV,  since  we  can  neglect  the  term  PV  for 
the  electrons  in  the  metal.  Thus  it  is  the  latent  heat  of  evaporation 
of  electrons  at  temperature  T.  Then  we  can  interpret  the  quantity 
No(dCp/dN)  merely  in  terms  of  the  change  of  latent  heat  with  tempera- 
ture, a  change  that  can  actually  depend  on  many  factors.  We  may 

fT 

expect  in  any  case  that  the  term  I    Nv(dCP/dN)dT  will  not  be  of  an  order 

JQ 

of  magnitude  greater  than  RT.  For  a  temperature  of  3000°  abs.,  a  high 
temperature  for  thermionic  experiments,  it  will  then  be  of  the  order  of 
magnitude  of  6  kg.-cal.  per  gram  mol,  small  compared  to  I/0,  which  is  of 
the  order  of  100  kg.-cal.  per  gram  mol,  so  that  presumably  the  latent 
heat  does  not  change  by  a  very  large  fraction  of  itself  in  the  usable  range 
of  temperatures. 

Finally  the  quantity  i  in  Eq.  (2.2)  is  the  chemical  constant  of  an 
electron  gas.     This  is  given  by  Eq.  (3.16),  Chap.  VIII: 

(2  4) 


All  the  quantities  in  Eq.  (2.4)  are  familiar  except  gQ,  which  we  now  dis- 
cuss. As  we  saw  in  Chap.  XXI,  Sec.  2,  the  orientation  of  the  electron 
spin  is  quantized  in  space,  so  that  it  has  two  possible  stationary  states 
of  orientation,  in  which  the  spin  is  directed  in  either  of  two  opposing 
directions.  This  results  in  having  00,  the  a  priori  probability  of  the  lowest 
stationary  state,  equal  to  2,  so  that  we  have 


3.  Kinetic  Determination  of  Thermionic  Emission.  —  We  have  now 
found  the  pressure  of  an  electron  gas  in  equilibrium  with  a  metal,  at  an 


SEC.  3]        THEUMION1C  EMISSION  AND  THE  VOLTA  EFFECT  465 

arbitrary  temperature,  by  thermodynamic  methods.  Next  we  shall 
investigate  the  same  problem  by  the  kinetic  method.  We  shall  find  the 
number  of  electrons  per  second  that  enter  the  metal  from  the  electron 
gas,  at  the  equilibrium  pressure.  But  for  equilibrium  this  must  equal 
the  number  of  electrons  per  second  emitted  from  the  metal,  which  is  the 
quantity  we  seek.  To  find  the  number  of  electrons  entering  the  metal  per 
second,  we  first  find  the  number  striking  the  metal  per  second,  which  is  a 
simple  calculation  from  the  kinetic  theory  of  gases.  Then  we  assume  that 
a  fraction  r  of  these  electrons  will  be  reflected,  (1  —  r)  captured.  From 
what  we  have  said,  for  slow  electrons  such  as  we  are  dealing  with,  r  is 
small  compared  to  unity,  so  that  (1  —  r)  is  almost  unity.  Actually  the 
reflection  coefficient  will  presumably  depend  on  the  velocity  of  the  imping- 
ing electron,  but  for  simplicity  we  shall  ignore  this  dependence,  proceeding 
as  if  r  were  a  constant. 

Let  us  first  find  the  number  of  molecules  of  a  perfect  gas,  at  pressure* 
P,  temperature  T,  striking  a  square  centimeter  of  surface  per  second. 
Since  the  electrons  act  like  a  perfect  gas,  this  calculation  will  apply  to 
them  as  well  as  to  an  ordinary  gas.  The  calculation  is  similar  to  that  of 
Sec.  3,  Chap.  IV,  whore  we  found  the  pressure  by  tho  kinetic  method. 
Consider  the  molecules  whose  momentum  lies  in  the  range  dpx  dpv  dpz. 
As  in  Fig.  IV-2,  the  number  of  such  molecules  crossing  one  square  centi- 
meter perpendicular  to  the  x  axis  per  second  will  be  the  number  in  a  prism 
of  base  one  centimeter,  slant  height  along  the  direction  of  tho  velocity 
equal  to  p/m,  or  altitude  equal  to  px/m.  The  volume  of  this  prism  is 
Paj/ra,  and  the  number  of  suoh  molecules  per  unit  volume,  by  Eq.  (2.4), 
Chap.  IV,  is 


dpx  dpy  dp,.  (3.1) 

Thus  the  number  of  molecules  crossing  the  square  centimeter  per  second, 
found  by  integrating  over  all  values  of  pv  and  pg,  but  only  over  positive 
values  of  px,  is 


™        pt*          f  "         p*v  f  °°         Pa* 

^e   2mkT  dpA       e   *™kT  dpy\       e   *mkT  dp, 

o     Wl  ,7  —  00  J  —  oo 


)-H  =  P(2nntT)-Hf    (3.2) 


using  the  formulas  Eq.  (2.3)  of  Chap.  IV.  The  number  (3.2)  gives  the 
number  of  electrons  striking  a  square  centimeter  of  metallic  surface  per 
second,  if  the  pressure  is  given  by  Eq.  (2.2).  Since  the  fraction  (1  —  r) 
of  these  enter  the  metal,  we  have  as  the  number  of  electrons  entering  the 
metal  per  second 


466  INTRODUCTION  TO  CHEMICAL  PHYSICS    [CHAP.  XXVIII 


- 

For  equilibrium,  the  number  of  electrons  leaving  the  metal  per  second 
must  equal  this  value.     Thus  the  electron  current  is  this  multiplied  by  tho 

-V    -  CTdT  tTNJ*!*  dT 

electronic  charge  e,  or  is  A'T*e   Te  J  °  RT**  °   °~^      ,  where 

A!  =  (1  —  r)  —  p  —  =  (1  —  r)  120  amp.  per  sq.  cm.  per  degree,2 

V  =  ^  (3.4) 

_£ 

If  No(dCP/dN')  were  zero,  F"  (3.4)  would  have  the  form  A'T*e  T. 
This  is  the  familiar  formula  for  thermiunic  emission,  and  is  one  that  shows 
good  agreement  with  experiment,  except  that  the  experimental  value  of 
A'  does  not  agree  with  the  theoretical  value  given  in  Eq.  (3.4).  Actually 
Ns(dCP/dN)  is  undoubtedly  not  zero,  and  the  last  term  of  Eq.  (3.4)  must 
be  retained.  It  varies  slowly  with  temperature,  however,  while  the 

-5! 
factor  e    T  varies  extremely  rapidly.     Thus  we  can  expand  it  in  serins. 

What  we  shall  do  is  to  assume 


a  linear  function  of  temperature,  which  presumably  is  sufficiently  accurate 
for  the  temperature  range  used.     Then  we  have 


TdT 

0 


f 
e  ~  J 

For  the  thermionic  emission,  we  then  have 


where 

A  -  AV,        b  =  V  -  a.  (3.7) 

Formula  (3.7)  is  of  the  familiar  form,  often  called  a  Richardson  equation. 
We  notice  in  it  that  6  does  not  represent  the  latent  heat  at  the  absolute 
zero  but  a  slightly  different  quantity  without  theoretical  significance, 
since  the  expression  (3.5)  is  merely  a  convenient  approximation  for  a 
small  temperature  range.  And  A  is  not  the  value  of  Eq.  (3.4)  but  has 
the  additional  factor  e&.  This  factor  is  of  the  order  of  magnitude  of 


BBC.  4]        THERMIONIC  EMISSION  AND  THE  VOLT  A  EFFECT  467 

unity,  but  from  the  observed  values  of  the  A's  we  gather  that  it  varies 
from  something  like  10~3  to  103,  depending  on  circumstances. 

4.  Contact  Difference  of  Potential. — Suppose  we  have  two  metals, 
a  and  6,  in  the  same  container  at  the  same  temperature.  For  each  one, 
we  can  calculate  the  vapor  pressure  of  the  electron  gas  in  equilibrium  with 
it  by  Eq.  (2.2).  Since  this  equation  depends  on  the  properties  of  the 
metal,  we  shall  get  a  different  answer  in  the  two  cases.  That  is,  an  elec- 
tron gas  in  equilibrium  with  one  metal,  say  one  with  a  low  work  function, 
will  have  too  great  a  pressure  to  be  in  equilibrium  with  the  second  metal 
with  a  larger  work  function.  Let  us  use  a  kinetic  argument  to  see  what 
will  happen  and  what  sort  of  final  equilibrium  we  may  expect.  Suppose 
the  metal  a,  of  low  work  function,  has  established  its  equilibrium  pressure 
in  the  electron  gas,  and  then  the  metal  b  is  introduced  into  the  container 
and  brought  to  the  same  temperature.  The  gas  pressure  is  too  great  for 
equilibrium  with  6,  so  that  more  electrons  will  strike  its  surface  and 


c*  b 

Fia.  XXVIII-2. — Potential  energy  of  an  electron  between  two  metals,  in  equilibrium. 

condense  than  will  be  emitted.  Thus  there  will  be  a  net  flow  of  electrons 
into  metal  6,  and  it  will  become  negatively  charged.  It  will  then  tend  to 
repel  electrons  near  it,  by  electrostatic  repulsion,  and  this  will  diminish 
the  electron  current  toward  it,  until  finally  the  flow  of  electrons  will 
stop,  the  electrical  difference  of  potential  being  just  great  enough  to  reduce 
the  number  of  electrons  coming  toward  6  to  equality  with  the  number 
leaving  it.  The  net  result  is  that  the  metals  have  become  charged  in  such 
a  way  that  there  is  a  definite  difference  of  potential  between  them.  This 
is  known  as  the  Volta  effeet,  and  the  difference  of  potential  is  the  contact 
difference  of  potential. 

Let  us  now  investigate  the  Volta  effect  more  quantitatively.  When 
equilibrium  is  established,  there  will  be  a  difference  of  potential  between 
the  empty  space  outside  the  two  metals.  The  potential  energy  of  an 
electron  in  this  space  will  then  be  as  in  Fig.  XXVIII-2.  The  jump  in 
potential  at  the  surface  of  each  metal  is  as  in  Fig.  XXVIII-1,  but  now  the 
potential  varies  from  one  metal  to  another.  Since  metal  6  is  negatively 
charged,  the  potential  in  its  neighborhood  is  less  than  near  metal  a,  and 
the  potential  energy  of  an  electron,  which  is  —  e  times  the  electrostatic 
potential,  will  be  greater.  The  difference  of  potential  between  empty 


468  INTRODUCTION  TO  CHEMICAL  PHYSICS    [CHAP.  XXVIII 

space  outside  the  two  metals  is  the  contact  difference  of  potential  and  is 
shown  in  Fig.  XXVIII-2. 

The  electrons  in  the  empty  space  now  form  a  perfect  gas  in  an  external 
force  field.  This  problem  has  been  discussed  in  Section  4,  Chap.  IV. 
There  we  found  that  in  such  a  case  the  temperature  of  the  gas  is  constant 
throughout,  but  the  pressure  and  the  density  vary  from  point  to  point,  the 

_a- 

number  of  molecules  per  unit  volume  being  proportional  to  e  kT  .  In 
such  a  case,  where  the  pressure  is  not  constant  throughout,  we  cannot  use 
the  thermodynamic  method  of  the  Gibbs  free  energy,  for  that  is  based 
on  a  single  pressure,  constant  throughout  the  system,  which  can  be  used 
as  a  thermodynamic  variable.  Fortunately,  however,  we  can  get  as  far 

Epot 

as  we  need  to  here  by  means  of  the  Boltzmann  factor  e  kT  for  the  perfect 
gas.  It  is  just  as  well  to  remember  how  this  factor  comes  about  and  to 
see  that  it  is  closely  related  to  our  explanation  of  the  Volta  effect.  On 
account  of  the  difference  in  potential  between  the  neighborhood  of  a  and  6, 
an  electron  with  energy  shown  in  E\  in  Fig.  XXVIII-2  will  be  slowed 
down  and  stopped  as  it  emerges  from  a  and  tries  to  reach  6.  Instead,  it 
will  be  turned  around  and  will  fall  back  to  a  again.  Only  electrons  with 
energies  like  Eo  will  be  able  to  enter  the  metal  6.  It  is  this  stopping  of 
the  slower  particles  by  the  regions  of  high  potential  energy  which  keeps 
the  density  smaller  there,  and  which  in  this  case  keeps  too  many  electrons 
from  striking  metal  6. 

It  is  now  clear  what  conditions  we  must  have  for  equilibrium  between 
metal  a,  metal  6,  and  the  gas.  The  vapor  pressure  outside  metal  a  must 
be  the  correct  one  for  thermal  equilibrium  with  that  metal,  the  pressure 
outside  6  must  be  the  correct  one  for  equilibrium  with  it,  and  the  pressures 
outside  the  two  metals  must  be  related  according  to  the  Boltzmann 
factor,  to  ensure  equilibrium  between  the  different  parts  of  the  gas.  Thus 
let  the  potential  energy  of  a  mole  of  electrons  outside  the  metal  a  bo 
Ea,  and  outside  metal  6  be  Eb.  Then  the  Boltzmann  factor  leads  to  the 
relation 


P  .(*»-*•). 

IT   =  «    "*T     ,  (*-l) 

•*  o 


where  Pa,  Pb  are  the  pressures  outside  metals  a  and  &.     That  is, 

-a).  (4.2) 


Equation  (4.2)  is  the  condition  of  equilibrium  of  the  gas.  At  the  same 
time,  we  must  have  the  gas  in  equilibrium  with  each  metal,  which  means 
that  Pa  and  Pb  must  be  given  by  Eq.  (2.1).  Thus  we  have 


SEC.  4]        THERMIONIC  EMISSION  AND  THE  VOLT  A  EFFECT  469 


or 


The  significance  of  Eq.  (4.3)  is  clearer  if  we  neglect  the  terms  in 


which  are  small  though  not  entirely  negligible.     Then  it  is 

Ea   —   Eh    ==   La   —  Lb* 

Neglecting  the  No(dCp/dN)'*,  the  L's  are  the  latent  heats.  Thus  we 
have  the  important  statement  that  the  contact  difference  of  potential 
between  two  metals  equals  the  difference  of  their  latent  heats,  or  approxi- 
mately of  their  work  functions.  This  relation  is  found  to  be  verified 
experimentally.  The  contact  difference  of  potential  can  be  found  by 
purely  electrostatic  experiments,  and  the  work  functions  by  thermionic 
emission;  the  results  obtained  in  these  two  quite  different  types  of  experi- 
ment are  in  agreement.  The  small  correction  terms  arising  from  the 
No(dCp/dNYs  lie  almost  within  the  errors  of  the  experiments,  so  that  we 
hardly  need  consider  them  in  our  statement  of  the  general  theorem. 
It  is  a  characteristic  of  thermal  equilibrium  that  it  is  the  same,  no 
matter  what  agency  or  process  brings  it  about.  Thus  in  particular,  two 
metals  in  thermal  equilibrium  will  take  up  a  difference  of  potential  equal 
to  the  contact  difference,  so  long  as  there  is  any  agency  whatever  by 
which  charge  can  flow  from  one  to  the  other.  In  the  case  we  considered, 
the  electron  gas  formed  the  agency;  it  is  a  conductor  and  carried  current 
from  one  metal  to  the  other.  But  we  can  readily  imagine  conditions 
where  the  electron  gas  would  not  be  an  effective  agency.  For  instance, 
we  may  have  the  metals  at  room  temperature.  The  density  of  the 
electron  gas  at  room  temperature  is  so  extremely  small  that  for  all  prac- 
tical purposes  it  has  no  electrons  at  all  in  it,  and  it  would  take  an  exces- 
sively long  time  to  transfer  appreciable  charge  or  produce  equilibrium. 
Nevertheless  in  time  the  equilibrium  would  be  established  and  from 
Eq.  (4.3),  since  the  temperature  dependent  terms  are  small,  the  contact 
difference  of  potential  at  room  temperature  must  be  about  the  same  as  at 
high  temperature.  But  there  are  other  much  more  effective  ways  of 


470  INTRODUCTION  TO  CHEMICAL  PHYSICS    [CHAP.  XXVIII 

transferring  charge  from  one  metal  to  another  at  room  temperature :  they 
may  be  connected  by  a  wire  or  other  conductor.  Thermodynamics  now 
requires  that  if  this  is  done,  the  metals  will  automatically  come  to  such 
potentials  that  the  points  directly  outside  the  two  metals  will  differ  by 
the  contact  potential.  This  of  course  demands  that  the  metals  will 
automatically  become  charged  enough  to  produce  these  potentials  in 
outside  space. 

This  mechanism  furnishes  the  basis  of  one  electrostatic  method  of 
measuring  contact  potentials.  The  two  metals  whose  difference  of 
potential  is  desired  are  made  into  plates,  so  that  they  can  be  brought 
close  together  like  the  plates  of  a  condenser.  When  they  are  close 
together  they  are  connected  by  a  wire,  so  that  they  will  set  up  the  contact 
difference  of  potential;  then  they  are  disconnected.  The  capacity  of  a 
condenser  is  inversely  proportional  to  the  distance  between  the  plates,  so 
that  it  is  very  large  when  the  plates  are  close  together,  and  it  requires  a 
large  charge  to  produce  the  required  potential  difference.  Then  the 
plates,  insulated  from  each  other,  are  removed  to  a  long  distance  apart. 
Their  capacity  becomes  much  smaller  and  the  charge,  which  of  course 
remains  the  same,  raises  them  to  a  large  potential  difference,  large  enough 
so  that  it  can  be  readily  measured  with  an  electrostatic  voltmeter. 

We  notice  that  in  equilibrium,  a  metal  with  a  low  work  function 
becomes  positively  charged,  one  with  a  high  work  function  negatively 
charged,  so  as  to  set  up  the  contact  difference  of  potential.  It  is  interest- 
ing to  notice  the  close  similarity  between  this  and  the  corresponding 
situation  of  two  atoms,  one  of  low  ionization  potential  and  the  other  of 
high  ionization  potential,  in  equilibrium,  as  we  considered  them  in  Chap. 
XX,  Sec.  4.  If  either  one  lacks  an  electron,  we  found  that  it  would  be  the 
electropositive  one,  the  one  with  low  ionization  potential.  Similarly 
here  we  may  consider  the  metals  of  low  work  function  to  be  electropositive 
ones,  which  lose  electrons  easily.  We  cannot  push  this  analogy  too  far, 
however,  for  while  there  is  some  parallelism  between  ionization  potential 
and  work  function,  it  is  by  no  means  very  complete,  so  that  we  should  not 
arrange  the  metals  in  the  same  series  of  electropositive  or  negative  charac- 
ter by  means  of  the  work  functions  that  we  should  from  ionization 
potentials. 

There  is  an  interesting  graphical  interpretation  of  our  result  that  the 
contact  difference  of  potential  between  two  metals  equals  the  difference 
of  work  function.  In  Fig.  XXVIII-2,  let  us  go  down  from  the  energy 
Ea  by  the  amount  La,  and  down  from  the  energy  Eb  by  the  amount  L&. 
Then  from  Eq,  (4.3),  the  resulting  energies,  Ea  —  La,  and  Eb  —  L&,  will 
be  approximately  the  same,  so  that  the  horizontal  lines  drawn  at  these 
heights  in  the  two  metals,  in  Fig.  XXVIII-2,  will  be  at  the  same  height 
in  these  two,  or  any  two,  metals  in  equilibrium.  Let  us  see  what  inter- 


SEC.  4]       THERMIONIC  EMISSION  AND  THE  VOLT  A  EFFECT          471 

pretation  we  can  give  to  these  levels.  The  quantity  La  represents  the 
work  done  on  an  electron  (or  a  mole  of  electrons,  to  be  more  precise),  in 
removing  it  from  metal  a  at  the  absolute  zero.  We  might  adopt  a  very 
crude  picture  of  a  metal:  we  might  suppose  it  to  be  a  region  of  constant 
potential  energy  for  electrons,  in  which  the  electrons  simply  formed  a 
perfect  gas  of  high  density.  Then  if  the  potential  energy  of  an  electron 
jumped  by  the  amount  La  in  going  out  of  the  surface,  we  should  under- 
stand the  interpretation  of  the  work  function.  With  this  simple  inter- 
pretation, the  equality  of  Ea  —  La  and  Eb  —  L&  would  mean  that  the  two 
metals  set  themselves  so  that  the  potential  energy  of  an  electron  wavS  the 
same  in  each;  if  they  were  joined  by  a  wire,  there  would  be  no  jump  in 
potential  in  going  from  one  metal  to  another.  And  if  the  electrons 
satisfied  Maxwell-Boltzmann  statistics,  the  density  of  electrons  within 

ii 
either  metal  would  be  greater  by  a  factor  eRT  than  the  density  outside. 

Since  L0  is  much  larger  than  RT  at  ordinary  temperatures,  this  would 
mean  a  very  large  density  of  electrons  within  the  metal  but,  when  one 
calculates  it,  it  is  not  unreasonably  large.  It  would  also  mean  that  the 
densities  of  free  electrons  should  be  the  same  within  any  two  metals. 
This  would  not  be  exactly  the  case,  however,  when  we  recalled  the  addi- 
tional terms  in  Eq.  (4.3),  coming  from  the  No(dCp/dNYs.  These  terms 
can  result  in  slight  differences  of  potential  within  the  metals  and  in 
differences  of  electron  densities. 

