Skip to main content

Full text of "Chemistry, an experimental science"

See other formats


CxxBIVxIoX^JbvY 


AN      EXPERIMENTAL      SCIENCE 


CHEMICAL  EDUCATION  MATERIAL  STUDY 


'*&?.£&"** 


<wi$,  the  one-millionth  copy  of1 

CHEMISTRY: 

AN  EXPERIMENTAL  SCIENCE 

isvrescntedto 
die  President  and  Regents  oftlic  University  of-* 
California  to  commemorate  the  University s 
contribution  to  tlu  eminently successfub  : 
development  of  a  hujh  school  chemistry 
cutriculum.rm7his project  was  made-^ 
possible  by  arrant  from  thelsational 
Science  Foundation.  CRoy aides  were-^ 
returned  to  tlieU.S.Treasury  as  complete 
repayment  f the  aran  v.  <~jlieyrojecZy 
directed  by  faculty  of  tlu  University  of 
California  and  of  Harvey  Mudd College, 
resulted  iru>  tins  textbook,  a  laboratoiy 
manual,  a  teadier'sguidej  ancLo 
twenty-seven  associated  funis. 


18  March  iy$3 


Si 

a 
s 


2J  >H  vO  ^ 

•    §5    *S    *g    5 


Q>  n.  °& 

>  CO        g 


N        g  ^        ON        ck 


lr> 


5  s*  1 


M 


O  CN 


$    !j?    SJ    § 


^55 


^    5    5 


w 


^    v| 


8  s  a 


On 


f*     cj     >s     5? 


CO        O        Ci         CN 

*0  >  ON  ^ 


fc        °°"        £ 

"^      to      co 


Cj 

K 


6| 


V 

& 

CI 
Cj 

€ 

K 
ca 


3  *■  J5*   5*   £*>   $*   510   3 


INTERNATIONAL  ATOMIC  WEIGHTS 


ATOMIC 

ATOMIC 

ATOMIC 

ATOMIC 

NAME 

SYMBOL 

NUMBER 

WEIGHT 

NAME 

SYMBOL 

NUMBER 

WEIGHT 

Actinium 

Ac 

89 

(227) 

Mercury 

Hg 

80 

200.6 

Aluminum 

Al 

13 

27.0 

Molybdenum 

Mo 

42 

95.9 

Amencium 

Am 

95 

(243) 

Neodymium 

Nd 

60 

144.2 

Antimony 

Sb 

51 

121.8 

Neon 

Ne 

10 

20.2 

Argon 

Ar 

18 

39.9 

Neptunium 

Np 

93 

(237) 

Arsenic 

As 

33 

74.9 

Nickel 

Ni 

28 

58.7 

Astatine 

At 

85 

(210) 

Niobium 

Nb 

41 

92.9 

Barium 

Ba 

56 

137.3 

Nitrogen 

N 

7 

14.01 

Berkelium 

Bk 

97 

245 

Osmium 

Os 

76 

190.2 

Beryllium 

Be 

4 

9.01 

Oxygen 

O 

8 

16.00 

Bismuth 

Bi 

83 

209.0 

Palladium 

Pd 

46 

106.4 

Boron 

B 

5 

10.8 

Phosphorus 

P 

15 

31.0 

Bromine 

Br 

35 

79.9 

Platinum 

Pt 

78 

195.1 

Cadmium 

Cd 

48 

112.4 

Plutonium 

Pu 

94 

(242) 

Calcium 

Ca 

20 

40.1 

Polonium 

Po 

84 

210 

Californium 

Cf 

98 

(251) 

Potassium 

K 

19 

39.1 

Carbon 

C 

6 

12.01 

Praseodymium 

Pr 

59 

140.9 

Cerium 

Ce 

58 

140.1 

Promethium 

Pm 

61 

(147) 

Cesium 

Cs 

55 

132.9 

Protactinium 

Pa 

91 

(231) 

Chlorine 

CI 

17 

35.5 

Radium 

Ra 

88 

(226) 

Chromium 

Cr 

24 

52.0 

Radon 

Rn 

86 

(222) 

Cobalt 

Co 

27 

58.9 

Rhenium 

Re 

75 

186.2 

Copper 

Cu 

29 

63.5 

Rhodium 

Rh 

45 

102.9 

Curium 

Cm 

96 

(247) 

Rubidium 

Rb 

37 

85.5 

Dysprosium 

Dy 

66 

162.5 

Ruthenium 

Ru 

44 

101.1 

Einsteinium 

Es 

99 

(254) 

Samarium 

Sm 

62 

150.4 

Erbium 

Er 

68 

167.3 

Scandium 

Sc 

21 

45.0 

Europium 

Eu 

63 

152.0 

Selenium 

Se 

34 

79.0 

Fermium 

Fm 

100 

(253) 

Silicon 

Si 

14 

28.1 

Fluorine 

F 

9 

19.0 

Silver 

Ag 

47 

107.9 

Francium 

Fr 

87 

(223) 

Sodium 

Na 

11 

23.0 

Gadolinium 

Gd 

64 

157.3 

Strontium 

Sr 

38 

87.6 

Gallium 

Ga 

31 

69.7 

Sulfur 

S 

16 

32.1 

Germanium 

Ge 

32 

72.6 

Tantalum 

Ta 

73 

180.9 

Gold 

Au 

79 

197.0 

Technetium 

Tc 

43 

(99) 