The  picture  of  a  metal  which  we  have  just  mentioned  was  elaborated 
greatly  some  years  ago,  and  was  considered  to  be  a  reliable  approximate 
picture.  One  difficulty  remained  with  it,  however.  The  electrons 
within  the  metal  formed  a  perfect  gas,  and  there  was  no  reason  why  they 
should  not  have  the  classical  heat  capacity  of  a  perfect  gas,  f R  per 
mole  for  CV,  independent  of  temperature.  This  would  give  a  contribution 
to  the  specific  heat  of  a  metal,  in  addition  to  that  coming  from  atomic 
vibration,  and  would  increase  the  specific  heat  far  beyond  the  value  of 
Dulong  and  Petit.  Yet  experimentally  metals  agree  fairly  accurately 
with  the  law  of  Dulong  and  Petit,  showing  that  the  electrons  can  con- 
tribute only  a  little,  if  anything,  to  their  specific  heat.  This  difficulty 
showed  that  there  was  something  fundamentally  wrong  with  the  simple 
free  electron  picture  of  a  metal,  and  that  has  proved  to  be  the  assumption 
of  the  Maxwell-Boltzmann  statistics  for  the  electrons.  In  fact,  electrons 
have  been  found  to  satisfy  the  Fermi  statistics.  In  the  next  chapter  we 
consider  the  application  of  this  form  of  statistics  to  a  detailed  model  of 
the  free  electrons  in  a  metal. 


CHAPTER  XXIX 
THE  ELECTRONIC  STRUCTURE  OF  METALS 

The  electrons  in  a  metal,  like  those  in  an  atom,  are  governed  by  the 
quantum  theory,  and  a  complete  study  of  their  motions  is  impossibly 
difficult,  on  account  of  the  enormous  number  of  electrons  in  a  finite  piece 
of  metal,  all  exerting  forces  on  each  other.  The  only  practicable  approxi- 
mation is  similar  to  that  used  in  Chap.  XXI,  Sees.  2  and  3,  where  we  have 
taken  up  the  structure  of  atoms.  There  we  replaced  the  force  acting 
on  an  electron,  which  actually  depends  on  the  positions  of  all  other 
electrons,  by  an  averaged  force,  averaged  over  all  the  positions  which 
the  other  electrons  take  up  during  their  motion.  In  the  case  of  a  metal, 
then,  we  have  a  structure  consisting  of  a  great  many  positive  nuclei, 
arranged  in  a  regular  lattice  structure,  with  electrons  moving  about 
them.  Each  nucleus  will  be  surrounded  by  a  group  of  electrons  forming 
its  inner,  or  x-ray,  shells,  just  as  we  should  find  in  individual  atoms.  The 
remaining  electrons,  however,  the  valence  electrons,  will  move  very 
differently  from  the  way  they  would  in  separated  atoms.  The  reason  is 
that  the  jiimensions  of  th^  orb^s  qf  th<*  valence  electrons  in  the  fltoms 
are  of  the  same  order  of  magnitude  as  the  interat.nTrnn  spring's  fn 
so  that  electrons  of  neighboring  fl.tQ.mfl  w^nH  foTH  fo 


and  profoundly  affect  their  motion.  Thus  we  must  treat  the  problem  as 
if  each  valence  electron  moved  in  the  field  of  the  positive  ions,  consisting  of 
the  inner  shells  and  nuclei  of  all  the  atoms,  and  the  averaged  out  field 
of  all  the  other  valence  electrons.  Our  problem  of  the  electronic  structure 
of  metals,  then,  is  divided  into  two  parts:  first,  we  must  find  what  this 
field  is  like  in  which  the  electrons  move;  secondly,  we  must  investigate 
their  motion  in  the  field.  Then  we  can  build  up  a  model  of  the  whole 
metal,  treating  each  electron  as  if  it  moved  in  the  field  of  which  we  have 
spoken  and  remembering  that  on  account  of  the  Pauli  exclusion  principle, 
or  the  Fermi  statistics,  no  two  electrons  of  the  same  spin  can  be  found  in 
the  same  stationary  state.  First,  we  investigate  the  field  inside  a  metal. 
1.  The  Electric  Field  Within  a  Metal.  —  We  have  soen  in  Chap.  XXI, 
Sec.  2,  that  an  electron  in  an  isolated  atom  is  acted  on  by  a  central  force 
on  the  average,  equal  to  the  attraction  exerted  by  the  nucleus,  diminished 
by  a  certain  shielding  effect  on  account  of  the  other  electrons.  The 
potential  energy  of  the  electron  in  such  an  atom  was  illustrated  in  Fig. 
XXII-7.  When  such  atoms  are  placed  near  each  other,  the  potential 

472 


SBC.  1] 


THE  ELECTRONIC  STRUCTURE  OF  METALS 


473 


energy  of  an  electron  at  points  between  atoms  decreases,  as  we  saw  in 
Fig.  XXII-7  (6).  In  a  crystal  of  a  metal,  with  a  lattice  of  equally  spaced 
atoms,  we  have  a  similar  situation,  with  a  potential  energy  as  shown  in 
Fig.  XXIX-1.  Here  we  show  in  (a)  the  potential  in  a  single  atom,  as 
in  Fig.  XXII-7  (a),  and  in  (6)  the  corresponding  thing  for  the  whole 
crystal.  It  will  be  seen  that,  at  points  well  within  the  crystal,  the 
potential  energy  is  a  periodic  function,  reducing  near  each  nucleus  to  the 
value  that  we  should  have  near  the  nucleus  of  an  isolated  atom  but  coming 
to  a  maximum  between  each  pair  of  nuclei.  The  field,  or  the  force  on  an 
electron,  is  given  by  the  slope  of  the  potential  energy  curve.  We  see  that 
it  fluctuates  violently  from  point  to  point,  depending  on  where  the 
electron  may  be  in  an  atom.  The  average  field,  however,  is  zero,  if  we 
average  over  many  atomic  diameters.  For  this  average  field  is  found 
from  the  difference  of  potential  energy  between  widely  separated  points, 
divided  by  the  distance  between,  and  we  see  that  on  account  .of  the 
periodicity  of  the  potential  energy  function,  the  difference  of  potential 

Potent/at  enery 


FIG.  XXIX-1. — Potential  energy  of  an  electron  in  (a)  the  central  field  representing  an 
atom;  (6)  a  periodic  field  representing  a  crystalline  solid. 


at  least  between  corresponding  points  near  different  atoms  will  be  zero. 
It  is  this  average  field  that  one  speaks  about  in  ordinary  electrical  prob- 
lems, where  we  do  not  analyze  things  on  a  microscopic  or  atomic  scale. 
We  see,  then,  that  Fig.  XXIX-1  corresponds  to  a  metal  in  which  no 
current  is  flowing,  so  that  by  Ohm's  law  the  electric  field  within  the  metal 
is  zero,  or  there  is  no  difference  of  potential  between  different  points.  A 
different  figure  would  have  to  be  drawn  in  the  case  of  a  current  flow, 
consisting  of  a  curve  like  Fig.  XXIX-1,  superposed  on  a  gradual  change 
of  potential  energy,  representing  the  field  within  the  metal.  Such  a  curve 
is  shown  in  Fig.  XXIX-2,  though  the  over-all  slope  of  the  curve  as  drawn, 
representing  the  applied  field,  is  much  greater  than  one  would  find  in  an 
actual  experimental  case.  We  shall  work  at  first  with  a  metal  in  which 
no  current  flows,  so  that  the  mean  field  is  zero,  as  in  Fig.  XXIX-1. 

As  we  go  outside  the  metal,  as  Fig.  XXIX-1  shows,  the  potential 
energy  of  an  electron  rises  and  approaches  an  asymptotic  value  at  infinity, 
much  as  the  potential  energy  of  an  electron  outside  an  atom  approaches 
an  asymptote  at  infinity.  It  is  worth  while  looking  a  little  in  detail  at 
the  nature  of  this  asymptotic  behavior.  The  force  acting  on  a  particular 


474 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 


electron  somewhat  outside  a  metal  of  course  depends  on  the  distribution 
of  the  remaining  valence  electrons,  which  are  still  attached  to  the  metal. 
It  is  well  known  from  electrostatics  that  a  negative  electron  outside  the 
metal  will  induce  a  compensating  positive  surface  charge  on  the  surface 
of  the  metal,  as  indicated  in  Fig.  XXIX-3,  where  we  show  the  electrical 


FIG.  XXIX-2. — Potential  energy  of  an  electron  in  a  field  representing  a  metal  with  a  cuirent 
flowing,  and  an  average  potential  gradient. 

lines  of  force  running  between  the  electron  and  the  induced  surface  charge. 
All  the  lines  from  the  electron  terminate  on  the  induced  charge;  thus  its 
net  amount  must  be  just  the  charge  of  one  electron.  But  this  is  what  we 
should  expect.  If  the  uncharged  metal  had  N  electrons,  and  one  is 
removed,  there  are  N  —  1  electrons  left,  and  the  positive  surface  charge 
represents  simply  the  averaged  out  deficiency  left  by  the  removed  elec- 


X      ///     l\Vx 

V//i\\\>- 


'!\ 


FIG.  XXIX-3. — Lines  of  force  between  an  electron  outside  a  metallic  surface  and  the 
positive  charge  induced  on  the  surface. 

tron.  Now  it  is  easy  to  show  that  the  lines  of  force  in  Fig.  XXIX-3  are 
just  like  those  found  from  the  electron  of  charge  —  e  at  a  distance  d  from 
the  surface,  and  an  equal  and  opposite  charge  +e  at  a  distance  d  behind 
the  surface.  This  imaginary  charge  within  the  metal  is  called  the  electric 
image  of  the  electron.  The  dotted  lines  within  the  metal  in  Fig.  XXIX-3 
represent  the  imaginary  continuation  of  the  lines  of  force,  which  really 
terminate  on  the  surface  of  the  metal,  to  the  electric  image.  But  if  we 


SBC.  21  THE  ELECTRONIC  STRUCTURE  OF  METALS  475 

had  the  two  charges  +e  and  —  e  at  a  distance  2d  from  each  other,  the  force 
exerted  by  each  on  the  other  would  be  an  attraction  of  magnitude  e2/(2d)2. 
This  force  is  called  the  image  force,  and  its  potential,  —  e2/4d,  is  the 
limiting  value  of  the  potential  energy  of  the  electron,  at  large  distances 
from  the  surface  of  the  metal.  It  is  this  function  which  the  curve  of  Fig. 
XXIX-1  approaches  asymptotically  at  large  distances. 

To  solve  the  problem  of  the  motion  of  an  electron  in  a  potential  field 
like  that  of  Fig.  XXIX-1  is  a  very  difficult  problem  in  quantum  theory. 
We  shall  describe  its  solution  in  a  later  section,  but  first  we  shall  take  up 
an  approximation  to  it,  the  free  electron  theory,  which  has  enough 
resemblance  to  the  correct  theory  so  that  it  can  be  used  satisfactorily  for 
some  purposes.  In  this  approximation,  it  is  assumed  that  the  field  acting 
on  the  electrons  in  the  metal  is  not  only  zero  on  the  average,  but  zero 
everywhere,  or  the  potential  energy  is  constant,  equal  perhaps  to  the 
average  value  of  the  potential  energy  over  the  periodic  potential  of  Fig. 
XXIX-1.  At  the  surf  ace  of  the  metal,  — y 
of  course,  the  potential  energy  must 
rise;  as  the  simplest  approximation  we  j 


may  assume  that  it  rises  discontinu-       .,      VVT__  .     ~ ..-  ,       A    A.  . 

,       „  .  .  i  .  ,      •      i  FIG.    XXIX-4. — Simplified    potential 

OUSly   from   the    Value    which    it    has    energy    function    for    the    free    electron 

within  the  metal  to  the  asymptotic  model  of  a  metal, 
value  at  infinity.  That  is,  we  replace  the  potential  energy  curve 
of  Fig.  XXIX-1  by  the  simplified  curve  of  Fig.  XXIX-4.  We  shall 
find  in  the  next  section  that  we  can  work  out  the  motion  of  electrons 
in  such  a  potential  energy  completely,  applying  the  results  to  the  prop- 
erties of  metals. 

2.  The  Free  Electron  Model  of  a  Metal. — In  Fig.  XXIX-4,  there  is  a 
volume,  which  we  shall  take  to  be  V,  in  which  the  potential  energy  has 
a  constant  value,  which  we  choose  to  be  zero.  Outside  this  volume,  the 
potential  energy  is  greater,  by  an  amount  Wa-  An  electron  whose 
energy  is  less  than  Wa  is  then  free  to  wander  freely  through  the  volume 
but  cannot  leave  it,  while  an  electron  with  energy  greater  than  Wa  can 
travel  anywhere,  but  it  suffers  a  decrease  of  Wa  in  its  kinetic  energy  on 
leaving  the  volume.  Electrons  of  the  first  sort  form  the  picture  furnished 
by  this  model  for  the  electrons  bound  in  the  metal,  while  those  of  the 
second  sort  represent  the  fast  electrons  which  can  be  emitted  in  thermionic 
emission.  Let  us  consider  first  the  electrons  with  energy  less  than  We. 
These  act  exactly  like  molecules  of  a  perfect  gas,  confined  to  the  volume 
V.  We  have  already  investigated  such  a  perfect  gas.  We  have  found  its 
distribution  of  energy  levels  in  Chap.  IV,  Sec.  1,  and  have  applied  the 
Fermi-Dirac  statistics  to  it  in  Chap.  V,  Sec.  5.  Only  one  change  must 
be  made  in  the  formulas  of  that  section  to  adapt  them  to  our  present 
use.  In  developing  the  Fermi  statistics,  we  assumed  that  only  one  mole- 


476  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

cule  could  occupy  each  stationary  state.  With  electrons,  however,  one 
electron  of  each  spin,  or  two  in  all,  can  occupy  each  state,  so  that  the 
allowed  number  of  electrons  per  energy  range  is  twice  what  we  assumed 
before.  Then  from  Eq.  (1.9)  of  Chap.  IV,  we  find  that  if  all  energy  levels 
are  filled,  the  number  of  electrons  with  energy  less  than  €  is 

m)*eH  (2.1) 


and  the  number  with  energy  in  the  range  de  is 

f 

At  the  absolute  zero,  then,  if  the  gas  contains  N  electrons,  these  will  be 
distributed  in  energy  according  to  Eq.  (2.2),  up  to  a  maximum  energy 
given  by  setting  N(e)  =  N  in  Eq.  (2.1).  This  maximum  energy,  which 
was  called  €0o  in  Chap.  V  and  which  is  usually  called  W%  in  the  theory  of 
metals,  is  given,  by  analogy  with  Eq.  (5.3),  Chap.  V,  by 


We  have  seen,  in  Eq.  (5.5)  of  Chap.  V,  that  such  energies  are  of  the  order 
of  magnitude  of  several  electron  volts.  In  Table  XXIX-1  we  give  values 
of  these  energies  for  a  series  of  metals,  computed  on  the  assumption  that 
the  number  of  electrons  equals  the  number  of  atoms,  so  that  V/N  is  the 
volume  per  atom.  This  can  be  computed  easily  from  the  lattice  spacings 
in  Table  XXVII-1,  together  with  Eq.  (3.1)  of  Chap.  XIII,  V/N  =  cr3, 
where  r  is  the  lattice  spacing,  c  a  constant  computed  in  Eqs.  (2.6),  (2.7), 
and  (2.8)  of  Chap.  XXVII.  We  have  no  particular  justification  for  the 
assumption  that  the  number  of  electrons  equals  the  number  of  atoms. 
For  the  alkali  metals,  where  each  atom  furnishes  one  valence  electron,  the 
assumption  seems  very  plausible,  and  more  elaborate  methods,  which 
will  be  described  later,  justify  it.  For  other  metals,  we  should  think  at 
first  sight  that  the  number  of  electrons  per  atom  should  be  greater  than 
unity,  since  each  atom  has  several  valence  electrons.  On  the  other  hand, 
the  more  advanced  theory  shows  in  this  case  that  the  extra  electrons  do 
not  act  very  much  like  free  electrons,  and  in  some  ways  it  is  more  reason- 
able to  take  the  number  of  free  electrons  to  be  less,  rather  than  more,  than 
the  number  of  atoms. 

A  number  of  properties  of  the  electrons  in  a  metal  can  be  found  from 
our  model.  In  particular,  we  can  find  the  specific  heat  of  the  electrons 
and  can  see  in  a  qualitative  way  what  their  contribution  to  the  equation 
of  state  should  be.  For  the  heat  capacity,  Eq.  (5.9),  Chap.  V,  gives 

(2-4) 


SBC.  2]  THE  ELECTRONIC  STRUCTURE  OF  METALS  477 

TABLE  XXIX-1. — FERMI  ENERGY  Wi  OF  METALS 


Metal 

Wi,  volts 

Li  

4.7 

Be  

9.0 

Na  

3.1 

Mg  

4.5 

Al  

5.6 

K  

2.1 

Ca  

3.0 

Ti  

5.4 

V  

6.3 

Cr  

7.0 

Fe  

7.0 

Co  

6.2 

Ni  

7.4 

Cu  

7.0 

Zn  

5.9 

Rb  

1.8 

Sr  

2.5 

Zr  

4.5 

Mo  

5.9 

Ru  

6.4 

Rh  

6.3 

Pd  

6.1 

Ag  

5.5 

Cd  

4.7 

Cs  

1.6 

Ba  

2.3 

Ta  

5.2 

w           

5.8 

Os  

6.3 

Ir  

6.3 

Pt  

6.0 

Au  

5.6 

This  is  a  heat  capacity  proportional  to  the  temperature,  and  in  Sec.  5, 
Chap.  V,  we  computed  it  for  a  particular  case,  showing  that  it  amounted 
to  only  about  1  per  cent  of  the  corresponding  specific  heat  of  free  electrons 
on  the  Boltzmann  statistics,  at  room  temperature.  In  Table  XXIX-2 
we  show  the  value  of  the  electronic  specific  heat  at  300°  abs.,  computed 
from  the  values  of  Wi  which  we  have  already  found,  in  calories  per  mole. 
We  verify  the  fact  that  this  specific  heat  is  small,  and  for  ordinary  pur- 
poses it  can  be  neglected,  so  that  the  specific  heat  of  a  metal  can  be  found 
from  the  Debye  theory,  considering  only  the  atomic  vibrations.  At  low 
temperatures,  however,  Eq.  (2.4)  gives  a  specific  heat  varying  as  the  first 
power  of  the  temperature,  while  Debye's  theory,  as  given  in  Eq.  (3.8), 


478  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

TABLE  XXIX-2.— ELECTRONIC  SPECIFIC  HEAT,  FREE  ELECTRON  THEORY 


Cv/T 
calculated 

Cv/T 
observed 

Cv  (300°) 
calculated 

Cu  

1  .  20  X  10~4 

1  .  78  X  10~4 

0.036 

Ag  

1.52 

1.6 

0.046 

Zn               

1  41 

2.33  -  2.97 

0.042 

Hg           

1.84 

3  75 

0.055 

Tl  

2.10 

3.8 

0.063 

Sn  

1.97 

3.46 

0.059 

Pb                       .... 

2.21 

7  07 

0  066 

Ta  

1.62 

27. 

0.049 

Ni 

1  13 

17  4 

0  034 

Pd 

1  38 

31 

0  041 

Pt  

1.40 

16.1 

0.042 

Observations  are  taken  from  Jones  and  Mott,  Proc.  Roy.  Soc.,  162,  49  (1937).  Calculations  in 
that  paper  assume  different  numbers  of  electrons  per  atom,  instead  of  one  per  atom  as  we  have  done, 
and  secure  somewhat  better  agreement  with  experiment.  Specific  heats  are  tabulated  in  calories  per 
mole. 

Chap.  XIV,  shows  that  the  specific  heat  of  atomic  vibrations  varies  as  the 
third  power.  At  temperatures  of  a  few  dc»croou  oK^in^  fko  third 
Of  thejemperflf.iirp  is  sn  mn^fr  seller  than  the  firs|f  pnwpr  fhof  fhn 
specific  heat  can  be  neglected,  leaving  only  the  heat  capacity  of  the 
electrons.  The  observed  specific  heat  in  this  region  is  in  fact  proportional 
to  the  temperature,  and  the  constant  of  proportionality  is  roughly  in 
agreement  with  that  predicted  by  Eq.  (2.4).  In  Table  XXIX-2  we  tabu- 
late the  values  of  CV/T  as  observed,  as  well  as  those  calculated  by  Eq. 
(2.4)  from  the  values  of  W  i  in  Table  XXIX-1.  It  is  plain  that  the 
agreement,  while  not  exact,  is  good  enough  to  indicate  the  essential  cor- 
rectness of  our  methods,  for  most  of  the  metals,  particularly  for  the 
alkalies.  For  the  transition  metals,  however,  and  in  particular  for  the 
ferromagnetic  ones,  the  observed  value  of  Cv/T  is  much  greater  than 
the  calculated  one,  a  fact  whose  explanation  we  shall  give  later,  in  Sec.  6. 
To  understand  the  relation  of  the  electrons  to  the  equation  of  state 
of  the  metal,  we  may  consider  the  internal  energy  at  the  absolute  zero  as  a 

junction  Of  Volume.      This  quantify  nf  noiir^   ahrmlH  Imvfl  fl.  mim'mnm 


for  the  actual  volume  of  the  metal,  rising  as  it  is  compressed  or  expandeH. 
In  the  free  electron  model,  the  energy  will  depend  on  volume  in  two  ways. 
In  the  first  r>lfl.fift.  f.he  lcinp.ti/^^ii^i!igirjH7^1_-flQj^^iiiJ-Qii  vnlnrne.  on  account  of 
the  fact,  proved  in  Eq.  (5.7),  Chap.  V,  that  the  total  kinetin  energy  is 
\NWj,  where  Wj  is  proportional  to  V~H.  as  shown  in  Eq.  (2.3^.  Thus 


iriiiofin  ynpfgy  T 


+hft  vnlijrnP. 


varying 


where  r  is  foe  distant  between  atoms,  afoce  V**  is  proportional  to  r, 
This  leads  to  a  repulsive  term  in  the  energy.     In  the  second  place,  the 


SBC.  2] 


THE  ELECTRONIC  STRUCTURE  OF  METALS 


479 


wi|l  depefld  on  volnirift.  on  account  of  a 


change  of  the  quantity  W*  with  volume.  It  requires  a  little  more  careful 
analysis  to  see  just  what  this  change  means,  but  it  turns  out  that  the 
situation  is  as  shown  in  Fig.  XXIX-5.  Here  we  show  potential  energy 
curves  like  Fig.  XXIX-1  for  two  different  distances  of  separation.  As 
the  distance  decreases  and  the  atoms  overlap  more,  we  see  that  the 
potential  energy  of  an  electron  between  two  atoms  decreases.  In  fact,  if 
the  total  potential  energy  is  the  sum  of  Coulomb  attractions  to  the  various 
atoms  (a  crude  approximation,  on  account  of  the  other  electrons,  which 
complicate  the  situation),  the  potential  energy  at  a  given  point  of  the 
crystal  would  be  a  sum  of  terms  —  1/d,  where  the  d's  are  the  distances  to 


FIG.  XXIX-5. — Potential  energy  of  an  electron  in  a  metallic  crystal,  for  two  differ- 
ent distances  of  separation,  illustrating  the  decrease  of  average  potential  energy  with 
decreasing  lattice  spacing. 

the  various  atoms  of  the  crystal.  Then,  since  each  of  these  d's  is  propor- 
tional to  r,  the  lattice  spacing,  we  find  that  the  potential  energy  at  a  point 
between  the  atoms  should  decrease  approximately  like  —  1/r.  We  can 
take  an  average  value  over  the  periodic  potential  of  Fig.  XXIX-5  to 
represent  the  zero  of  potential  in  the  free  electron  picture.  We  see,  in 
other  words,  that  this  should  not  be  chosen  as  being  really  zero  potential, 
but  that  it  should  be  given  a  negative  value,  roughly  proportional  to 
1/r.  This  value,  then,  is  the  mean  potential  energy  of  the  electrons. 
The  net  result  is  that  we  might  expect  to  express  the  internal  energy  of 
the  crystal  at  the  absolute  zero  as 


A 


B 


(2.5) 


where  the  first  term  represents  the  potential  energy,  the  second  the 
kinetic  energy,  and  where  the  constant  B  is  given  theoretically  by 


B  - 


(2.6) 


480  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

using  Eq.  (2.3),  and  Eq.  (3.1),  Chap.  XIII,  for  the  definition  of  c.  The 
internal  energy  (2.5)  is  the  general  sort  we  expect,  with  a  minimum  and  an 
asymptotic  value  at  infinity,  not  entirely  unlike  a  Morse  curve.  As  a 
matter  of  fact,  it  is  not  very  satisfactory  for  actually  describing  the 
equation  of  state  of  a  metal,  on  account  of  various  approximations  that 
enter  into  it.  First,  the  electrons  of  real  metals  are  not  free  and  their 
kjpetifl  energy  is  not  given  at  all  accurately  by  the  free  electron  theory. 
Secondly,  the  potential  energy  does  not  vary  as  1/r,  being  really  much 
more  complicated  than  this.  And  finally,  the  electrons  do  not  really 
move  in  an  averaged  out  external  field  at  all  but  are  acted  on  by  all  the 
other  electrons  in  their  instantaneous  motions.  For  all  these  reasons,  as 
simple  an  expression  as  Eq.  (2.5)  is  not  adequate  and  is  not  actually  as 
satisfactory  as  the  Morse  function,  which  we  have  used  in  Chap.  XXVII, 
Sec.  2.  In  spite  of  this,  it  gives  some  insight  into  the  mechanism  of  the 
forces  in  the  metal,  describing  them  in  rather  different  light  from  that 
used  in  Chap.  XXII,  Sec.  4,  where  we  treated  them  as  a  somewhat  varied 
form  of  homopolar  bonds. 

3.  The  Free  Electron  Model  and  Thermionic  Emission. — We  have 
stated  in  the  last  section  that  electrons  whose  kinetic  energy  is  greater 
than  Wa  can  escape  from  the  volume  V  in  Fig.  XXIX-4,  furnishing  a 
model  for  thermionic  emission.  Using  this  free  electron  picture,  we  can 
easily  calculate  the  rate  of  thermionic  emission  from  a  metal  directly. 
This  of  course  must  lead  to  the  same  result  as  the  indirect  method  of 
Chap.  XXVIII,  Sec.  3,  for  that  is  entirely  justified  thermodynamically, 
but  it  may  lead  to  greater  insight  into  the  mechanism  of  thermionic 
emission.  In  the  free  electron  theory,  the  electrons  within  the  metal 
form  a  perfect  gasr  and  we  can  find  foe*  ^nmhp.r  of  electrons  emerging_per 
second  by  finding  the  number  hitt.inpr  fop  mirfan.ft  Inypy  ^f  t.l^  mftV1  frn™ 
the  inside  per  second,  and  by  multiplying  hy  (1  —  r)1  where  r  is  the 
reflection  coefficient,  as  in  Chap.  XXVIII.  We  must  take  account  of  one 
fact  here,  however,  which  was  absent  in  our  calculation  of  the  last  chapter: 
we  must  confine  ourselves  to  electrons  of  energy  greater  than  Wa,  so  that, 
after  passing  the  barrier,  they  will  still  have  a  positive  kinetic  energy  and 
a  real  velocity. 

The  situation  of  the  barrier  is  different  from  what  it  would  be  with  the 
Boltzmann  statistics,  as  can  be  seen  most  clearly  from  Fig.  XXIX-6. 
Here  we  have  drawn  energies,  both  inside  and  outside  the  metal,  as  in 
Fig.  XXIX-4.  The  zero  of  energy  is  taken  to  be  at  the  bottom  of  the 
picture.  Then  at  the  absolute  zero  there  will  be  filled  energy  levels  up  to 
the  energy  W%  and  empty  levels  above,  the  filled  ones  being  shaded  in 
Fig.  XXIX-6.  We  can  now  see  that  the  potential  energy  of  an  electron 
outside  the  metal,  Wa,  is  related  to  Wi  and  to  L0,  the  heat  of  vaporization 


SEC.  3] 


THE  ELECTRONIC  STRUCTURE  OF  METALS 


481 


of  a  mole  of  electrons  at  the  absolute  zero,  discussed  in  the  last  chapter,  by 
the  equation 


Wa 


(3.1) 


For  Lo  represents  the  energy  necessary  to  remove  a  mole  of  electrons  from 
the  metal  at  the  absolute  zero,  in  equilibrium.  For  equilibrium,  the 
metal  must  be  left  in  its  lowest  state,  so  that  the  removed  electrons  must 
come  from  the  top  of  the  Fermi  distribution,  and  they  must  have  no 
kinetic  energy  after  they  are  removed  from  the  metal.  Thus  each 
electron  is  raised  just  through  the  energy  LQ/N  in  the  figure.  In  the 
Fermi  statistics,  in  other  words,  the  work  function  represents  the  differ- 
ence in  energy  between  the  top  of  the  Fermi  distribution  and  space  outside 
the  metal.  And  the  result  of  Sec.  4,  Chap.  XXVIII,  that  on  account  of 
the  contact  potential  the  values  of  Ea  —  La  and  Eb  —  Lb  were  equal  for  two 
metals  at  the  absolute  zero,  means  graphically  that  two  metals  will  adjust 


Fia.  XXIX-6. — Occupied  energy  levels  for  the  free  electron  model  of  a  metal,  at  the 
absolute  zero,  illustrating  the  relation  between  Wa,  W%,  and  the  thermionic  work  function 
or  latent  heat  of  vaporization  of  electrons. 

their  potentials  so  that  the  top  of  the  Fermi  distribution  is  at  the  same 
height  in  each  metal,  at  the  absolute  zero.  Let  us  find  the  small  modifica- 
tions in  this  occurring  at  higher  temperatures. 

The  exact  statement  of  Eq.  (4.3),  Chap.  XXVIII,  is  that  in  any  two 
metals  in  equilibrium,  the  space  outside  the  metals  acquires  such  a 
potential  that  if  we  subtract  from  it  the  amount 


(3.2) 


the  result  is  the  same  for  all  metals.  If  it  were  not  for  the  second  term, 
as  we  have  just  mentioned,  this  would  take  us  down  to  the  top  of  the 
Fermi  distribution  for  all  metals.  We  shall  how  show  that  the  last  term 
m£ans  that  really  it  will  take  us  down  to  the  level  c0,  rather  than  Wi  or 
eoo,  as  defined  in  Chap.  V,  Sees.  3  and  4,  for  all  metals,  so  that  any  two 
metals  in  equilibrium  have  their  €0's  at  the  same  energy.  To  do  this,  let 
us  compute  the  second  term  of  Eq.  (3.2).  We  do  not  have  CP\  but  we 
shall  assume  for  the  present  that  the  metal  has  no  thermal  expansion, 
so  that  CP  =  Cvt  and  we  can  use  the  result  of  Eq.  (2.4).  Substituting  for 
Wi  from  Eq.  (2.3),  this  gives  CV  proportional  to  N*.  Thus  we  have 


482  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

dCv       1  Cv 


Then,  substituting  in  Eq.  (3.2),  and  using  Eq.  (5.6),  Chap.  V,  we  have 

Lo  +  tfo^  ^  -  Lo  +  tf  o(eoo  -  «o).  (3.4) 

But,  referring  to  Fig.  XXIX-6,  we  see  that  an  energy  lower  by  the  amount 
(3.4)  than  the  energy  outside  the  metal,  is  just  at  the  level  eo,  verifying  our 
statement  that  this  quantity  is  the  same  for  all  metals  in  equilibrium, 
while  the  energies  outside  the  metals,  and  the  bottoms  of  the  Fermi 
distributions,  are  different  for  different  metals. 

Now  we  can  proceed  with  our  calculation  of  thermionic  emission.     We 
wish  to  find  how  many  electrons,  with  energy  sufficient  to  surmount  the 

barrier  of  height  Wa  =  Wi  +  -^>  strike  a  square  centimeter  of  the  surface 

J\l  o 

of  the  metal  per  second.  Let  the  x  axis  be  normal  to  the  surface.  Then 
in  going  through  the  barrier,  the  x  component  of  kinetic  energy,  pl/2m, 
will  be  reduced  by  Wa,  while  the  y  and  z  components  will  be  unchanged. 
In  other  words,  the  electrons  we  are  interested  in  are  those  whose  x 
component  of  momentum  is  greater  than  \/2mWa,  while  their  y  com- 
ponents of  momentum  can  be  anything.  In  finding  this  number,  we  must 
integrate  over  px,  py,  pz,  rather  than  over  the  energy,  for  our  limits  of 
integration  depend  on  the  p's.  We  ask  first,  then,  how  many  electrons 
with  momentum  in  the  range  dpx  dpv  dpg  will  cross  1  sq.  cm.  per  second 
normal  to  the  x  axis.  This,  by  the  same  methods  used  in  the  last  chapter, 
will  be  p9/m  times  the  number  of  electrons  per  unit  volume  in  the  range 
dpx  dpv  dp,.  And  in  turn  the  number  per  unit  volume  in  that  range  will 
be  the  number  of  energy  levels  in  that  range,  multiplied  by  the  Fermi 
factor 

1 


(p«/2m-«o) 

e      kT      +  1 

giving  the  fraction  of  those  levels  occupied  by  electrons,  and  divided  by 
the  volume  V  of  the  gas  to  find  the  number  of  electrons  per  unit  volume. 
We  must  then  find  the  number  of  energy  levels  in  dpx  dpv  dp,.  As  in 
Chap.  IV,  Sec.  1,  this  is  (V/h*)dpx  dpv  dpz,  except  for  the  correction  arising 
from  the  electron  spin.  This  doubles  the  number  of  allowed  energy  levels, 

2V 
leading  to    rdp»  dpv  dp,  as  the  number  of  energy  levels  in  dpx  dpv  dp,. 


SBC.  3J  THE  ELECTRONIC  STRUCTURE  OF  METALS  483 

We  are  now  prepared  to  find  how  many  electrons,  of  energy  sufficient 
to  cross  the  barrier,  strike  1  sq.  cm.  of  the  boundary  of  the  metal  per 
second.  Using  the  statements  made  above,  this  is 


J-  > 


The  integral  in  Eq.  (3.5)  cannot  be  completely  evaluated  analytically. 
But  for  the  high  energy  of  the  electrons  concerned,  we  are  entirely  justi- 
fied in  neglecting  the  term  unity  in  the  denominator  of  the  Fermi  function. 
Then  the  calculation  can  be  carried  out  at  once,  giving 


mh\]  -  .  ""'I  _  „  ^  V  V2SHT. 


where  in  the  integral  over  px  we  can  introduce  pi  as  a  new  variable  to 
simplify  the  integration.  When  we  multiply  by  the  factor  (1  —  r), 
representing  the  fraction  of  electrons  that  penetrate  the  barrier,  and  when 
we  consider  Eqs.  (3.2)  and  (3.4),  and  Eq.  (3.3)  of  Chap.  XXVIII,  we  see 
that  this  result  agrees  exactly  with  that  found  in  Chap.  XXVIII.  As  a 
matter  of  fact,  there  is  nothing  in  this  simple  model  that  would  lead  to 
any  reflection  coefficient  at  all  for  electrons,  so  that  we  should  really  set 
r  =  0. 

We  have  seen,  in  other  words,  that  our  free  electron  model,  using  the 
Fermi  statistics,  leads  to  thermionic  emission  agreeing  with  our  previous 
deduction  from  thermodynamics.  This  is  hardly  remarkable,  for  any 
model  whatever,  correctly  worked  out  according  to  thennodynamic 
principles,  would  have  to  do  the  same  thing.  But  we  have  a  somewhat 
better  physical  understanding  of  the  reason  for  the  rapid  increase  in 
emission  with  increasing  temperature:  only  those  electrons  that  happen 
to  have  enough  energy  inside  the  metal  to  surmount  the  barrier  can 
escape;  the  number  of  such  electrons  increases  very  rapidly  with  increas- 
ing temperature,  the  more  rapidly  the  lower  the  work  function  is. 

The  model  we  have  used  is,  of  course,  too  simplified.  A  more  accurate 
model  would  likewise  have  to  agree  with  our  deduction  of  the  previous 
chapter,  but  it  could  differ  in  its  final  results  from  the  free  electron  model  in 
having  a  different  reflection  coefficient  and  a  different  value  of  No(dCp/dN). 
And  the  calculation  would  differ  in  that  it  is  only  with  free  electrons, 
unperturbed  by  atoms,  that  we  can  find  the  number  colliding  with  1  sq. 
cm.  of  surface  per  second  as  simply  as  we  have  done  here.  Then  in  wave 
mechanics  the  reflection  coefficient  is  not  so  simple  as  in  the  classical 


484  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

theory  of  free  particles.  Finally,  the  change  of  work  function  with 
temperature,  coming  from  the  term  No(dCP/dN),  is  a  decidedly  more 
complicated  thing  than  we  have  assumed.  On  account  of  thermal 
expansion,  the  volume  of  an  actual  metal  changes  with  temperature. 
Then  the  energy  levels  of  the  electrons  are  bound  to  change,  bringing  a 
change  explicitly  with  volume  but  actually  with  temperature,  in  the 
height  of  the  barrier,  in  €0,  and  so  on,  quite  apart  from  anything  we  have 
had  to  take  up.  All  these  complications  make  a  direct  calculation  of 
thermionic  emission  from  a  correct  model  a  very  difficult  thing.  And 
they  make  the  calculation  of  the  preceding  chapter  all  the  more  important, 
for  it  is  derived  from  straightforward  thermodynamics  and  from  the 
properties  of  the  electron  gas  in  empty  space,  about  which  there  can  be 
no  doubt,  and  it  does  not  depend  on  the  nature  of  the  metal  at  all. 

4.  The  Free  Electron  Model  and  Electrical  Conductivity. — By  slight 
extensions  of  the  free  electron  theory,  one  can  explain  the  electrical 
conductivity  of  a  metal.  Let  there  be  an  external  electric  field  E, 
impressed  on  the  metal.  Then  the  electrons  will  no  longer  move  with 
iniform  velocity  in  a  straight  line,  as  in  a  perfect  gas.  Instead,  they  will 
be  accelerated,  the  time  rate  of  change  of  momentum  of  each  electron 
equaling  the  external  force.  If  the  field  is  applied  at  a  certain  instant, 
this  external  field  by  itself  would  result  in  a  net  electric  current,  building 
up  proportionally  to  the  time,  if  no  other  forces  acted  on  the  electrons. 
We  must  assume  in  addition,  however,  that  the  electrons  meet  some  sort 
of  resistance,  proportional  to  their  average  or  .drift  velocity,  which  hence 
is  proportional  to  the  current.  When  we  consider  this  resistance,  we  find 
that  the  current,  instead  of  building  up  indefinitely,  approaches  an 
asymptotic  value,  proportional  to  the  external  field,  and  hence  obeys 
Ohm's  law.  The  time  taken  to  reach  this  steady  state  is  quite  negligible 
in  comparison  with  ordinary  times,  so  that  we  find  that  the  current  in 
the  conductor  is  proportional  to  the  field.  We  can  easily  formulate  this 
argument  mathematically  and  see  how  the  electrical  conductivity  depends 
on  various  properties  of  the  electrons.  The  argument  does  not  depend 
on  the  Fermi  statistics  and  follows  equally  well  from  Boltzmann  statistics. 

There  are  N/V  electrons,  each  of  charge  --e,  per  unit  volume  in  the 
metal.  These  will  have  a  certain  distribution  of  velocities  (the  Maxwell 
distribution  or  the  Fermi  distribution),  of  which  we  need  only  the  prop- 
erty that  the  mean  velocity  is  zero,  in  the  absence  of  an  external  field.  In 
the  field  E,  the  force  on  each  charge  will  be  —  eE,  so  that,  if  px  is  the 
component  of  momentum  of  an  electron  in  the  direction  of  the  field  (which 
we  take  to  be  along  the  x  axis),  we  shall  have 


SBC.  4]  THE  ELECTRONIC  STRUCTURE  OF  METALS  486 

But  the  velocity  of  an  electron  is  its  momentum  divided  by  its  mass,  and 
the  electric  current  density  u  is  the  number  of  electrons  per  cubic  centi- 
meter, times  the  charge  on  each  electron,  times  the  velocity  of  each. 
Thus  we  have 


so  that  Eq.  (4.1)  can  be  rewritten 


from  which  we  can  at  once  verify  that  the  current  increases  proportion- 
ally to  the  time.  But  now  let  us  assume  that  there  is  an  additional  force 
acting  on  the  electrons,  proportional  to  their  velocity,  of  the  nature  of  a 
viscous  resistance.  Let  the  force  acting  on  a  single  electron  be  —  p»/r, 
where  r  is  a  constant.  Then  Eq.  (4.1)  becomes 

rfj?  =  -eE  -  &.  (4.4) 

The  meaning  of  r  is  easily  seen  if  we  ask  for  a  solution  of  Eq.  (4.4)  in  the 
case  where  the  external  field  is  zero.  Then  we  have  at  once 

_* 
px  =  const,  e  r,  (4.5) 

so  that  r  is  the  time  in  which  the  original  momentum  of  an  electron  would 
be  reduced  to  l/e  of  its  value  on  account  of  the  friction.  The  quantity  r 
is  sometimes  called  a  relaxation  time.  Then  we  have 

fa^N^E-*,  (4.6) 

of  which  the  solution  satisfying  the  initial  condition  of  no  current  at 
t  =  Ois 


e  (4.7) 

which  reduces  to 

«•-?£*-»*>  (4-8) 

at  times  large  compared  to  the  relaxation  time,  where  by  definition  o  is 
the  electrical  conductivity,  given  on  the  free  electron  theory  by 


486  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

The  conductivity  increases  as  we  see  with  the  number  of  free  electrons 
available  to  carry  the  current  and  with  the  time  in  which  each  one  can 
he  speeded  up  by  the  field  before  it  reaches  a  stationary  speed  on  account 
of  the  resistance.  It  is  obvious  that  Eq.  (4.8),  though  it  gives  an  explana- 
tion of  Ohm's  law,  does  not  lead  to  a  calculation  of  the  conductivity  in 
terms  of  known  quantities,  because  though  we  have  seen  how  to  estimate 
N/V,  there  is  no  way  of  estimating  the  relaxation  time  r.  We  can,  of 
course,  reverse  the  argument,  and  from  known  conductivities  and  the 
values  of  N/V,  assumed  in  Table  XXIX-1,  find  what  values  of  relaxation 
time  would  be  required.  These  times  are  given  in  Table  XXIX-3,  from 
which  we  see  that  they  are  very  short,  of  the  order  of  10~14  sec. 

TABLE  XXIX-3. — RELAXATION  TIMES  FOR  ELECTRICAL  CONDUCTIVITY,  FREE  ELEC- 
TRON THEORY 

T,  seconds 

Li 0.89  X  10~14 

Na  3.26 

Mg 1.88 

Al 2.20 

K 4.07 

Ca ...  3.31 

Ti .  .  1.98 

Cr .  ...  ..  .  1.62 

Fe .  0.42 

Co  .  .  ...  ....  0.52 

Ni          ..    .  .  0.56 

Cu  ...  ..  2.50 

Rb.  .  ...  2.70 

Sr .  .    .  0.79 

Mo.      .  ...  0.97 

Pd.         .          .  .  0.50 

Ag      ..  .  ....  3.96 

Cd  ....  .  0.97 

Cs    .  .  7.50 

Ta .  0  41 

W .  .  1.00 

Os.         ...  ...  .      .  .  0.08 

Pt     .    .    .  .  .  ....  0.48 

Au 2.43 

The  electrical  conductivities  used  in  computing  this  table  are  either  for  0  or  20°C. 

There  is  nothing  in  the  free  electron  theory  to  explain  the  existence  of 
the  resisting  force  or  the  relaxation  time.  It  is  usually  described  as 
coming  from  collisions  between  the  electrons  and  the  atoms,  and  thus 
cannot  properly  be  explained  unless  we  take  the  atoms  into  account 
specifically.  We  can  see  something  about  the  mechanism  of  the  collisions, 
however,  by  considering  the  motion  of  the  electrons  in  a  momentum 
space,  similar  to  that  of  Chap.  IV,  Sec.  1,  in  which  px,  pv,  pg  are  plotted  as 
variables,  and  each  electron  is  represented  by  a  point.  As  we  saw  in  that 


SBC.  4]  THE  ELECTRONIC  STRUCTURE  OF  METALS  487 

section,  we  may  imagine  a  lattice  of  points  in  momentum  space,  one  to  a 
volume  h*/V,  each  point  corresponding  to  an  energy  level,  so  that  there 
can  be  two  electrons,  one  of  each  spin,  at  each  point.  According  to  the 
Fermi  distribution,  at  the  absolute  zero,  the  N  points  will  be  located  at 
the  points  within  a  sphere  of  radius  p  =  \/2mWi  about  the  origin,  so 
that  the  density  of  electrons  within  the  sphere  will  be  the  maximum 
allowable  value,  while  none  will  be  found  outside  the  sphere.  At  higher 
temperatures,  the  distribution  will  vary  gradually,  rather  than  discon- 
tinuously,  from  the  maximum  density  within  the  sphere  to  zero  outside, 
but  the  change  will  come  within  a  very  narrow  region  about  the  surface  of 
the  sphere1.  Now  if  there  is  an  external  impressed  field,  the  momentum 
of  each  electron  will  increase  proportionally  to  the  time.  That  is,  each 
of  the  points  will  move  in  the  x  direction  (if  the  field  is  along  x)  with  a 
velocity  given  by  Eq.  (4.1).  Since  this  velocity  is  the  same  for  all  points, 
the  whole  sphere  of  points  will  drift  along  the  x  axis  with  uniform  velocity, 
so  that  after  a  time  t  it  will  be  displaced  a  distance  —  eEt  along  the  px 
axis,  since  —eE  =  dpx/dt  is  the  velocity  of  the  points. 

Now  let  us  consider  the  effect  of  collisions  on  the  distribution  function. 
The  original  distribution  was  in  equilibrium,  according  to  the  Fermi 
distribution,  so  that  collisions  will  leave  it  unchanged,  but  the  displaced 
distribution  is  not.  Collisions  of  electrons  with  lattice  points  correspond 
to  a  sudden  jump  of  a  representative  point  from  one  region  of  momentum 
space  to  another.  There  are  at  least  two  principles  governing  such  jumps 
that  we  can  understand  easily.  In  the  first  place,  an  electron  cannot 
jump  to  a  stationary  state  which  is  already  occupied,  on  account  of  the 
exclusion  principle.  In  the  second  place,  if  the  collision  of  aiTelecTffon 
and  an  atom  is  elastic,  as  we  shall  assume,  it  takes  place  with  constant 
energy  or  only  slight  dissipation  of  energy  in  the  form  of  elastic  vibrations 
of  the  lattice,  so  that  the  representative  point  jumps  from  one  location  to 
another  at  the  same  or  slightly  smaller  distance  from  the  origin,  in  the 
momentum  space.  Then  we  can  see  in  Fig.  XXIX-7  that  there  are  only 
a  few  collisions  possible.  In  this  figure  we  show  the  undisplaced  sphere, 
the  displaced  sphere,  and  the  crescents  A  and  J5,  in  which  A  includes 
points  having  more  energy  than  Wi,  representing  electrons  that  have 
been  accelerated  by  the  field,  and  B  represents  points  of  less  energy  than 
Wi,  from  which  electrons  have  been  removed  by  action  of  the  field.  The 
likely  collisions  are  essentially  those  in  which  a  point  from  'crescent  A 
jumps  to  an  unoccupied  point  in  5.  These  collisions  have  the  effect  of 
disturbing  the  distribution,  reducing  the  number  of  points  with  positive 
momentum,  -and  increasing  the  number  with  negative  momentum,  verify- 
ing our  statement  that  the  distribution  was  not  in  equilibrium,  so  that 
collisions  disturb  it.  We  can  now  understand  the  general  nature  of  the 
motion  of  representative  points,  subject  to  the  external  field  and  to 


488 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 


collisions.  They  stream  steadily  in  the  direction  of  increasing  x,  on 
account  of  the  field,  but  as  points  enter  the  crescent  A  they  are  likely  to 
have  collisions  taking  them  back  to  B  and  starting  the  process  all  over 
again.  An  equilibrium  will  be  set  up,  in  which  the  number  of  points 
entering  A  per  second  will  be  equal  to  the  number  of  collisions  sending 
points  from  A  to  B  per  second.  These  collisions  prevent  the  sphere  from 
moving  indefinitely  far  to  the  right;  the  farther  it  goes7~tEe"~gfeater  the 
crSce^"TBeco"mes,~the  more  collisions  there  are,  so  that  equilibrium 
corresponds  io Ta  finite  displacement  of  the  sphere,  though  the  individual 
points  are  "moving  as  we  have  just  stated. 


Undispl&ced 
sphere 


-Displaced 
sphere 


Fia.  XXIX-7. — Diagram  of  occupied  levels  in  momentum  space,  in  free  electron  model 
of  a  metal.  The  points  of  the  displaced  sphere  (shaded)  are  occupied  when  the  electrons 
have  been  accelerated  by  an  external  field. 

The  current  density  is  proportional  to  the  mean  momentum  of  the 
electrons  or  to  the  displacement  of  the  sphere  from  the  center  of  the 
momentum  space.  The  time  rate  of  change  of  current,  or  the  velocity 
of  the  center  of  the  sphere,  then  depends  on  two  things:  on  the  velocity 
of  the  individual  points,  which  is  proportional  to  the  external  field,  and 
on  the  number  of  points  leaving  A  and  entering  B  per  second.  The 
collisions,  represented  by  these  points  jumping  from  A  to  B,  have  the 
effect  of  slowing  down  the  motion  of  the  sphere  as  a  whole,  though  not  of 
its  individual  points.  The  number  of  collisions  is  proportional  to  the 
number  of  points  in  A,  and  this  is  proportional  to  its  area,  which  in  turn 
is  proportional  to  the  displacement  of  the  sphere  or  to  the  current  density. 
Thus  we  see  from  our  mechanism  that  we  have  essentially  the  two  terms 
in  dux/dt  given  in  Eq.  (4.6),  though  the  second  term,  proportional  to  the 
current  density,  arises  from  collisions  rather  than  from  a  frictional  force. 
It  is  then  clear  that  if  we  can  calculate  the  probability  of  collision  we  can 


SBC.  5J  THE  ELECTRONIC  STRUCTURE  OF  METALS  489 

compute  the  conductivity,  using  Eq.  (4.9).  We  cannot  go  further  with 
this,  however,  without  considering  the  interaction  of  electrons  and  atoms 
more  in  detail.  We  can  only  say  that  the  probability  of  collision  is 
inversely  proportional  to  the  relaxation  time  ry  since  doubling  the  number 
of  collisions  will  halve  the  time  required  for  dissipating  the  momentum,  so 
that  the  conductivity  should  be  inversely  proportional  to  the  probability 
of  collision,  or  the  specific  resistance  should  be  directly  proportional  to  the 
probability  of  collision,  a  very  reasonable  result. 

5.  Electrons  in  a  Periodic  Force  Field. — In  the  three  preceding  sec- 
tions, we  have  been  dealing  with  the  free  electron  approximation,  in 
which  we  assume  that  the  electrons  in  a  metal  are  not  acted  on  by  any 
forces.  Now  we  shall  give  a  brief  and  qualitative  discussion  of  the 
changes  brought  about  when  we  remember  that  the  electrons  are  really 
acted  on  by  a  periodic  force  field,  as  shown  in  Fig.  XXIX-1.  It  is  quite 
impossible  to  understand  these  changes  without  knowing  a  few  simple 
facts  about  wave  mechanics,  and  we  shall  proceed  to  give  some  simple 
illustrations  of  the  wave  nature  of  the  electron. 

It  has  been  shown,  both  theoretically  and  experimentally,  that  an 
electron  of  momentum  p  has  many  of  the  properties  of  a  wave  of  wave 
length  X  =  h/p,  where  h  is  Planck's  constant.  This  is  shown  most 
clearly  in  an  experimental  way  by  the  phenomenon  of  electron  diffraction, 
in  which  a  beam  of  electrons  striking  a  crystal  is  diffracted  much  as  a 
beam  of  x-rays  of  the  same  wave  length  would  be.  Theoretically  the 
wave  conception  of  electrons  is  shown  most  clearly  in  the  explanation  of 
the  quantum  condition,  the  condition  used  in  Chap.  Ill,  Sec.  3,  to  fix  the 
energy  levels  in  the  quantum  theory.  We  shall  illustrate  this  by  the 
problem  in  which  we  are  particularly  interested  at  present,  the  perfect 
gas.  Consider  a  particle  moving  under  the  action  of  no  forces  in  a  region 
bounded  by  x  =  0,  x  =  X,  y  =  0,  y  =  F,  z  =  0,  z  =  Z,  or  in  a  volume 
XYZ  =  V.  Classically,  if  the  particle  starts  out  with  components  of 
momentum  px,  pv,  pz,  it  will  suffer  various  reflections  at  the  walls  of  the 
region,  traveling  between  collisions  with  the  walls  in  a  straight  line  with 
constant  velocity  and  momentum.  At  each  reflection,  the  component  of 
momentum  perpendicular  to  the  wall  will  be  changed  in  sign  but  not  in 
magnitude,  while  the  other  two  components  will  be  unchanged.  When 
we  take  account  of  all  possible  reflections,  then,  we  shall  find  the  particle 
traveling  equal  fractions  of  the  time  with  the  eight  possible  momenta 
given  by  the  possible  combinations  of  sign  in  ±px,  ±pv,  ±pz.  Corre- 
sponding to  this,  in  wave  mechanics,  a  particle  of  momentum  p*,  pv,  pz 
is  associated  with  a  plane  wave  of  the  form 


sin 


490  INTRODUCTfO^mfLCHtiMlCAL  PHYSICS      [CHAP.  XXIX 


analogous  to  Eq.  (2.8),  Chapter  JHl,  where  we  were  discussing  elastic 
waves.  The  quantities  Z/X,  w/X,  n/X,  in  Eq.  (5.1),  are  given  by  the 
relations 

1    _  P*  m  _  Pv  n  _  P*  /K  9\ 

x"T      x"P      x"T  (6'2) 

the  vector  form  of  our  previous  relation  X  =  A/P-  Now  when  we  com- 
bine the  waves  corresponding  to  the  various  combinations  of  ±  signs,  we 
find  a  standing  wave,  just  as  we  did  in  a  similar  case  in  Chap.  XIV,  Sec.  2. 
It  turns  out  that  the  wave  must  satisfy  boundary  conditions  of  reducing 
to  zero  on  the  surfaces  of  the  enclosure,  at  x  =  0,  x  =  X,  etc.  Then, 
just  as  in  Eqs.  (2.11)  and  (2,12),  Chap.  XIV,  we  must  have  a  wave 
represented  by  the  function 


A    .     0     .    .  .     2irmy    .     2trnz  ,_  ON 

A  sin  2irvt  sm  —  —  sin  —  r-2  sin  —  r—  >  (5.3) 

A  A  A 

where  in  order  to  satisfy  the  condition  that  the  amplitude  is  zero  at 
x  =  X,  etc.,  we  must  have 

21X  2mY  2nZ  ,e  A. 

-y-  =  **,         -y-  =  *ir,         —  =  *•>  (5.4) 

where  8Xt  sv,  s»  are  positive  integers.  Using  Eq.  (5.2),  we  can  rewrite 
these  equations  as 

sxh  syh  szh  ,      . 

P*  =  2X>        Pv  =  2Y'        P*  =  2Z'  (6<5) 

The  conditions  (5.5)  determine  the  momentum,  and  hence  the  energy,  of 
the  electron.  But  they  are  just  the  conditions  that  would  be  determined 
by  the  quantum  theory,  as  in  Chap.  Ill,  Sec.  3.  There  we  determined 
energy  levels  by  considering  a  phase  space  and  by  demanding  that  the 
area  of  the  curve  enclosed  by  the  path  of  the  representative  point  in  the 
phase  space  be  an  integer  times  h.  In  Fig.  XXIX-8  we  show  the  x  —  px 
projection  of  the  phase  space  for  the  present  case.  As  the  particle  travels 
from  x  =  0  to  x  =  X,  its  x  component  of  momentum  is  px.  At  X  there 
is  a  collision  with  the  wall  and  the  momentum  changes  to  —  pxy  again 
reversing  when  the  particle  rejburns  to  x  =  0.  The  area  of  the  rectangular 
path  is  then  2Xpx,  so  that  we  must  have 

2Xp,  =  a,        P*  =  ^>  (5.6) 

where  sx  is  an  integer,  with  similar  relations  for  the  y  and  z  components. 
But  Eqs.  (5.6)  are  identical  with  Eqs.  (5.5),  showing  that  the  wave  con- 
ception of  the  electron  leads  to  the  same  quantum  conditions  that  we 
found  earlier  by  quantizing  the  areas  of  the  cells  in  the  phase  space. 


SBC.  5] 


THE  ELECTRONIC  STRUCTURE  OF  METALS 


491 


The  conditions  (5.5)  or  (5.6)  are  not  exactly  the  same  that  we  used 
earlier,  in  Eq.  (1.6),  Chap.  IV,  in  considering  the  same  problem.  There 
we  found 


P* 


Pv  =  -- 


nnh 
P*  -  -> 


(5.7) 


where  nxj  Uy,  n,  were  integers.     The  difference  is  that  in  Chap.  IV  we 
were  not  considering  the  collisions  with  the  wall  and  the  reversal  of 
momentum  produced  by   those  collisions.     Our  present  treatment  is 
correct,  both  according  to  ordinary  quantum 
theory  and  to  wave  mechanics,  for  a  gas  really 
confined  in  a  box.     The  results  are  just  the 
same,  as  far  as  the  distribution  of  energy    Px 
levels  is  concerned.     In  a  momentum  space, 
Eq.  (5.5)  gives  a  lattice  of  allowed  momentum 
values,  each  value  corresponding  to  a  volume 
(h/2X)(h/2Y)(h/2Z)  =  A3/87  of  momentum 
space.     On  the  other  hand,  Eq.  (5.7)  leads      o 
to  a  volume  h?/V  for  each  allowed  value,  so 
that  there  are  only  an  eighth  as  many  mo- 
menta in  a  given  range  in  Eq.  (5.7)  as  in  Eq. 
(5.5).     But  the  integers  in  Eq.  (5.7)  can  be 
positive  or  negative,  while  those  in  Eq.  (5.5)      x 
must  be  positive.     Thus  in  Eq.  (5.7)  we  have       FlG.  xxix-8.-Path  of  repre- 

points    in    all    eight    OCtailts    of    momentum    scntative    point    in    momentum 
i  •!      •      TI        /f  f\    Ji  n       ^    space,  for  a  particle  reflected  from 

space,  while  in  Eq.  (5.5)  they  are  confined  ^alls  at  x  =  0  and  x  -  x. 
to  the  first  octant.     The  number  of  allowed 

points  within  a  sphere  of  given  radius,  corresponding  to  given  energy, 
is  then  the  same  according  to  either  equation,  leading  in  either  case 
to  Eq.  (1.9),  Chap.  IV,  for  the  number  of  states  with  energy  less  than 
«,  which  we  have  used  as  the  basis  of  our  treatment  of  the  perfect  gas. 
While  the  two  expressions  (5.5)  and  (5.7),  then,  are  equivalent  as  far 
as  the  distribution  of  energy  levels  is  concerned,  the  argument  on 
which  Eq.  (5.7)  is  based  is  more  satisfactory  for  treating  electrical 
conductivity  and  the  flow  of  electrons.  If  we  consider  electrons  in  a  box, 
there  can  be  no  net  flow  of  current,  for  as  many  electrons  will  be  traveling 
in  one  direction  as  in  the  opposite  direction,  on  account  of  reflection  at  the 
boundaries.  To  get  conductivity,  it  must  be  possible  for  more  electrons 
to  travel  one  way  than  the  other.  This  is  most  easily  handled  by  neg- 
lecting the  walls  of  the  box  and  by  assuming  that  an  electron  continues 
with  the  fixed  momentum  p,  as  in  the  derivation  of  Eq.  (5.7).  The 
corresponding  treatment  according  to  wave  mechanics  proves  to  be  to 
use  only  one  wave,  a  traveling  wave  like  Eq.  (5.1),  but  to  apply  the 


492  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

boundary  conditions  that  the  wave  must  reduce  to  the  same  phase  at 
x  =  0  and  x  =  X,  at  y  =  0  and  y  =  F,  and  at  z  =  0  and  z  =  Z.  This 
demands 


where  nx,  nu,  nz  are  integers,  which  reduces  to  the  conditions  (5.7).  As  in 
Eq.  (5.7),  the  integers  n  can  be  positive  or  negative,  so  that  the  corres- 
ponding particles  can  be  traveling  in  either  direction. 

We  can  now  consider  the  effect  of  a  periodic  potential  field  on  the 
electrons.  The  relation  of  waves  in  a  constant  potential  field  to  waves  in 
a  periodic  field  is  very  much  like  that  between  the  vibrations  of  a  continu- 
ous medium,  treated  in  Chap.  XIV,  and  vibrations  of  a  weighted  medium 
or  periodic  set  of  mass  points,  discussed  in  Chap.  XV.  The  calculation 
which  we  made  there  of  y2,  the  square  of  the  frequency,  is  similar  to  that 
which  is  made  here  of  the  energy  of  the  electron.  Thus,  there,  for  the 
continuous  medium,  we  had  Eq.  (2.18),  Chap.  XIV,  giving 


=  ©I(j)2n?;+^n>         (5.9) 


while  here  we  have 


P2        A2  [YnA2       /nA2       /n.V~| 

=  2^  =  2^L  w  +  w  +  w  J'        (5ilo) 


In  Chaps.  XIV  and  XV,  we  found  that  Eq.  (5.9)  did  not  really  hold  for  a 
solid  composed  of  atoms.  We  found  that  there  was  instead  a  periodic 
dependence  of  v2  on  sx,  sy,  and  sx,  as  illustrated  in  Fig.  XV-1.  This  arose 
because  of  a  periodic  dependence  of  the  wave  itself  on  the  s's,  as  illustrated 
in  Fig.  XV-2,  and  it  was  closely  tied  up  with  the  fact  that  the  shortest 
wave  that  could  be  propagated  in  the  crystal  had  a  half  wave  length  equal 
to  the  interatomic  distance,  so  that  this  wave  corresponded  to  opposite 
displacements  of  neighboring  atoms.  Similarly  in  the  wave  mechanics 
of  electrons  in  a  periodic  potential,  there  is  a  periodicity,  and  the  shortest 
wave  is  that  which  is  in  opposite  phases  at  successive  atoms.  We  cannot 
give  any  sort  of  derivation  of  the  behavior,  without  much  more  knowledge 
of  wave  mechanics  than  we  have  developed  here,  but  the  analogy  with  the 
mechanical  vibrations  is  correct  in  most  details.  In  place  of  the  reciprocal 
space  which  we  had  in  Chap.  XIV,  Sec.  2,  and  Chap.  XV,  Sec.  2,  we  have 
our  momentum  space,  each  wave  function  being  represented  by  a  point  in 
this  space.  The  momentum  space  is  divided  into  certain  polyhedra,  or 
Brillouin  zones,  such  that  the  energy  repeats  periodically  in  each  zone 
and  each  polyhedron  contains  as  many  stationary  states  as  there  are 
atoms  in  the  crystal.  There  is  one  difference  between  this  case  and  that 


SBC.  5] 


THE  ELECTRONIC  STRUCTURE  OF  METALS 


493 


of  elastic  vibrations,  however.  In  the  case  of  vibrations,  there  were  a 
number  of  independent  solutions  of  the  equation  giving  v2  as  a  function  of 
5*,  sy,  s,;  the  number  was  equal  to  the  number  of  atoms  in  the  unit  cell,  in 
the  simple  case  of  one-dimensional  vibration  which  we  took  up  there,  but 
in  general  it  equals  three  times  this  number,  when  we  consider  both 
longitudinal  and  transverse  vibration.  Here,  however,  the  relation 
giving  E  as  a  function  of  nx,  ny,  ng  has  an  infinite,  rather  than  a  finite, 


\ 


\ 


X 


(a) 


(b) 


(c) 

Fia.  XXIX-9. — Energy  as  a  function  of  momentum  px,  for  an  electron  in  a  one- 
dimensional  lattice,  (a)  Almost  free  electron.  (6)  Same  as  (a),  but  plotted  in  the  first 
Brillouin  zone,  (c)  More  tightly  bound  electron. 

number  of  solutions.  These  are  often  called  energy  bands.  As  in  Chap. 
XV,  Sec.  2,  we  can  handle  these  different  bands  in  cither  of  two  ways:  in 
the  first  place,  we  can  confine  our  attention  to  the  central  Brillouin  zone, 
plotting  E  as  a  function  of  the  n's  for  each  energy  band;  or  in  the  second 
place,  we  can  plot  one  band  in  one  zone,  another  in  another,  in  such  a 
way  that  they  fit  together  in  a  reasonable  way.  The  corresponding  two 
ways  of  plotting  the  elastic  spectrum  were  shown  in  Fig.  XV-4  (c). 

Now  we  are  prepared  to  understand  the  actual  behavior  of  the  elec- 
tronic energy  levels  in  a  crystal,  as  a  function  of  the  momentum.  In 
Fig.  XXIX-9  we  plot  curves  of  E  vs.  pxj  for  a  one-dimensional  model  of  a 


494 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 


crystal,  in  two  cases.  In  (a)  we  show  a  case  in  which  there  is  only  slight 
departure  from  constancy  of  the  potential,  so  that  we  have  almost  the 
free  electron  case.  For  free  electrons,  of  course  E  =  p2/2w,  so  that  the 
curve  would  be  a  parabola.  The  curve  in  (a)  is  plotted  so  as  to  show  its 
resemblance  to  a  parabola,  each  energy  band  being  plotted  in  a  different 
zone.  It  will  be  seen  that  for  most  energies  and  momenta  the  parabola  is 


FIG.  XXIX-10. — Energy  bands  as  a  function  of  internuclear  distance.  The  graph  is 
drawn  for  metallic  sodium,  showing  the  bands  that  go  into  the  2a,  3s,  and  3p  levels  of  the 
atom  at  infinite  separation.  The  2s  level  is  an  x-ray  level,  and  3«  is  the  valence  electron 
level.  The  energy  gap  which  appears  between  3s  and  3;>  in  the  figure,  at  distances  of  less 
than  6 A,  is  really  filled  with  bands  from  higher  atomic  levels. 

a  good  approximation,  but  there  is  a  discontinuity  of  energy  near  the 
momenta  at  the  edges  of  the  zones.  In  (6)  we  plot  the  same  curves 
reduced  to  the  central  Brillouin  zone.  The  case  (c)  shows  a  much  greater 
departure  from  free  electrons.  This  is  a  case  more  like  what  is  met  in  an 
actual  metal.  We  plot  this  only  in  the  central  zone;  it  is  so  far  from 
the  case  of  free  electrons  that  there  is  no  sense  trying  to  make  the  curves 
resemble  a  parabola  by  plotting  as  in  (a).  Here  there  are  a  number  of 
bands  of  very  low  energy,  with  almost  a  constant  energy  throughout  the 


SBC.  61  THE  ELECTRONIC  STRUCTURE  OF  METALS  495 

band.  These  correspond  to  the  inner,  x-ray  energy  levels  of  an  electron 
in  a  force  field  representing  a  single  atom,  and  as  we  should  expect  from 
the  fact  that  the  orbits  of  the  corresponding  electrons  do  not  overlap  at 
all,  these  energy  levels  are  in  almost  exact  agreement  with  those  of  iso- 
lated atoms.  The  higher  bands,  however,  are  fairly  broad;  that  is,  the 
energy  difference  between  the  top  and  bottom  of  a  band  is  considerable. 
These  correspond  to  the  valence  electron  levels  of  the  separated  atoms  and 
are  broadened  on  account  of  the  perturbation  of  the  valence  electron  of 
one  atom  when  it  overlaps  another. 

It  is  very  instructive  to  plot  the  energy  bands  in  another  way,  as  in 
Fig.  XXIX-10.  Here  we  show  the  top  and  bottom  of  each  band,  as  a 
function  of  internuclear  distance,  as  we  vary  the  size  of  the  crystal.  This 
shows  very  clearly  the  way  in  which  the  energy  levels  at  infinite  separation 
go  into  sharp  levels,  as  in  the  isolated  atoms.  On  the  other  hand,  as  the 
distance  is  decreased,  the  bands  broaden,  the  broadening  beginning  at 
about  the  interatomic  distance  where  the  orbits  in  question  begin  to 
overlap,  so  that  the  valence  electron  levels  are  broadened  at  the  normal 
distance  of  separation  in  the  metal,  while  the  x-ray  levels  are  not  but 
would  be  broadened  if  the  crystal  were  compressed  to  a  much  smaller 
lattice  spacing.  Fig.  XXIX-10  is  drawn  for  a  three-dimensional  lattice, 
rather  than  a  one-dimensional  one  as  in  Fig.  XXIX-9.  In  the  one- 
dimensional  case,  there  is  an  energy  gap  between  each  band  and  its 
neighboring  band,  a  gap  in  which  there  are  no  energy  levels.  On  the 
other  hand,  in  three  dimensions,  this  is  not  in  general  the  case.  With 
the  lower,  widely  separated  bands  there  are  gaps,  but  in  the  case  of  the 
valence  electrons  the  bands  overlap  and  the  gaps  are  filled  up. 

6.  Energy  Bands,  Conductors,  and  Insulators. — In  discussing  atomic 
structure  in  Chap.  XXI,  we  first  found  out  about  the  energy  levels  of  an 
electron  in  a  central  field,  and  then  we  built  up  a  model  of  the  atom  by 
assuming  that  the  electrons  of  the  atom  were  distributed  among  these 
energy  levels,  the  lower  ones  being  filled  with  two  electrons  each,  one  of 
each  spin,  and  the  upper  ones  being  empty.  Similarly  here  we  have  taken 
up  the  energy  levels  of  an  electron  in  a  periodic  potential  field,  and  next 
we  must  ask  what  levels  are  actually  occupied  in  the  metal.  We  have 
mentioned  that  each  of  the  energy  bands  consists  of  N  stationary  states, 
if  there  are  N  atoms  in  the  metal.  For  the  x-ray  levels,  these  stationary 
states  all  have  almost  exactly  the  same  energy,  while  for  the  higher  levels 
they  vary  considerably  in  energy.  In  a  crystal,  then,  each  band  can 
accomodate  2N  electrons,  N  of  each  spin,  or  two  electrons  per  atom,  one 
of  each  spin.  We  first  fill  up  the  x-ray  levels,  having  just  the  same 
number  of  electrons  per  atom  as  in  the  isolated  atoms  and  with  just  about 
the  same  energy.  The  remaining  electrons  go  into  the  valence  electron 
bands.  It  then  becomes  a  question  of  great  importance  how  many  such 


496  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

electrons  there  are,  and  how  much  they  fill  the  bands  up.  To  consider 
this  question,  let  us  consider  a  series  of  elements,  such  as  Ne,  Na,  Mg, 
etc.,  each  containing  one  more  electron  than  the  one  before.  If  we 
formed  a  crystal  of  Ne  atoms,  there  would  be  ten  electrons  per  atom,  just 
enough  to  fill  the  energy  bands  coming  from  the  atomic  Is,  2s,  and  2p 
electrons.  These  are  all  fairly  narrow  bands,  not  overlapping  with  others. 
Thus  all  the  electrons  of  a  neon  crystal  would  be  in  filled  energy  bands. 
With  sodium,  however,  there  is  one  more  electron  per  atom,  and  it  must 
go  into  a  valence  electron  band.  The  band  coming  from  the  atomic  3.s 
electron  is  broadened  a  good  deal  at  the  actual  distance  of  separation  of 
sodium  atoms,  and  in  fact  its  E  vs.  p  curve  is  a  good  deal  like  that  for  free 
electrons.  This  band  can  hold  two  electrons  per  atom,  but  only  one  is 
available,  so  that  it  is  only  half  full.  In  the  momentum  space,  if  the 
energy  is  approximately  proportional  to  p2,  as  with  free  electrons,  the 
occupied  levels  will  then  fill  an  approximately  spherical  volume  half  as 
large  as  a  Brillouin  zone,  or  just  the  same  size  that  we  should  have  for 
free  electrons  with  one  free  electron  per  atom.  In  magnesium,  with  two 
valence  electrons  per  atom,  we  might  think  at  first  sight  that  both  would 
go  into  the  3s  band,  filling  it.  But  actually  there  is  no  gap  between  this 
3s  band  and  the  one  coming  from  the  atomic  3p  levels.  Some  of  the 
levels  of  the  3s  band  lie  higher  than  the  lowest  ones  of  the  3p  band.  The 
two  valence  electrons  per  atom,  or  2N  for  the  crystal,  will  go  into  the  low- 
est 2N  states  of  the  combined  bands,  meaning  that  some  will  go  into  the 
3s  band,  some  into  the  3p,  neither  one  being  entirely  filled.  As  more  and 
more  electrons  are  added,  the  bands  fill  up  more  and  more,  but  for  the 
characteristically  metallic  elements  the  levels  all  overlap,  so  that  we  never 
have  the  situation  of  a  certain  number  of  filled  levels,  all  the  rest  being 
empty,  with  a  gap  between. 

It  can  now  be  shown  that  if  a  crystal  has  all  its  electrons  in  filled 
bands,  it  must  be  an  insulator;  conductivity  comes  essentially  from 
partly  filled  bands.  To  see  this,  we  need  to  know  the  effect  of  an  external 
field  on  the  energies  of  the  electrons,  in  a  periodic  field.  It  turns  out 
that,  just  as  in  Sec.  4,  the  external  field  brings  about  a  constant  rate  of 
change  of  momentum,  the  points  representing  the  various  electrons  drift- 
ing with  uniform  velocity  in  the  momentum  space.  The  relation  between 
momentum  and  energy,  however,  is  as  we  have  found  in  this  section. 
Thus  we  may  see,  for  example  in  Fig.  XXIX-11,  what  will  happen.  At 
the  instant  when  the  field  is  applied,  the  levels  indicated  by  shading  in  (a) 
are  assumed  to  be  occupied.  This  corresponds  to  a  whole  Brillouin 
zone,  or  a  whole  band,  being  filled.  As  time  goes  on,  the  momenta  all 
increase,  so  that  the  occupied  levels  have  shifted  along  to  those  shaded  in 
(6).  But  on  account  of  periodicity,  the  levels  that  have  been  filled  in 
going  from  (a)  to  (b)  are  exactly  equivalent  to  those  vacated,  so  that 


SBC.  6] 


THE  ELECTRONIC  STRUCTURE  OF  METALS 


497 


there  has  been  no  net  change  of  the  electrons  at  all,  and  no  resultant  elec- 
tric current  has  been  set  up.  Furthermore,  no  change  of  distribution 
within  this  band  can  be  brought  about  by  collisions,  for  thorn  are  no 
empty  levels  to  which  electrons  can  jump.  In  other  words,  the  electric 
field  has  no  effect  on  a  filled  band.  This  argument  is  not  exactly  correct; 
it  neglects  the  polarization  effect,  which  the  field  of  course  can  produce  on 
electrons  in  closed  shells.  But  it  is  correct  in  showing  the  lack  of  con- 
ductivity from  filled  bands.  In  a  partly  filled  band,  on  the  contrary,  the 
behavior  is  qualitatively,  though  not  quantitatively,  like  that  described  in 
Sec.  4.  Only  a  part  of  a  Briilouin  zone  is  filled  with  electrons,  so  that 


(a) 


(b) 


FIG.  XXIX-11. — Occupied  electronic  levels  in  Brillouin  zones,  one-dimensional  crystal 
(a)  At  time  of  application  of  external  field.  (6)  After  lapse  of  time,  illustrating  that  on 
account  of  the  periodicity  the  occupied  levels  really  do  not  change  with  time. 

when  the  filled  region  is  shifted  by  the  field,  the  levels  that  are  filled  are 
different  from  those  that  are  vacated,  and  there  is  a  net  change  in  the 
distribution,  resulting  in  a  current. 

Using  the  statements  just  made,  we  see  that  a  crystal  of  neon,  or 
other  material  having  only  filled  electron  bands,  will  be  an  insulator,  while 
a  metal,  having  partly  filled  bands,  will  be  a  conductor.  One  can  set  up 
energy  bands  for  chemical  compounds,  as  well  as  for  elements,  and  if  the 
compounds  are  held  together  by  homopolar  bonds,  it  turns  out  that  the 
energy  bands  are  such  that  certain  bands  are  entirely  filled,  the  others 
entirely  empty,  so  that  these  materials  are  insulators.  For  metals,  on 
the  other  hand,  even  though  the  conductivity  is  explained  essentially  as 
on  the  free  electron  theory,  the  wave  mechanical  picture  can  contribute 
several  important  points  to  the  theory.  In  the  first  place,  the  current 
produced  by  a  certain  change  of  momentum,  or  by  a  given  field  acting 
through  a  given  time,  is  not  the  same  as  in  the  free  electron  theory.  We 


498  INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 

can  see  this  from  the  fact  that  a  filled  band  of  electrons  produces  no  net 
current,  though  it  would  according  to  the  free  electron  theory.  As  a 
matter  of  fact,  as  the  number  of  electrons  in  a  band  increases,  the  current 
produced  by  a  given  field  at  first  increases  proportionally  to  the  number 
of  electrons,  as  if  they  were  free,  but  as  the  band  is  more  nearly  filled  the 
current  increases  less  rapidly  than  the  number  of  electrons,  then  reaches 
a  maximum,  and  decreases  to  zero  when  the  band  is  filled.  As  a  result  of 
this  effect,  the  current  produced  is  less  than  if  the  electrons  were  really 
free.  We  may  define  an  effective  number  of  free  electrons,  equal  to  the 
number  that  would  produce  the  same  current  as  the  electrons  actually 
present.  For  the  alkali  metals,  this  effective  number  of  free  electrons  is 
almost  exactly  equal  to  the  actual  number  of  valence  electrons,  but  for 
other  metals  it  is  much  less,  so  that  even  though  there  are  more  valence 
electrons  than  in  the  alkalies,  the  effective  number  of  free  electrons  is  less. 
A  second  point  in  which  the  exact  theory  affects  the  conductivity  is 
in  the  matter  of  the  collisions  of  the  electrons  with  the  atoms,  resulting  in 
the  time  of  relaxation  which  we  have  discussed  in  Sec.  4  in  connection 
with  the  resistance.  We  found  in  Eq.  (4.9)  that  the  specific  conductivity 
was  proportional  to  the  relaxation  time,  or  the  specific  resistance  propor- 
tional to  the  probability  of  collision.  Now  in  wave  mechanics  the  picture 
we  form  of  the  collision  of  an  electron  with  an  atom  is  the  scattering  of  the 
wave  representing  the  electron,  by  the  irregularity  of  potential  represent- 
ing the  atom.  But  we  really  have  scattering,  not  by  a  single  atom,  but 
by  a  whole  crystalline  arrangement  of  atoms.  If  these  are  regularly 
spaced,  the  electron  wave  will  not  be  scattered,  any  more  than  a  light 
wave  is  scattered  in  passing  through  a  crystal.  It  is  only  the  deviations 
from  homogeneity,  the  irregularities  in  density,  that  produce  scattering. 
Thus  if  the  lattice  is  perfectly  regular,  electrons  will  travel  through  it 
undeviated.  This  is  the  case  at  the  absolute  zero  of  temperature.  At 
higher  temperatures,  however,  there  will  be  fluctuations  of  density,  on 
account  of  the  temperature  vibrations  of  the  atoms.  The  amount  of 
scattering  of  the  wave,  or  the  probability  of  collision  of  the  electron  with 
atoms,  will  be  proportional  to  the  mean  square  deviation  of  density  from 
the  mean,  or  to  the  square  of  the  amplitude  of  atomic  vibration,  which 
in  turn  is  proportional  to  the  energy  of  the  vibrating  atoms,  or  to  the 
temperature.  Thus  we  expect  the  probability  of  collision  and  the  specific 
resistance  to  be  proportional  to  the  temperature.  Of  course,  it  is  a  well- 
known  experimental  fact  that  this  is  true,  and  this  simple  explanation  of 
the  temperature  variation  of  specific  resistance  is  one  of  the  most  impor- 
tant results  of  the  wave  mechanical  theory  of  conductivity.  More 
elaborate  methods  make  it  possible  to  estimate  the  magnitude  of  the 
scattering  and  of  the  resistance,  and  the  agreement  with  experiment  is 
good. 


Sac.  6]  THE  ELECTRONIC  STRUCTURE  OF  METALS  499 

Further  applications  of  the  theory  of  energy  bands  can  be  made  to 
the  interatomic  forces  in  metals  and  other  solids.  We  have  already 
spoken,  in  Sec.  2,  of  the  relation  of  the  free  electron  theory  to  the  equation 
of  state.  This  relation  can  be  made  much  more  accurate  and  quantitative 
by  means  of  the  energy  band  theory.  In  Fig.  XXIX-10  we  have  shown 
the  way  in  which  the  energy  bands  vary  with  interatomic  distance.  In 
an  approximate  way,  we  can  find  how  the  energy  of  the  crystal  varies  with 
lattice  spacing  by  adding  the  energies  of  the  various  electrons  it  it,  though 
further  corrections  must  be  made  to  get  accurate  results.  We  see  that 
most  atomic  energy  levels  widen  out  as  the  atoms  approach,  their  centers 
of  gravity  staying  roughly  constant,  but  then  rising  as  the  atoms  come 
very  close  together.  If,  then,  the  crystal  had  only  filled  bands  of  elec- 
trons, we  should  expect  the  energy  to  be  roughly  constant  for  large 
interatomic  distances  but  to  rise  for  smaller  distances.  That  is,  there 
would  be  no  attractions  between  atoms,  apart  from  Van  dcr  Waals  forces, 
which  are  neglected  in  this  treatment,  but  at  smaller  distances  there  would 
be  repulsions,  resulting  in  the  impenetrability  of  atoms.  This  is  what  we 
should  expect  in  an  inert  gas  like  neon,  for  instance.  But  if  we  consider 
a  crystal  like  sodium,  we  have  a  band  of  valence  electrons  only  half  full, 
the  bottom  half  being  occupied.  This  bottom  half  represents  a  band  of 
electrons  which  spreads  out  as  the  interatomic  distance  decreases  (as 
we  found  in  Sec.  2  to  be  the  case  with  free  electrons),  but  whose  center  of 
gravity  first  decreases,  before  its  final  increase  at  very  small  interatomic 
distance.  Thus  the  mean  energy  has  a  minimum;  this  is  what  is  responsi- 
ble for  the  binding  of  the  metallic  crystal.  Calculations  using  this  model 
for  the  alkali  metals  give  very  good  agreement  with  experiment.  It  is 
clear  from  this  example  that  the  essential  for  metallic  binding  is  a  band 
whose  lower  half  is  filled,  while  its  upper  half  is  empty.  The  broader  the 
band  is  and  the  more  electrons  it  holds,  the  tighter  the  binding.  Thus 
in  the  transition  groups  of  elements,  we  have  noted  in  Chap.  XXVII, 
Sec.  2,  that  the  binding  increases  in  strength  as  we  add  more  d  electrons, 
weakening  again  as  the  d  shell  is  filled.  These  d  electrons  have  a  band 
which  is  rather  narrow  but  which  is  capable  of  holding  ten  electrons.  The 
first  five  go  into  the  bottom  half  of  the  band,  contributing  to  the  binding, 
while  the  last  five  go  into  the  upper  half,  weakening  the  binding  again. 

We  have  just  stated  that,  though  the  d  band  in  the  transition  group 
elements  is  rather  narrow,  resulting  from  the  small  overlapping  of  the 
rather  small  d  orbits,  still  it  can  contain  ten  electrons.  This  means  that 
the  number  of  energy  levels  per  unit  range  of  energy  must  be  very  much 
greater  than  it  would  be  for  free  electrons.  This  has  an  interesting 
application  to  the  electronic  specific  heat  of  these  metals.  The  general 
expression  for  the  specific  heat  of  a  system  obeying  the  Fermi  statistics 
is  given  in  Eq.  (4.8),  Chapter  V, 


500 


INTRODUCTION  TO  CHEMICAL  PHYSICS      [CHAP.  XXIX 


=  (**£\  LVr, 


(6.1) 


where  dN/d,€  is  the  number  of  energy  levels  per  unit  range.  Since  this  is 
much  larger  for  the  d  shell  than  for  free  electrons,  we  infer  that  the  elec- 
tronic specific  heat  should  be  unusually  large  for  a  transition  metal.  This 
is  observed  experimentally,  as  we  have  already  shown  in  the  discussion  of 
Table  XXIX-2. 

The  binding  of  valence  crystals  can  also  be  explained  from  the  stand- 
point of  energy  bands.  In  Fig.  XXIX-12  we  show  energy  bands  for 
diamond,  a  typical  crystal  held  by  homopolar  bonds.  We  see  that  the 


FIG.  XXIX-12. — Energy  bands  in  diamond.     The  lower  band  is  occupied,  the  upper  one 
unoccupied,  illustrating  the  energy  gap  above  the  occupied  levels. 

bands  look  rather  different  from  what  they  do  in  a  metal,  in  that  they  are 
divided  into  an  upper  and  a  lower  half,  with  a  gap  between.  There  are 
enough  electrons  in  carbon  to  fill  the  bands  coming  from  the  Is  atomic 
level,  and  the  lower  band  coming  from  the  atomic  2s  and  2p.  Then  there 
is  a  wide  gap  between  the  top  of  the  filled  level  and  the  first  empty  level, 
showing  that  diamond  must  be  an  insulator.  But  now  we  notice  that  the 
filled  level  dips  down  sharply  as  the  interatomic  distance  decreases,  before 
commencing  its  rise.  Thus  diamond  is  strongly  bound,  as  we  know  from 
its  high  heat  of  vaporization  and  low  compressibility.  Similar  bands 
occur  in  other  valence  compounds,  though  they  have  not  been  extensively 
investigated. 

From  the  examples  considered,  we  see  that  the  effect  of  the  periodic 
potential  on  the  motion  of  the  electrons  is  essential  to  an  understanding 


Sac.  61  THE  ELECTRONIC  STRUCTURE  OF  METALS  501 

of  interatomic  or  intermolecular  binding  in  crystals,  as  well  as  of  the 
electrical  properties.  We  have  been  able  to  do  no  more  in  this  chapter 
than  give  a  suggestion  of  the  type  of  results  to  be  obtained.  To  go 
further,  one  must  make  a  great  deal  more  study  than  we  have  of  the 
principles  of  wave  mechanics.  The  same  remark  might,  in  fact,  be  made 
about  a  great  many  of  the  topics  taken  up  in  this  book.  A  well-trained 
chemical  physicist  should  be  an  expert  in  the  quantum  theory  and  wave 
mechanics,  as  well  as  in  thermodynamics  and  statistical  mechanics.  Our 
effort  in  this  book,  however,  has  been  to  show  that  one  can  really  get 
surprisingly  far  and  can  understand  nature  surprisingly  well,  with  rela- 
tively elementary  facts  about  the  fine  structure  of  matter  and  the  prin- 
ciples governing  its  behavior. 


PROBABLE  VALUES  OP  THE  GENERAL  PHYSICAL 

CONSTANTS 


Velocity  of  light 
Mechanical  equivalent 
of  heat 

Normal  atmosphere 
Gas  constant 


Ice  point 

Volume  of   1   mole  of 

perfect  gas,  n.t.p. 
Avogadro's  number 


Boltzmann's  constant       k 


Planck's  constant 
Electronic  charge 

Electron  volt 


e  =  2.998  X  1010  cm.  per  second 

1  cal.  =  4.185  X  107  ergs 

1  kg.-cal.  =  4.185  X  10l°  ergs 

1  atm.  =  1.013  X  106  dynes  square  centimeter 

R  =  8.314  X  107  ergs  per  degree  per  mole 

_  8.314  X  107 

""  4.185  X  107 

=  1.987  cal.  per  degree  per  mole 

—  .08205  1.  atm.  per  degree  per  mole  * 
Temp.  =  273.2°  abs. 

Volume  =  0.08205  X  273.2  =  22.41  1. 
NQ  =  6.03  X  1023  molecules  per  mole 
8.314  X  107 


6.03  X  1023 
=  1.379  X  10~lfl  erg  per  degree 
h  =  6.61  X  10"27  erg-sec. 
e  -  4.80  X  10~10  e.s.u. 


Mass  of  atom  of  unit 

atomic  weight 
Mass  of  electron 


^  1.60  X  10~12  X  6.03  X  1028 

4.185  X  1010 
=  23.05  kg.-cal.  per  mole 


Mass  -  1.66  X  10"24  gm. 
m  -  9.10  X  10^8gm. 

—  isW  X  mass  of  unit  atomic  weight 
Rydberg  energy  Rhc  =  2.17  X  10"11  erg 

=  13.56  electron  volt 
=  313  kg.-cal.  per  mole 
Radius  of  first  hydro- 
gen orbit  GO  =  0.53  A. 

The  values  for  the  large-scale  constants  are  taken  from  the  well-known  tabulation 
of  Birge,  Phys.  Rev.  Supplement ,  Vol.  1,  1929.  The  atomic  and  electronic  constants 
are  taken  or  computed  from  the  more  recent  values  tabulated  by  Dunnington,  Bull. 
Am.  Phys.  Soc.,  14  (1),  17  (1939). 


508 


SUGGESTED  REFERENCES 

The  reader  may  very  likely  want  to  refer  to  other  texts  dealing  with  the  same 
general  subject  as  the  present  one.  There  are  of  course  a  great  many  books,  both  old 
and  new,  treating  thermodynamics,  statistical  mechanics,  or  both.  We  may  mention 
first  the  classical  treatises,  L.  Boltzmann's  "  Vorlesungen  tiber  Gastheorie,"  reprinted 
by  Barth  in  1923,  and  Willard  Gibbs's  "Elementary  Principles  in  Statistical  Mechan- 
ics," reprinted  in  his  "Collected  Works"  by  Longmans,  Green  and  Company,  in 
1931.  Somewhat  more  recent  are  two  books  by  M.  Planck:  "Treatise  on  Thermo- 
dynamics," published  by  Longmans,  Green  and  Company,  and  "Heat  Radiation," 
published  by  P.  Blakiston's  Sons  and  Company,  the  latter  containing  many  of  the 
statistical  ideas  connected  with  the  introduction  of  the  theory  of  quanta.  Another 
standard  text,  dealing  principally  with  phase  equilibrium,  is  the  "Lehrbuch  der 
Thermodynamik,"  by  J.  D.  van  der  Waals.  "The  Dynamical  Theory  of  Gases,"  by 
J.  H.  Jeans,  Cambridge  University  Press,  is  a  standard  text  on  kinetic  theory;  more 
recent  texts  on  the  same  subject  are  "Kinetic  Theory  of  Gases,"  by  Loeb,  and  a  book 
by  the  same  title  by  E.  H.  Kennard,  both  published  by  McGraw-Hill  Book  Company, 
Inc.  "Kinetische  Theorie  der  Warme,"  by  K.  Herzfeld,  published  by  Vieweg,  com- 
bines the  principles  of  statistical  mechanics  with  application  to  matter,  in  somewhat 
the  same  way  as  the  present  text.  Among  more  recent  texts  of  thermodynamics  one 
may  mention  "Thermodynamics,"  by  E.  Fermi,  Prentice-Hall,  Inc.,  1937;  "Textbook 
of  Thermodynamics,"  by  P.  Epstein,  John  Wiley  &  Sons,  Inc.,  1937;  "Heat  and 
Thermodynamics,"  by  J.  K.  Roberts,  Blackie  and  Son,  Ltd.,  1933;  "Modern  Thermo- 
dynamics by  the  Method  of  Willard  Gibbs,"  by  E.  A.  Guggenheim,  Methuen  and 
Company,  Ltd.,  1933;  and  "Heat  and  Thermodynamics,"  by  M.  W.  Zemansky, 
McGraw-Hill  Book  Company,  Inc.,  1937.  Recent  texts  on  statistical  mechanics 
include  the  standard  treatise,  "Statistical  Mechanics,"  by  R.  H.  Fowler,  Cambridge 
University  Press,  1929;  (a  revision  and  condensation  of  this  work,  in  collaboration 
with  Guggenheim,  is  understood  to  be  in  preparation);  "Statistical  Mechanics  with 
Applications  to  Physics  and  Chemistry,"  by  R.  C.  Tolman,  Chemical  Catalog  Com- 
pany, 1927;  the  successor  to  that  volume,  "Principles  of  Statistical  Mechanics,"  by 
Tolman,  Oxford  University  Press,  1938;  "Statistical  Physics,"  by  Landau  and  Lif- 
schitz,  Oxford  University  Press,  1938;  and  "  Quantenstatistik  und  ihre  Anwendungen 
auf  die  Elektronentheorie  der  Metalle,"  by  L.  Brillouin,  Springer,  1931.  A  number 
of  texts  on  physical  chemistry  and  chemical  thermodynamics  bear  closely  on  the  topic 
of  this  book.  First  of  course  is  the  well-known  "Thermodynamics  and  the  Free 
Energy  of  Chemical  Substances,"  by  Lewis  and  Randall,  McGraw-Hill  Book  Com- 
pany, Inc.,  1923.  Others  are  "A  Treatise  on  Physical  Chemistry,"  by  H.  S.  Taylor, 
D.  Van  Nostrand  Company,  Inc.,  1932;  and  "A  System  of  Physical  Chemistry,"  by 
W.  C.  McC.  Lewis,  Longmans  Green  &  Company,  1921.  Closely  related  is  "The  New 
Heat  Theorem,  its  Foundation  in  Theory  and  Experiment,"  by  Nernst,  E.  P.  Dutton 
Company,  Inc.,  1926,  describing  some  of  the  early  applications  of  quantum  theory  to 
thermodynamics. 

Thermodynamics  and  statistical  mechanics,  as  well  as  the  structure  of  matter,  can 
hardly  be  understood  without  a  study  of  atomic  and  molecular  structure,  and  of  the 
quantum  theory  which  underlies  them.  Suggested  texts  in  this  general  field  are 

505 


506  INTRODUCTION  TO  CHEMICAL  PHYSICS 

"Introduction  to  Modern  Physics,"  by  F.  K.  Richtmyer,  McGraw-Hill  Book  Com- 
pany, Inc.,  "Atoms,  Molecules  and  Quanta,"  by  Ruark  and  Urey,  McGraw-Hill  Book 
Company,  Inc.,  "Introduction  to  Atomic  Spectra,"  by  H.  E.  White,  McGraw-Hill 
Book  Company,  Inc.,  "The  Structure  of  Line  Spectra,"  by  Pauling  and  Goudsmit, 
McGraw-Hill  Book  Company,  Inc.  For  somewhat  more  mathematical  treatments 
of  the  same  field,  including  wave  mechanics,  useful  treatments  are  found  in  "Introduc- 
tion to  Quantum  Mechanics,"  by  Pauling  and  Wilson,  McGraw-Hill  Book  Company, 
Inc.,  "Elements  of  Quantum  Mechanics,"  by  S.  Dushman,  John  Wiley  &  Sons,  Inc., 
"Wave  Mechanics,  Elementary  Theory,"  by  J.  Frenkel,  Oxford  University  Press; 
"The  Fundamental  Principles  of  Quantum  Mechanics,"  by  E.  C.  Kemble,  McGraw- 
Hill  Book  Company,  Inc.;  as  well  as  the  treatment  in  "Introduction  to  Theoretical 
Physics,"  by  Slater  and  Frank,  McGraw-Hill  Book  Company,  Inc. 

Many  texts  deal  with  specific  subjects  taken  up  in  one  chapter  or  another  of  the 
present  book.  One  may  mention  "The  Phvsics  of  High  Pressure."  by  P.  W.  Bridg- 
man,  The  Macmillan  Company,  1931,  and  "The  Thermodynamics  of  Electrical 
Phenomena  in  Metals,"  also  by  Bridgman  and  published  by  The  Macmillan  Com- 
pany; "Metallography,"  by  C.  H.  Desch,  Longmans,  Green  and  Company,  for  a 
treatment  of  phase  equilibrium;  "Measurement  of  Radiant  Energy,"  by  W.  E. 
Forsythe,  McGraw-Hill  Book  Company,  Inc.,  for  black-body  radiation;  "Crystal 
Chemistry,"  by  C.  W.  Stillwell,  McGraw-Hill  Book  Company,  Inc.,  with  more 
detailed  treatments  of  many  of  the  points  taken  up  in  the  Part  III  of  the  present 
book;  "Valence  and  the  Structure  of  Atoms  and  Molecules,"  by  G.  N.  Lewis,  Chemical 
Catalog  Company,  and  "The  Nature  of  the  Chemical  Bond,"  by  L.  Pauling,  Cornell 
University,  1939,  for  discussion  of  valence  bonds;  "The  Crystalline  State,"  by  W.  H. 
and  W.  L.  Bragg,  The  Macmillan  Company,  1934,  and  "The  Structure  of  Crystals," 
by  R.  W.  G.  Wyckoff,  Chemical  Catalog  Company,  for  more  detailed  information 
about  crystal  structure;  "Photoelectric  Phenomena,"  by  Hughes  and  DuBridge, 
McGraw-Hill  Book  Company,  Inc.,  for  treatment  of  electronic  questions;  "Properties 
of  Metals  and  Alloys,"  by  Mott  and  Jones,  Oxford  University  Press,  "Elektronen- 
theorie  der  Metalle,"  by  H.  Frohlich,  Springer,  and  "The  Theory  of  Metals,"  by  A.  H. 
Wilson,  Cambridge  University  Press,  for  the  structure  of  metals. 

In  addition  to  these  texts,  various  handbooks,  tables,  etc.,  are  of  great  service. 
Both  the  "Handbuch  der  Physik"  and  the  "Handbuch  der  Experimentalphysik " 
contain  several  volumes  dealing  with  the  general  subjects  treated  in  the  present 
book,  treated  both  experimentally  and  theoretically.  Some  specific  references  are 
made  to  them  in  the  text.  For  numerical  data,  Landolt-Bornstein's  "Physikalisch- 
Chemische  Tabellen"  are  invaluable,  supplemented  by  the  "International  Critical 
Tables."  In  the  field  of  atomic  spectra,  "Atomic  Energy  States,"  by  Bacher  and 
Goudsmit,  McGraw-Hill  Book  Company,  Inc.,  has  very  complete  data,  and  the  field 
of  crystal  structure  is  covered  by  the  "Strukturbericht,"  a  supplement  to  the  "Zeit- 
schrift  fur  Kristallographie."  Band  spectra  and  molecular  structure  are  included  in 
the  volume  of  tables  in  "  Molekulspektren  und  ihre  Anwendungen  auf  Chemische 
Probleme,"  by  H.  Sponer,  Springer.  Finally,  though  we  have  not  listed  many  refer- 
ences to  original  articles,  the  reader  will  do  well  to  consult  review  articles  in  Chemical 
Reviews  and  Reviews  of  Modern  Physics,  as  well  of  course  as  becoming  familiar  with  the 
large  literature  in  the  Journal  of  Chemical  Physics,  Journal  of  the  American  Chemical 
Society,  as  well  as  foreign  periodicals.1 


INDEX 


A  priori  probability,  36,  38,  127 
Absorption  of  radiation,  309-310, 322-333 
Absorptivity,  309-310,  325-326 
Acetic  acid,  data  regarding  melting  point, 
259 

heat  of  vaporization,  434 

structure  of  molecule,  427 

Van  der  Waals  constants,  408 
Acetylene,  structure  of  molecule,  428 

Van  der  Waals  constants,  408 
Activation,  energy  of,  159-164,  257 
Adiabatic  processes,  13,  17-19 
Aliphatic  compounds,  420-428 
Alkali  halides,  data  regarding  crystals, 
381,  393,  395 

data  regarding  melting  point,  259 

equation  of  state,  390-396 

residual  rays,  254-255 

specific  heat,  253 
Alloys,  270-290,  458-459 
Aluminum,  crystal  structure,  447 

data  regarding  melting  point,  259,  261 

Debye  temperature,  237 

equation  of  state  and  crystal  structure, 
451,  454 

molecular  volume,  261 

specific  heat,  236 

thermal  expansion,  261 
Ammonia,  data  regarding  melting  point, 
259 

formation     of    ammonium     complex, 
272-273 

heat  of  vaporization,  414 

valence  structure  of  molecule,  401 

Van  der  Waals  constants,  408 
Ammonium    bromide,    data    regarding 
crystals,  382 

rotation  of  ions,  300 

Ammonium     chloride,     data    regarding 
crystals,  381 

rotation  of  ions,  293,  300 
Ammonium  iodide,  data  regarding  crys- 
tals, 382 


Ammonium  ion,  357,  378 

Angular   momentum,    quantization,    40, 

135,  339 
Annealing,  282 
Antimony,  crystal  structure,  447,  449 

data  regarding  melting  point,  259 
Argon,  atomic  volume,  384 

data  regarding  crystals,  416 

data  regarding  melting  point,  259 

specific  heat,  130 

Van  der  Waals  constants,  408 
Aromatic  compounds,  428-432 
Arsenic,  crystal  structure,  447,  449 
Assembly,  canonical,  46—53 
fluctuations  in,  101-111 

microcanonical,  46 

in  molecular  phase  space,  69-72 

statistical,  32-35 
Atomic  number,  336 
Atoms,  structure,  321-351 
Attractive    forces    between    molecules, 
130-133 

and  atomic  theory,  352-376 

in  solutions,  271-277 

and  Van  der  Waals'  equation,  182-184, 

194-196 

Avogadro's  law,  60 
Avogadro's  number,  60-61,  128 

B 

Barium,  crystal  structure,  447 

equation  of  state,  451 
Barium  chloride,  data  regarding  melting 

point,  259 
Barium  oxide,  sulphate,  selenide,  tellu- 

ride,  data  regarding  crystals,  381 
Barometer  formula,  62-64 
Benzene,  data  regarding  melting  point, 

259 

structure  of  molecule,  428-431 

Van  der  Waals  constants,  408 

Beryllium,  crystal  structure,  447 

equation  of  state,  451 
Beryllium  oxide,  sulphide,  selenide,  tellu- 
ride,  data  regarding  crystals,  382 


507 


508 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Binary  systems,  phase  equilibrium,  270- 

290 

Bismuth,  crystal  structure,  447-449 
data  regarding  melting  point,  259 
Black-body  radiation,  307-320,  325-326 

and  Einstein-Bose  statistics,  85 
Body-centered     cubic     structure,     and 

metals,  445-447 
and  molecular  vibrations,  232 
and  order-disorder,  293 
Bohr,  frequency  condition,  323 

hydrogen  atom,  340 
Boiling  point,  166-169,  171-180 

of  chain  compounds,  422 
Boltzmann  factor,  and  thermionic  emis- 
sion, 468 

Boltzmann  statistics,  application  to  per- 
fect gas,  115-129 
fluctuations,  101-104,  109 
and  kinetic  method,  86-96 
relation  to  Fermi-Dirac  and  Einstein- 
Bose  statistics,  68-72 
Boltzmann's  constant,  33,  61 
Boltzmann's  H  theorem,  90 
Boltzmann's  relation  between  probability 

and  entropy,  34 
Boundary  conditions,  vibrating  chain  of 

atoms,  244 

vibrating  solid,  227-228 
Boyle's  law,  30,  60 

deviations  from,  190-198 
Brass,  phase  equilibrium,  270,  287-288 
Bridgman,  24,  181,  201,  220 
Brillouin  zone,  and  electrons  in  metals, 

493-501 

and  molecular  vibrations,  233 
Bromine,  characteristic  temperature,  for 

rotation,  136 
for  vibration,  142 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
and  homopolar  bonds,  400-408 
and  organic  compounds,  425-426 
Butane,  heat  of  vaporization,  434 
structure  of  molecule,  421 
Van  der  Waals  constants,  408 


Cadmium,  crystal  structure,  447 
data  regarding  melting  point,  259,  261 
Debye  temperature,  237 
equation  of  state  and  energy,  451,  454 


Cadmium,  molecular  volume,  267 

thermal  expansion,  261 
Cadmium  oxide,  sulphide,  selenide,  tel- 

luride,  data  regarding  crystals,  382 
Caesium,  compressibility,  202 
crystal  structure,  447 
data  regarding  melting  point,  259 
equation  of  state  and  energy,  451,  454 
Caesium  bromide,  fluoride,  iodide,  data 

regarding  crystals,  381 
Caesium  chloride,  crystal  structure,  378- 

379,  381 

Calcite  stnicture,  397-398 
Calcium,  crystal  structure,  447 

equation  of  state  and  energy,  451,  454 
phase  equilibrium  in  alloys,  274 
Calcium  carbonate,  structure,  397-398 
Calcium  chloride,  data  regarding  melting 

point,  259 
Calcium   oxide,   sulphide,   selenide,   tel- 

luride,  data  regarding  crystals,  381 
Caloric  theory  of  heat,  4-6 
Calorie,  6 

numerical  value,  8 
Canonical  assembly,  46-51 

and   Maxwell-Boltzmann  distribution, 

52-53 

Carbon,  and  homopolar  bonds,  400-407 
and  organic  compounds,  420-434 
(diamond),  crystal  structure,  379,  424 
Debye  temperature,  237 
energy  bands,  500 
energy  constants,  455 
(graphite),  crystal  structure,  429 
(molecule),  characteristic  temperature, 

of  rotation,  136 
of  vibration,  142 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
Carbon  bisulphide,  Van  der  Waals  con- 
stant, 408 
Carbon  dioxide,  data  regarding  melting 

point,  259 

heat  of  vaporization,  414 
triple  point,  167 

valence  structure  of  molecule,  405 
Van  der  Waals  constants,  408,  411 
vapor  pressure  and  latent  heat  of 

vaporization,  188-189 
Carbon    monoxide,    characteristic    tern* 

perature,  for  rotation,  136 
for  vibration,  142 


INDEX 


509 


Carbon  monoxide,  crystal  structure,  417 
data  regarding  melting  point,  259 
dissociation,  133 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
heat  of  vaporization,  414 
Van  der  Waals  constants,  408,  411 
vibrational  specific  heat,  144—145 
Carbon  tetrachloride,  boiling  point,  426 

data  regarding  melting  point,  259 
Carbonate  ion,  valence  structure,  406 
Carboxyl  group,  427 
Cells  in  phase  space,  38-43 
and    Fermi-Dirac    and    Einstein-Bose 

statistics,  65-85 
and  interatomic  forces,  369 
and  kinetic  method,  86-100 
CH,  characteristic  temperature,  for  rota- 
tion, 136 

for  vibration,  142 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
Change  of  phase,  23,  166-190,  2S&-269 
Characteristic   temperature,  for  atomic 
vibrations  in  crystals,  Debye,  235- 
236 

for  ionization,  322,  335 
for  rotation,  136 
for  vibration,  diatomic  molecules,  141- 

142 

Charles's  law,  60 

Chemical  constant,  and  chemical  equilib- 
rium, 155-156 
of  diatomic  gas,  140 
of  monatomic  gas,  118,  120,  128 
and  thermionic  emission,  463-464 
and  vapor  pressure,  178-180 
Chemical  equilibrium,  150-165 
Chlorine,  characteristic  temperature,  for 

rotation,  136 
for  vibration,  142 
data  regarding  melting  point,  259 
heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
and  homopolar  bond,  400-408 
and  organic  compounds,  425-426 
Van  der  Waals  constants,  408 
Chloroform,  boiling  point,  426 
heat  of  vaporization,  434 
Van  der  Waals  constants,  408 
Chromium,  crystal  structure,  447 
data  regarding  melting  point,  259 


Chromium,  equation  of  state  and  crystal 

structure,  451,  454 
Clapeyron's  equation,  174-181,  220 
Clay,  structure,  440 
Cobalt,  crystal  structure,  447 

data  regarding  melting  point,  259 

equation  of  state  and  energy,  451,  454 
Cold-working,  457 

Collisions,  and  chemical  reactions,   150- 
154,  158-165 

effect  on  approach  to  equilibrium,  86- 
92,  96-100 

elastic,  327 

and  electrical  resistance,  487-489 

of  electron  and  metal,  460-467 

inelastic,  327 

and  radiation,  326-333 

of  second  kind,  331 
Complexion,  33-43 

and    Fermi-Dirac   and    Einstein-Bose 

statistics,  65-85 

Compressibility,  of  alkali  halides,  392- 
395 

and  entropy  of  melting,  261 

and  fluctuations,  110 

of  metals,  452-454 

of  solids,  200-205,  218-221 

and  thermal  expansion,  238-240 

and  thermodynamics,  19 

of  water  and  ice,  170 
Concentration,  and  chemical  equilibrium, 
153-154 

and  mixtures  of  gases,  121-124 

and  phase  equilibrium  in  binary  sys- 
tems, 270-290 
Conductivity,    electrical,   456,    484-489, 

495-501 
Configuration,  atomic,  342 

table  of,  346-347 
Conservation  of  energy,  3-9 
Conservative  force,  3 
Contact  difference  of  potential,  467-471, 

480-482 
Copper,  alloys  with  nickel,  458 

crystal  structure,  447 

data  regarding  melting  point,  259 

Debye  temperature,  237 

energies  of  electrons  in  atom,  341 

equation  of  state  and  energy,  451,  454 

order-disorder  in  alloys,  293-304 

phase  equilibrium  in  alloys,  270,  279- 
282,  287-288 


510 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Corresponding    states,    and    Van    der 

Waals'  equation,  187 
Coulomb  energy,  353,  361-363,  368-369 
Covalent  bond,  373-376,  400-407 

and  organic  compounds,  420-434 

and  silicates,  435-443 
Cristobalite  structure,  441 
Critical  point,  166-169 

and  phase  changes  of  second  order,  291 

and  Van  der  Waals  equation,  184-187 
Crystals,  electric  field  in,  473,  489-501 

equation  of  state,  211-221 

ionic,  375,  377-399 

and  liquid,  256-258 

of  metals,  444-451 

molecular,  414-419 
Cuprous  chloride,  bromide,  iodide,  data 

regarding  crystals,  382 
Curie  point,  292,  297-304 
Cyclohexane,  structure  of  molecule,  424 


D 


Pebye,  ionic  crystals,  390-393 

specific  heat  of  solids,  222-240,  245, 

253-255 
Decanc,  heat  of  vaporization,  434 

Van  der  Waals  constants,  408 
Degeneracy,  in  space  quantization,  139, 

339 

Degradation  of  energy,  12 
Degree  of  order,  294-304 
Dependent  variables,  17-18 
Detailed  balancing,  91 

and  radiation,  324 

Diamond,    crystal   structure,   379,    424, 
444-447 

energy  bands,  500 

melting  point,  449 
Diatomic  molecules,  130-145,  400-414 

comparison  with  metals,  455 
Dichlorbenzene,   structure  of  molecule, 

431 
Dielectric  constant,  365 

and  Van  der  Waals  forces,  411 
Diffusion,  12-13 

Dimethylamine,  structure  of  molecule, 
427 

Van  der  Waals  constants,  408 
Dipoles,  354-361 

moment  of  HC1  and  HBr,  404-405 

moments  of  molecules,  table,  408 


Disorder,  10-12,  32-38,  43-46 

in  alloys,  293-304 

Dispersion  of  elastic  waves,  223,  234,  244 
Dissociation,  152-154 
Distribution  function,  33-35,  46-51 

in     Fermi-Dirac    and     Einstein-Bose 
statistics,  65-85 

for  fluctuations,  104-107 

in   Max  well-Bo  tzmann  statistics,  52- 

64 

Double  bonds,  401-402,  428,  431 
Ductility,  456 
Dulong  and  Petit's  law,  213 

deviations  from,  222-240 


E 


Effective  volume  of  molecules,  and  Van 

der  Waals'  equation,  183,  195-196 
Einstein,  black-body  radiation,  324-326 
photoelectric  effect,  316-320 
specific  heat  functions,  for  gases,  144, 

147 

for  solids,  214r-215,  253 
Einstein-Bose  statistics,  52,  65-85 
and  black-body  radiation,  326 
and  fluctuations,  108-109 
and  kinetic  method,  96-100 
and  perfect  gas,  126 
Elastic  collisions,  327 
Elastic  vibrations  of  solids,  and  specific 

heat,  222-255 
Electron  affinity,  338 
Electron  gas,  and  Fermi-Dirac  statistics, 

81 

and  metallic  structure,  475-484 
Electron  volt,  numerical  value,  132—133, 

318 
Electrons  and  atomic  structure,  337-351 

and  structure  of  metals,  472-501 
Electrostatics,  and  field  in  metal,  472- 

489 

and  interatomic  forces,  353-367 
and  ionic  crystals,  385-390 
Emission  of  radiation,  309-310,  317-320, 

322-333 

Emissive  power,  309-310,  325-326 
Energy    (see    Conservation    of   energy; 
Internal    energy;    Kinetic    energy; 
Potential  energy) 

Energy  of  activation,  159-164,  257 
Energy  bands  in  metals,  493-501 


INDEX 


511 


Energy  density,  in  radiation,  310-316, 

324-326 
Energy  levels,  41-42 

of  atomic  systems,  322,  338-344 
Ensemble  (see  Assembly) 
Enthalpy,    and    Joule-Thomson    effect, 

197-198 

and  latent  heats,  175-178 
and  thermodynamics,  20-21 
Entropy,  of  diatomic  gas,  140 
and  equilibrium  of  phases,  170-173 
in     Fermi-Dirac     and     Einstein-Bose 

statistics,  69-72 
and  fluctuations,  107 
of  fusion,  171-180,  258-269 
and  kinetic  method,  89-91,  98-99 
of  mixture  of  gases,  121-123,  128-129 
of  perfect  gas,  117-119,  127 

Fermi-Dirac  statistics,  78-79 
and   phase   change   of   second   order, 

291-304 

and  phase  equilibrium  in  binary  sys- 
tems, 272-290 
of  solids,  207-218 

and  statistical  mechanics,  32-35,  43-51 
and  thermodynamics,  9-14,  17-18,  21 
of  vaporization,  171-180 
Equation  of  state,   of  imperfect  gases, 

182-198 

of  ionic  crystals,  385-396 
of  metals,  450-456,  479-480 
of  perfect  gas,  58-61 

Fermi-Dirac  statistics,  82 
and  phase  equilibrium,  169-170 
of  solids,  199-221 
and   thermodynamics,    16-18,    22-23, 

2&-30 

Equilibrium,   between  atoms  and   elec- 
trons, 333-335 

between  liquid  and  solid,  266-269 
between  metals  and  electron  gas,  463- 

464 
between  phases,  166-181,  184-190 

in  binary  systems,  270-290 
chemical,  150-165 
and  radiation,  325 

thermal,  13-15,  23, 37-38,  46-51,  96,  98 
Equipartition  of  energy,  and  Maxwell's 

distribution,  57-58 
and  specific  heat,  of  polyatomic  gases, 

134,  144,  146 
of  solids,  213 


Ethane,  heat  of  vaporization,  434 
hindered  rotation,  147-148 
structure  of  molecule,  402,  420-421 
Van  der  Waals  constants,  408 

Ethyl   alcohol,   data   regarding  melting 

point,  259 

heat  of  vaporization,  434 
Van  der  Waals  constants,  408 

Ethyl  ether,  heat  of  vaporization,  434 
structure  of  molecule,  427 
Van  der  Waals  constants,  408 

P^thylene,  structure  of  molecule,  402,  428 
Van  der  Waals  constants,  408,  411 

Eutectic,  284-285 

Exchange,  and  interatomic  forces,  367- 
374 

Excitation  of  atoms,  321-333,  343 

Exclusion  principle,  342 

and  interatomic  forces,  369-372 

Explosion,  158-159 

External  work,  3,  7-9,  17,  21-22 
and  statistical  mechanics,  49 


Face-centered  cubic  structure,   descrip- 
tion and  figure,  415 

in  inert  gases,  416 

and  metals,  445-447 

and  molecular  vibrations,  232 

and  order-disorder,  293 
Fermi-Dirac  statistics,  52,  65-85 

and  atomic  structure,  342 

and  exchange  effect,  369 

and  fluctuations,  108-109 

and  kinetic  method,  96-100 

and  metals,  471,  475-484 

and  perfect  gas,  126 
Ferromagnetism,  292-293 
Fibers,  silicate,  439 

Field,   electric,   and   interatomic   forces, 
35&-360,  366 

in  metal,  472-501 
First  law  of  thermodynamics,  7-8,  19 

and  statistics,  49-51 
Fluctuations,  32.  101-111 
Fluorite  structure,  396-397 
Forces  between  molecules,  130-133 

interpretation    from    atomic    theory, 
352-376 

in  solids,  271-277 

and  Van  der  Waals'  equation,  182-184, 
194-196 


512 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Formic  acid,  structure  of  molecule,  427 
Free  electrons  in  metals,  475-489 
Free  energy,  Gibbs,  and  chemical  equili- 
brium, 154-168 
of  diatomic  gas,  140 
and  equilibrium  of  phases,  170-180 
and  melting,  265-269 
of  mixture  of  gases,  123-124 
of  perfect  gas,  120 
and  phase  changes  of  second  order, 

296-304 
and    phase    equilibrium    in    binary 

systems,  270,  278-290 
of  solids,  205-211 
and  thermionic  emission,  463-464 
and  thermodynamics,  22-23 
and  Van  der  Waals'  equation,  184- 

189 
Helmholtz,  and  Fermi-Dirac  and  Ein- 

stein-Bose  statistics,  73,  79,  82 
and  melting,  265-269 
of  perfect  gas,  119,  126 
and  second  virial  coefficient,  193-194 
of  solids,  205-211,  216-218 
and  statistical  mechanics,  50-51 
and  thermodynamics,  21-22 
Free  expansion  of  gas,  30,  196-198 
Freezing  (see  Melting) 
Frequency  of  oscillation,  diatomic  mole- 
cule, 141 

molecular  solid,  241-255 
solid,  213-240 
Friction,  3 
Fusion,  23,  166-169,  171-176,  256-269 


G 


Gallium,  crystal  structure,  447 

data  regarding  melting  point,  259 
Gas  constant,  33 

numerical  values,  60 
Gases,  and  equilibrium  with  other  phases, 

166-180 

imperfect,  and  Van  der  Waals'  equa- 
tion, 182-198 
perfect,  17,  30 

chemical  equilibrium  in,  150-165 
and    Maxwell-Boltzmann    distribu- 
tion, 53-64 
polyatomic,  130-149 
thermodynamic  and  statistical  treat- 
ment, 115-129 


Gases,  perfect,  translational  energy  levels 
in  quantum  theory,  54-55 

Van  der  Waals  constants,  408 
Gauss  error  curve,  106 
Germanium,  crystal  structure,  444,  447- 
449 

melting  point,  449 
Gibbs,  32,  44,  107 

Gibbs  free  energy  (see  Free  energy,  Gibbs) 
Gibbs's  paradox,  129 
Glass,  256-258 

structure,  442 

variability  of  composition,  273 
Gliding,  457 
Gold,  crystal  structure,  447 

data  regarding  melting  point,  259 

Debye  temperature,  237 

equation  of  state  and  energy,  451,  454 

order-disorder  in  alloys,  293-294 
Graphite  structure,  429 
Gravity,  4 
Grtineisen,  thermal  expansion,  217-221, 

238-240 

of  ionic  crystals,  392-394 
of  metals,  451-456 

H 

H  theorem,  90 

Hafnium,  crystal  structure,  447 

equation  of  state,  451 
Halogens,  characteristic  temperature,  for 

rotation,  136 
for  vibration,  142 
data  regarding  melting  point,  259 
heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
and  homopolar  bonds,  400-408 
and  organic  compounds,  425-426 
Heat  absorption,  7-9,  12-13,  20 
and  statistical  mechanics,  49 
Heat,    of    dissociation,    diatomic    mole- 
cules, table,  132 
latent     (see   Latent   heat,    of   fusion; 

Latent  heat,  of  vaporization) 
of  reaction,  156-158 

and  equilibrium  of  ions  and  electrons, 

334 

specific  (see  Specific  heat) 
Heat  capacity  (see  Specific  heat) 
Heat  engine,  13,  172 
Heat  flow,  12-13 
Heitler-London  method,  367-368 


INDEX 


513 


Helium,  specific  heat,  130 

Van  der  Waals  constants,  408 
Helmholtz  free  energy  (see  Free  energy, 

Helmholtz) 

Hexagonal   close-packed   structure,    de- 
scription and  figure,  417 
and  metals,  445-447 
and  molecular  vibration,  232 
Hexane,  structure  of  molecule,  423 
Hindered  rotation,  147-149,  417-418 
Homopolar  valence  attraction,  373-376, 

400-407 

and  organic  compounds,  420-434 
and  silicates,  435-443 
Hydrogen,     characteristic    temperature, 

for  rotation,  136 
for  vibration,  142 
combination    with    oxygen    to    form 

water,  151-164 

data  regarding  melting  point,  259 
heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
heat  of  vaporization,  414 
and  homopolar  bonds,  400-408 
interatomic  potential,  371 
and  organic  compounds,  420-434 
specific  heat,  137-138 
Van  der  Waals  constants,   408,   411 
Hydrogen  bromide,  data  regarding  melt- 
ing point,  259 
valence  structure,  404-405 
Van  der  Waals  constants,  408 
Hydrogen   chloride,   characteristic  tem- 
perature, for  rotation,  136 
for  vibration,  142 

crystal  structure  and  hindered  rota- 
tion, 417 

data  regarding  melting  point,  259 
dipole  moment,  358 
heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
heat  of  vaporization,  414 
valence  structure,  404-405 
Van  der  Waals  constants,  408 
Hydrogen    sulphide,    valence    structure, 

405 
Van  der  Waals  constants,  408 


Ice,  crystal  structure,  412,  418-419 
polymorphic  forms,  167-170 
structure,  260 


Image  force,  461,  474-475 
Impenetrability  of  matter,  130 
Imperfect  gases,  and  phase  equilibrium, 

166-170 

and  Van  der  Waals'  equation,  182-198 
Independent  variables,  17-18 
Indium,  crystal  structure,  447 

data  regarding  melting  point,  259 
Induced  emission,  325 
Inelastic  collisions,  327 
Inert  gases,  crystals,  416 
and  periodic  table,  345-350 
volumes  of  atoms,  384 
Insulators,  and  energy  bands,  495—501 
Integrals  independent  of  path,  8,  13 
Interatomic  distances,  in  crystals,  and 
formulas  for  thermodynamic  quan- 
tities, 212-213 

in  crystals  of  inert  gases,  416 
in  diatomic  molecules,  table,  132 
in  ionic*  crystals,  table,  381-382 
in  metals,  447 

in  organic  compounds,  420-434 
Interatomic  forces,  130-133 
interpretation    from    atomic    theory, 

352-376 

in  ionic  crystals,  385-390 
in  metals,  451-456 
in  organic  compounds,  433 
and  second  virial  coefficient,  191-196 
and  Van  der  Waals'  equation,  182-184 
and  vibrations  of  atoms  in  crystals, 

211-240 
Interference     of    light,     and     quantum 

theory,  319-320 

Intermolecular  forces,  in  gases,  410-414 
Internal  energy,  6-9,  17 
at  absolute  zero,  179 
and  melting,  258-269 
of  mixture  of  gases,  123 
of  perfect  gas,  Boltzmann  statistics, 

117 

Fermi-Dirac  statistics,  77-78,  81-82 
and  phase  change  of  second  order,  295, 

301 

of  solids,  205-220 
of  solutions,  275-277 
Internal  pressure,  182-184 
Iodine,   characteristic   temperature,   for 

rotation,  136 

for  vibration,  142 

crystal  structure,  418 


514 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Iodine,  dissociation,  133 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 

and  homopolar  bond,  400-408 

and  organic  compounds,  426 
Ionic  crystals,  375,  377-399 
Ionic  radii,  382-385 
lonization,  of  atoms,  321-335 
lonization  potential,  322,  334,  343 

table  of,  348 

Ions,  and  atomic  structure,  321-335, 337- 
338,  351 

forces  between,  357-358 

formation  in  solution,  272-274,  290 
Iridium,  crystal  structure,  447 
Iron,  crystal  structure,  447 

data  regarding  melting  point,  259,  261 

Debye  temperature,  237 

equation  of  state  and  crystal  structure, 
451,  454 

molecular  volume,  261 

thermal  expansion,  261 
Irreversible  process,  11-13,  16 

and  kinetic  approach  to  equilibrium, 
86-92,  96-98 

and  statistical  mechanics,  43-46 
Isomers,  423 
Isothermal  processes,  19 
Isothermals,  of  solid,  200 

and  Van  der  Waals'  equation,  184^186 
Isotopes,  336-337 


Joule,  5 

Joule-Thomson  effect,  196-198 

Joule's  law,  30,  115 

K 

Kinetic  energy,  and  exchange  effect,  369 
and  Maxwoll-Boltzmann  law,  60 
of  polyatomic  molecules,  134,  144 
Kinetic  theory,  15,  86-100 
and  chemical  reactions,  151-154,  158- 

165 

and  radiation,  324-333 
and    thermionic    emission,    465-471, 

480-484 

Kirchhoff's  law,  309-310 
Krypton,  atomic  volume,  384 
data  regarding  crystals,  416 


Krypton,  specific  heat,  130 
Van  der  Waals  constants,  408 


Lanthanum,  crystal  structure,  447 

equation  of  state,  451 
Latent  heat,  of  evaporation  of  electrons, 

464,  469-470,  481-484 
of  fusion,  171-180,  258-269 
of  vaporization,  171-180,  258-260 
of  metals,  452-454 
of  organic  compounds,  table,  434 
table  of,  414 

and  Van  der  Waals'  equation,  189 
Lattice  energies,  alkali  halides,  395-396 
Lattice  spacings,  ionic  crystals,  381-382 

metals,  447 

Lead,  crystal  structure,  447 
data  regarding  melting,  259 
Debye  temperature,  237 
equation  of  state,  451 
Lead    bromide,    chloride,    iodide,    data 

regarding  melting  point,  259 
Lewis,    G.    N.,    and    homopolar    bond, 

400-408 

Linear  oscillator,  and  black-body  radia- 
tion, 314r-316 

and  equipartition  of  energy,  58 
quantum  theory,  39-40,  42 
and  vibration  of  atoms  and  molecules 

in  crystals,  211-240 
and  vibration  of  diatomic  molecules, 

140-149 

Liouville's  theorem,  37-38,  44-46,  88 
Liquefaction  of  gases,  198 
Liquids,  166-174,  256-269 
Liquidus,  281 

Lithium,  compressibility,  202 
crystal  structure,  447 
equation  of  state  and  energy,  451,  454- 

455 
molecule,   characteristic  temperature, 

for  rotation,  136 
for  vibration,  142 
dissociation,  133 
heat    of    dissociation,    interatomic 

distance,  Morse  constant,  i32 
Lithium     fluoride,     chloride,     bromide, 
iodide,  data  regarding  crystals,  381, 
393,  395 


INDEX 


615 


Lithium  nitrate,  data  regarding  melting 

point,  259 
Longitudinal  waves  in  solids,  222-240 


M 


Madelung,  electrostatic  energy  of  crys- 
tals, 385-388 

Magnesium,  crystal  structure,  447 
data    regarding   melting   point,    259, 

261 

equation  of  state  and  energy,  451,  454 
molecular  volume,  261 
phase    equilibrium    in    alloys,    274, 

287-288 

thermal  expansion,  261 
Magnesium    oxide,    sulphide,    selenide, 

data  regarding  crystals,  381 
Manganese,  crystal  structure,  447-448 
data  regarding  melting  point,  259 
equation  of  state,  451 
Mass  action  law,  151-158 

applied  to  atomic  processes,  334r-335 
Maxwell-Boltzmaim  distribution,  52-64 
and  activation,  159 
and    Fermi-Dirac    and    Einstein-Bose 

statistics,  74,  84 
and  fluctuations,  101-104 
and  mean  moment  of  rotating  dipole, 

361,364 

Maxwell's  demon,  45-46 
Maxwell's     distribution     of    velocities, 

55-58 

in  arcs,  333 

and  kinetic  methods,  91-96 
Maxwell's  relations,  26 
Mean  free  path,  328 
Mechanical  equivalent  of  heat,  5,  8 
Mechanical  work  (see  External  work) 
Melting,  23,  166-169,  171-176,  256-269 

of  alloys,  278-290 
Melting  points,  of  chain  compounds,  422 

ionic  crystals,  table,  381-382 
Mercuric  bromide,  data  regarding  melt- 
ing point,  259 
Mercuric  iodide,  data  regarding  melting 

point,  259 
Mercuric    sulphide,    selenide,    telluride, 

data  regarding  crystals,  382 
Mercury,  crystal  structure,  447 
data  regarding  melting  point,  259 


Metallic  bond,  374-376,  451-464 
Metals,  376,  444-470,  472-501 
Metasilicate  ion,  structure,  437 
Metastable  equilibrium,  163 
Methane,  boiling  point,  426 
data  regarding  melting  point,  259 
heat  of  vaporization,  434 
valence  structure  of  molecule,  401 
Van  dcr  Waals  constants,  408,  411 
Methyl  alcohol,  data  regarding  melting 

point,  259 

heat  of  vaporization,  434 
structure  of  molecule,  427 
Van  der  Waals  constants,  408 
Methyl  chloride,  boiling  point,  426 
heat  of  vaporization,  434 
structure  of  molecule,  425 
Van  der  Waals  constants,  408 
Methyl  ether,  Van  der  Waals  constants, 

408 
Methylamine,  structure  of  molecule,  427 

Van  dcr  Waals  constants,  408 
Mica,  structure,  439 
Microcanonical  assembly,  46 
Microscopic  properties,  32-35 
Microscopic  reversibility,  88,  331 
Mixture  of  gases,  120-124 

and  chemical  equilibrium,  154-158 
Molecular  orbitals,  368 
Molecular  phase  space,  and  Fermi-Dirac 

and  Einstein-Bose  statistics,  65-86 
and  Maxwell-Boltzmann  distribution, 

52-64 
Molecular  volume,  of  liquids,  table,  408 

table  of,  261 

Molecules,  sizes  of,  409-410 
in  valence  compounds,  375-376 

(See  also  Diatomic  molecules;  Poly- 
atomic molecules) 

Molybdenum,  crystal  structure,  447 
Debye  temperature,  237 
equation  of  state  and  crystal  Btnicture, 

451,  454 
Moment  of  inertia,  diatomic  molecule, 

134r-135 

Momentum,  36,  38-43 
and  approach  to  equilibrium,  87 
of  radiation,  311 

Monatomic  gases,   and    Maxwell-Boltz- 
mann distribution,  53-64 


516 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Morse  curves,  133 
and  metals,  452 
Multiplets,  344 

N 

Naphthalene,  heat  of  vaporization,  434 
structure  of  crystal,  432 
structure  of  molecule,  430 
Van  der  Waals  constants,  408 
Neon,  atomic  volume,  384 
data  regarding  crystals,  416 
specific  heat,  130 
Van.  der  Waals  constants,  408 
Newton's  second  law  of  motion,  3 
NH,     characteristic     temperature,     for 

rotation,  136 
for  vibration,  142 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
Nickel,  alloys  with  copper,  458 
crystal  structure,  447 
data  regarding  melting  point,  259 
equation  of  state  and  energy,  451,  454 
phase  equilibrium  in  alloys,  274,  279- 

282 
Nitrate  ion,  357 

valence  structure,  406 
Nitric  oxide,  characteristic  temperature 

for  rotation,  136 
for  vibration,  142 
data  regarding  melting  point,  259 
heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
specific  heat,  137 
valence  structure,  403 
Van  der  Waals  constants,  408 
Nitrogen,  characteristic  temperature,  for 

rotation,  136 
for  vibration,  142 
crystal  structure,  417 
data  regarding  melting  point,  259 
dissociation,  133 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
heat  of  vaporization,  414 
and  homopolar  bonds,  400-408 
Van  der  Waals  constants,  408 
Nitrous  oxide,  Van  der  Waals  constants, 

408 

Nonconservative  force,  3 
Normal  modes  of  vibration,  215,  222-255 
Nucleus,  atomic,  336 


Octane,  Van  der  Waals  constants,  408 
OH,  characteristic  temperature,  for  rota- 
tion, 136 

for  vibration,  142 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
Ohm's  law,  484-489 
Opalescense,  111 

Order-disorder  transition,  293-304 
Organic  compounds,  376,  420-434 
Orthosilicate  radical,  structure,  435 
Osmium,  crystal  structure,  447 
Overtones,  215,  222-255 
Oxygon,  characteristic  temperature,  for 

rotation,  136 
for  vibration,  142 
combination  with  hydrogen   to  form 

water,  151-164 

data  regarding  melting  point,  259 
heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
heat  of  vaporization,  414 
and  homopolar  bonds,  400-408 
Van  der  Waals  constants,  408 


Palladium,  crystal  structure,  447 

equation  of  state,  451 
Partial  derivatives,  16,  18-20,  23-30 
Partial  pressure,  120-121 

and  chemical  equilibrium,  151-158 
Partition  function,  and  fluctuations,  106 

and  melting,  265-269 

for  perfect  gas,  125-128 

for  rotation,  diatomic  molecule,  138- 
140 

and  second  virial  coefficient,  191-196 

for  solids,  213-218 

and  statistical  mechanics,  50-51 

for  vibration,  diatomic  molecule,  143 
Pauli  exclusion  principle,  342 

and  exchange  effect,  369 

and  metals,  475-476 
Perfect  gas,  17,  30 

chemical  equilibrium  in,  150-165 

and  Maxwell-Boltzmann  distribution, 
53-64 

polyatomic,  130-149 


INDEX 


517 


Perfect  gas,  thermodynamic  and  statis- 
tical treatment,  115-129 
translational  energy  levels  in  quantum 

theory,  54-55 

Periodic  potential  in  metal,  473,  489-501 
Periodic  table,  344-351 
Permutations,  67 

Phase  changes  of  second  order,  291-304 
Phase  diagram,  166-168 

binary  systems,  281-287 
Phase  equilibrium,  166-170 

binary  system,  270-290 
Phase  space,  36-43 
and  Maxwell-Boltzmann  distribution, 

52-64 

Phosphine,   valence   structure   of  mole- 
cule, 405 

Van  dcr  Waals  constants,  408 
Phosphonium  ion,  378 
Photoelectric  effect,  316-320 
Photon,  316-320,  323-326 
Planck,  black-body  radiation,  307-320, 

325-326 

probability  and  entropy,  34 
Planck's  constant,  39 
Platinum,  crystal  structure,  447 
data  regarding  melting  point,  259 
Debyc  temperature,  237 
equation  of  state  and  energy,  451,  454 
Poisson's  ratio   and  velocity  of  elastic 

waves,  238,  240 
Polarization,     and    interatomic    forces, 

363-367,  398-399,  410-414 
of  light,  308 

Polyatomic  gases,  130-149 
internal  coordinates,  124 
Polymorphic  phases,  167-170,  180-181, 

22O-221,  448 

Potassium,  compressibility,  202 
crystal  structure,  447 
data  regarding  melting  point,  259,  261 
Debye  temperature,  237 
equation  of  state  and  energy,  451,  454- 

455 

molecular  volume,  261 
molecule,   characteristic   temperature, 

for  rotation,  136 
for  vibration,  142 
dissociation,  133 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
thermal  expansion,  261 


Potassium  bromide,  data  regarding  crys- 
tals, 381,  393,  395 

data  regarding  melting  point,  259,  261 
molecular  volume,  261 
thermal  expansion,  261 
Potassium  chloride,  data  regarding  crys- 
tals, 381,  393,  395 

data  regarding  melting  point,  259,  261 
Debye  temperature,  391 
molecular  volume,  261 
thermal  expansion,  261 
Potassium   dichromate,   data   regarding 

melting  point,  259 

Potassium  fluoride,  data  regarding  crys- 
tals, 381,  393,  395 
data  regarding  melting  point,  259 
Potassium    hydroxide,    data    regarding 

melting  point,  259 

Potassium  iodide,  data  regarding  crys- 
tals, 381,  393,  395 

Potassium  nitrate,  data  regarding  melt- 
ing point,  259 

Potential,  electrostatic,  353-367 
Potential  energy,  3 
of  interatomic  forces,   131-133,   191- 

196,  352-376 
and  Maxwell-Boltzmann  distribution, 

54,  62-64 
Pressure,  17 

and  chemical  equilibrium,  151-158 
and  equation  of  state  of  ionic  crystals, 

392-396 

and  equation  of  state  of  metals,450-456 
and  equation  of  state  of  solids,  199-221 
and  equilibrium  of  phases,  166-181 
of  imperfect  gases,  and  Van  der  Waals' 

equation,  182-198 

of  mixtures  of  gases,  and  partial  pres- 
sure, 120-121,  128 
of  perfect  gases,  58-60 

Fermi- Dirac  statistics,  79 
of  radiation,  311 
Probability,  a  priori,  36,  38,  127 
and  Maxwell-Boltzmann  distribution, 

53 

thermodynamic,  34 
of  transition,  324-333 
Propane,  heat  of  vaporization,  434 
structure  of  molecule,  421 
Van  der  Waals  constants,  408 
Pyrometer,  312-313 


518 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Q 

Quadrupole  moment,  356-357 
Quantum  defect,  340-341 
Quantum  theory,  36,  38-43,  45-46 

and  atomic  structure,  339-344 

and  equation  of  state  and  specific  heat 
of  solids,  215-221,  234-240 

and  identical  particles,  129 

and  kinetic  method,  96-100 

and  liquids,  265 

and  radiation,  314-320 

and  specific  heat  of  polyatomic  gases, 
138-149 

and  structure  of  metals,  489-501 

and  translational  energy  levels,  54H>5 
Quasi-ergodic  motion,  38 
Quenching,  181 

R 

Radiation,  307-326 
Radii,  of  atoms,  342 
table  of,  349 

of  ions,  382-385 
Ramsauer  effect,  330 
Randomness,  9-12,  32-35,  43-46 

and  melting,  262-264 
Rate  of  reaction,  150-154,  158-165 
Ray  leigh- Jeans  law,  314 
Reciprocal  space,  230,  245 
Recombination,  165 
Reduced  mass,  135,  141 
Reflection  of  electrons  by  metals,  462,  465 
Relaxation   time,    in   electrical    conduc- 
tivity, 485-486 

Representative  point,  36-38,  65-66 
Repulsive  forces,   between   atoms,   and 
exclusion  principle,  369-372 

between  ions,  ionic  crystal,  388-390 

between  molecules,  130-133 

and  Van  der  Waals'  equation,  182-184, 

194-196 

Residual  rays,  254r-255 
Resistance,  electrical,  484-489,  498-501 
Resonance  potential,  322 
Reststrahlen,  254-255 
Reversible  processes,  10-13,  16 

and  statistical  mechanics,  44-51 
Rheology,  256-258 
Rhodium,  crystal  structure,  447 
Richardson  equation,  thermionic  emis- 
sion, 466 


Rigid  sphere  atomic  model,  131 
Rotator,  diatomic  molecule,  134-140 

and  equipartition  of  energy,  58 

in  quantum  theory,  40,  42 
Rubidium,  compressibility,  202 

crystal  structure,  447 

data  regarding  melting  point,  259 

equation  of  state  and  energy,  451,  454 
Rubidium    fluoride,    chloride,    bromide, 
iodide,  data  regarding  crystals,  381, 
393,  395 
Rumford,  5 

Ruthenium,  crystal  structure,  447 
Rydberg  number,  340 


8 


Saha,  335 

Saturation  of  valence,  374-376 

Second  law  of  thermodynamics,  12-14,  16 

and  statistics,  19,  49-51 
Second  order,  phase  change,  291-304 
Second  virial  coefficient,  190-196 
Secondary  emission  of  electrons,  461-462 
Selenium,  crystal  structure,  444,  447,  450 

data  regarding  melting  point,  259 
Shielding  constant,  340-342 
Shot  effect,  108 
Silica  gel,  438 
Silicates,  435-443 
Silicon,  crystal  structure,  444,  447 

and  homopolar  bonds,  400 

melting  point,  449 

Silicon    hydride,    valence    structure    of 
molecule,  405 

Van  der  Waals  constants,  408 
Silver,  crystal  structure,  447 

data  regarding  melting  point,  259,  261 

Debye  temperature,  237 

equation  of  state  and  energy,  451,  454 

molecular  volume,  261 

thermal  expansion,  261 
Silver  bromide,  data  regarding  crystals, 
381 

data  regarding  melting  point,  259-261 

molecular  volume,  261 

thermal  expansion,  261 
Silver  chloride,  data  regarding  crystals, 
381 

data  regarding  melting  point,  259,  261 

molecular  volume,  261 

thermal  expansion,  261 


INDEX 


519 


Silver  fluoride,  data  regarding  crystals, 

381 
Silver   nitrate,   data   regarding   melting 

point,  259 
Simple    cubic    structure,    and    atomic 

vibrations,  232 
Sodium,  crystal  structure,  447 

data  regarding  melting  point,  259,  261 
Debye  temperature,  237 
electronic  energy  bands,  494 
equation  of  state  and  energy,  451,  454- 

455 
and  thermodynamic  functions,  200-- 

211 

molecular  volume,  261 
molecule,    characteristic   temperature, 

for  rotation,  136 
for  vibration,  142 
dissociation,  133 

heat  of  dissociation,  interatomic  dis- 
tance, Morse  constant,  132 
thermal  expansion,  261 
Sodium  acetate,   structure  of  molecule, 

428 
Sodium  bromide,  iodide,  data  regarding 

crystals,  381,  393,  395 
Sodium  chloride,  crystal  structure,  378, 

381 

data  regarding  crystals,  381,  393,  395 
data  regarding  melting  point,  259,  261 
Debye  temperature,  391 
molecular  volume,  261 
thermal  expansion,  261 
water  solution,  285-287 
Sodium  fluoride,  data  regarding  crystals, 

381,  393,  395 

data  regarding  melting  point,  259 
Sodium  hydroxide,  data  regarding  melt- 
ing point,  259 
Sodium  nitrate,  data  regarding  melting 

point,  259 

Sodium  perchlorate,  data  retarding  melt- 
ing point,  259 
Solids,  binary  systems,  270-290 

equation   of  state  and  specific   heat, 

199-255 

equilibrium  with  other  phases,  166-181 
ionic  substances,  377-399 
melting,  256-269 
metals,  444-501 
Solidus,  281 
Solutions,  270-290 


Space  quantization,  139,  339 
Specific  heat,  17-20,  22-23 
of  compounds,  241-255 
difference  between  Cp  and  CV,  90 
electronic,  322 
and  fluctuations,  107-108 
of  free  electrons,  471, 476-479,  49&-500 
internal,  of  gas,  117 
of  ionic  crystals,  390-393 
of  liquids,  262-265 
of  monatomic  perfect  gases,  61-62 

Fermi-Dirac  statistics,  78-79 
and   phase   changes  of  second  order, 

291-304 

of  polyatomic  gases,  136-149 
of  solids,  203-205,  213-214,  222-255 
and  statistical  mechanics,  51 
jind   temperature   variation   of  latent 

heat,  177-178 
variation  with  pressure  and  volume, 

116 

Spectrum,  infrared,  of  crystals,  254-255 
optical,  and  Kirchhoff's  law,  300-310 
of  vibrational  frequencies,  solids,  225- 

255 
Spin,  electronic,  339 

and  interatomic  forces,  369-374 
and  metals,  476 
Spontaneous  emission,  324 
Standing  waves,  226-231,  242-252 

of  electrons  in  metals,  489-501 
Stationary  states,  of  electrons  in  atoms, 

321-323,  338-344 
of  oscillator  and  rotator,  41-42 
Statistical  mechanics,  14,  32-85 

applied  to  black  body  radiation,  307- 

320 
applied  to  chemical  equilibrium,  154- 

158 
applied  to  equation  of  state  of  solids, 

211-221 
applied  to  equilibrium  between  phases 

and  vapor  pressure,  178-180 
applied  to  melting,  265-269 
applied  to  perfect  gas,  124-129 
applied  to  polyatomic  gases,  138-140, 

142-145 
applied    to    second    virial    coefficient, 

190-196 

applied  to  solubility  and  phase  equilib- 
rium, 270-304 
Stefan-Boltzmann  law,  307-313 


520 


INTRODUCTION  TO  CHEMICAL  PHYSICS 


Stirling's  theorem,  70-72 

Streamline  flow  of  representative  points, 

37,94 

Stresses  and  strains,  17,  199 
Strontium,  crystal  structure,  447 

equation  of  state,  451 
Strontium  oxide,  sulphide,  selenide,  tellu- 

ride,  data  regarding  crystals,  381 
Sugar,  phase  equilibrium  in  solution,  270 
Sulphate  ion,  357,  398 
Sulphur    dioxide,    valence    structure    of 

molecule,  406 

Van  der  Waals  constants,  408 
Supercooling,  181,  256-258,  262-264 


Tantalum,  crystal  structure,  447 

equation  of  state,  451 
Tellurium,   crystal  structure,  444,  447, 

450 
Temperature,  9,  12-14,  17 

bath,  and  canonical  assembly,  46-47 

and  fluctuations,  101 
of  inversion,  Joule-Thomson  effect,  1 98 
and  kinetic  method,  96 
Temperature-entropy  diagram,  172 
Terms,  spectroscopic,  323 
Thallium,  crystal  structure,  447 

data  regarding  melting  point,  259 
Thallium  bromide,  data  regarding  melt- 
ing point,  259 

Thallium  chloride,  data  regarding  crys- 
tals, 382 

data  regarding  melting  point,  259 
Thallium  iodide,  data  regarding  crystals, 

QQO 
OoZ 

Thermal  equilibrium,  37-38,  46-51 

and  kinetic  method,  96,  98 
Thermal  expansion,   19,   200-220,   238- 
240,  261 

of  ionic  crystals,  392-394 

of  metals,  450-456 
Thermal  pressure,  217-218 
Thermionic  emission,  460-471,  480-484 
Thermodynamic  formulas,  16,  23-30 

table  of,  27-29 
Thermodynamic  probability,  34 

and    Fermi-Dirac   and    Einstein-Bose 

statistics,  69-72 

Thermodynamic   scale   of   temperature, 
30-31 


Thermodynamics,  14,  16-31 
applied  to  black-body  radiation,  307- 

320 
applied  to  chemical  equilibrium,  154- 

158 
applied  to  equation  of  state  of  solids, 

199-211 
applied  to  equilibrium,  between  atoms 

and  electrons,  333-335 
between  metal  and  gas,  463-464 
between  phases,  174-178 
applied  to  perfect  gas,  115-124 
applied  to  solubility  and  phase  equilib- 
rium, 270-304 
applied  to  Van  der  Waals'  equation, 

18^-189 

Threshold,  photoelectric,  318 
Tin,  crystal  structure,  444,  447-449 
data  regarding  melting  point,  259 
Debye  temperature,  237 
Titanium,  crystal  structure,  447 
Transition  probability,  42,  88 

of  atoms,  322-333 
Transverse    waves,    in    electromagnetic 

radiation,  313-314 
in  solids,  222-240 
Traveling  waves,  226-228 
Triethylamine,  Van  der  Waals  constants, 

408 
Trimethylamine,  structure  of  molecule, 

427 

Van  der  Waals  constants,  408 
Triple  point,  166-167,  171-172,  181 

and  eutectic,  285 
Tungsten,  crystal  structure,  447 

equation  of  state  and  energy,  451,  454 

U 

Uncertainty  principle,  40-41 
Undetermined  multipliers,  48 


Valence  forces,  130 

explanations  from  atomic  theory,  371- 

376 
Vanadium,  crystal  structure,  447 

equation  of  state,  451 
Van  der  Waals  constants,  182-198 

for  molecular  substances,  407-414 
Van  der  Waals'  equation,  182-198,  210 

and  molecular  substances,  407-414 


INDEX 


521 


Van  der  Waals  forces,  182-198 

and  atomic  structure,  356-374 

and  inert  gases,  385 

and  molecular  substances,  407-414 

and  solutions,  273 
Van't  Hoff's  equation,  154-158 
Vapor  pressure,  166-169,  174-180 

and  Van  der  Waals'  equation,  188-189 
Vaporization,  23 

entropy  of,  171-180 

heat  of,  171-180 

of  metals,  452-^54 

of  organic  compounds,  table,  434 

table  of,  414 

and   Van   der   Waals7   equation,    189, 

258-260 
Velocity,  of  elastic  waves,  227-240 

of  light,  308 
Vibration,  of  diatomic  molecules,  140-149 

and  light  waves,  313-314 

of  molecules   and   atoms  in   crystals, 

211-255 

Vibrational  degrees  of  freedom,  146 
Virial,  190-396 
Viscosity,  12,  13,  257 
Volt,  electron,  132-133,  318 
Volta  effect,  467-471 
Volume,  17 

and  chemical  equilbrium,  151-158 

of  imperfect  gases,  and  Van  der  Waals' 
equation,  182-198 

molecular,  table  of,  261 
of  liquids,  table,  408 

of  perfect  gas,  58-61 

and  phase  equilibrium,  168-178 

of  solids,  199-221 


W 


Water,  crystal  structure,  260,  418-419 
data  regarding  melting  point,  259 
dissociation  into  hydrogen  and  oxygen, 

151-164 

entropy  and  free  energy,  172-173 
equilibrium  between  phases,  166-169 
heat  of  vaporization,  414 


Water,  and  solubility,  270-275,  285-290 
valence  structure  of  molecule,  401 
Van  der  Waals  constants,  408, 412-414 
vapor   pressure   and   latent   heat   of 

vaporization,  188-189 
vibrational  specific  heat,  146-147 
Water  glass,  438 
Wave  mechanics,  41,  307 
and    electrons    in    periodic    potential, 

489-501 

and  radiation,  323 
Waves,    elastic,    in    continuous    media, 

227-234 
electromagnetic,  and  light,  308,  313- 

320 

in  molecular  media,  241-252 
Work,  3,  7-9,  17,  21-22 

and  statistical  mechanics,  49 
Work  function,  thermionic  and  photo- 
electric,    317-318,     464,     469-470, 
480-484 
Wurtzite,  crystal  structure,  379-380,  382 


Xenon,  atomic  volume,  384 
data  regarding  crystals,  416 
specific  heat,  130 
Van  dor  Waals  constants,  408 
X-ray  diffraction,  liquids,  256 
X-ray  levels  in  atoms,  344 


Z 


Zinc,  crystal  structure,  447 
data  regarding  melting  point,  259 
Debye  temperature,  237 
energy,  454 

order-disorder  in  alloys,  293-304 
phase    equilibrium    in    alloys,    270, 

287-288 
Zinc  oxide,  sulphide,  selenide,  telluride, 

data  regarding  crystals,  382 
Zincblende,   crystal  structure,   379-381, 

382 

Zirconium,  crystal  structure,  447 
equation  of  state,  451