Hafnium 

Hf 

72 

178.5 

Tellurium 

Te 

52 

127.6 

Helium 

He 

2 

4.00 

Terbium 

Tb 

65 

158.9 

Holmium 

Ho 

67 

164.9 

Thallium 

Tl 

81 

204.4 

Hydrogen 

H 

1 

1.008 

Thorium 

Th 

90 

232.0 

Indium 

In 

49 

114.8 

Thulium 

Tm 

69 

168.9 

Iodine 

I 

53 

126.9 

Tin 

Sn 

50 

118.7 

Iridium 

Ir 

77 

192.2 

Titanium 

Ti 

22 

47.9 

Iron 

Fe 

26 

55.8 

Tungsten 

W 

74 

183.9 

Krypton 

Kr 

36 

83.8 

Uranium 

U 

92 

238.0 

Lanthanum 

La 

57 

138.9 

Vanadium 

V 

23 

50.9 

Lead 

Pb 

82 

207.2 

Xenon 

Xe 

54 

131.3 

Lithium 

Li 

3 

6.94 

Ytterbium 

Yb 

70 

173.0 

Lutetium 

Lu 

71 

175.0 

Yttrium 

Y 

39 

88.9 

Magnesium 

Mg 

12 

24.3 

Zinc 

Zn 

30 

65.4 

Manganese 

Mn 

25 

54.9 

Zirconium 

Zr 

40 

91.2 

Mendelevium 

Md 

101 

(256) 

Parenthetical  names  refer  to  radioactive  elements;  the  mass  number  (not  the  atomic  weight)  of  the  isotope  with  largest 
half-life  is  usually  given. 

*  Latest  values  recommended  by  the  International  Union  of  Pure  and  Applied  Chemistry,  1961. 


CHEMISTRY 

AN   EXPERIMENTAL   SCIENCE 


CHEMISTRY 

Prepared  by 
CHEMICAL  EDUCATION  MATERIAL  STUDY 

Under  a  grant  from 
THE  NATIONAL  SCIENCE  FOUNDATION 


Editor:  GEORGE  C.  PIMENTEL,    University  of  California,  Berkeley,  California 

Associate  Editors 

BRUCE  H.  MAHAN,    University  of  California,  Berkeley,  California 

A.  L.  McCLELLAN,    California  Research  Corporation,  Richmond,  California 

KEITH  MacNAB,    Sir  Francis  Drake  High  School,  San  Anselmo,  California 

MARGARET  NICHOLSON,    Acalanes  High  School,  Lafayette,  California 

An  Experimental  Science 


Contributors 

ROBERT  F.  CAMPBELL 

Miramonte  High  School,  Orinda,  California 

JOSEPH  E.  DAVIS,  JR. 

Miramonte  High  School,  Orinda,  California 

SAUL  L.  GEFFNER 

Forest  Hills  High  School,  Forest  Hills,  New  York 

THEODORE  A.  GE1SSMAN 

University  of  California,  Los  Angeles,  California 

MELVIN  GREENSTADT 

Fairfax  High  School,  Los  Angeles,  California 

CARL  GRUHN 

South  Pasadena  High  School,  South  Pasadena,  California 

EDWARD  L.  HAENISCH 

Wabash  College,  Crawfordsville,  Indiana 

ROLFE  H.  HERBER 

Rutgers  University,  New  Brunswick,  New  Jersey 

C.  ROBERT  HURLEY 

Sacramento  State  College,  Sacramento,  California 

LAWRENCE  D.  LYNCH,  JR. 

Beverly  Hills  High  School,  Beverly  Hills,  California 


LLOYD  E.  MALM 

University  of  Utah,  Salt  Lake  City,  Utah 

CLYDE  E.  PARRISH 

Cubberley  Senior  High  School,  Palo  Alto,  California 
ROBERT  W.  PARRY 

University  of  Michigan,  Ann  Arbor,  Michigan 

EUGENE  ROBERTS 

Polytechnic  High  School,  San  Francisco,  California 

MICHELL  J.  SIENKO 

Cornell  University,  Ithaca,  New  York 

ROBERT  SILBER 

American  Chemical  Society,  Washington,  D.C. 

HARLEY  L.  SORENSEN 

San  Ramon  Valley  Union  High  School,  Danville,  California 

LUKE  E.  STEINER 

Oberlin  College,  Oberlin,  Ohio 

MODDIE  D.  TAYLOR 

Howard  University,  Washington,  D.C. 

ROBERT  L.  TELLEFSEN 

Napa  High  School,  Napa,  California 


Director:  J.  ARTHUR  CAMPBELL,     Harvey  Mudd  College,  Claremont,  California 
Chairman:  GLENN  T.  SEABORG,    University  of  California,  Berkeley,  California 


W.    H.    FREEMAN   AND   COMPANY,   Cooperating  Publishers 

SAN    FRANCISCO 


©  Copyright  1960,  1961,  1962,  1963  by  The  Regents  of  the  University  of  California. 

The  University  of  California  reserves  all  rights  to  reproduce  this  book, 
in  whole  or  in  part,  with  the  exception  of  the  right  to  use 
short  quotations  for  review  of  the  book. 

Printed  in  the  United  States  of  America. 

Library  of  Congress  Catalog  Card  Number:  63-18323. 

ISBN:  0-7 167-0001 -8 

20 


Preface 


Chemistry  deals  with  all  of  the  substances  that 
make  up  our  environment.  It  also  deals  with  the 
changes  that  take  place  in  these  substances — 
changes  that  make  the  difference  between  a  cold 
and  lifeless  planet  and  one  that  teems  with  life 
and  growth.  Chemistry  helps  us  understand  and 
benefit  from  nature's  wondrous  ways. 

Chemistry  is  an  important  part  of  what  is 
called  science.  Since  every  phase  of  our  daily 
life  is  affected  by  the  fruits  of  scientific  activity, 
we  all  should  know  what  scientific  activity  is, 
what  it  can  do,  and  how  it  works.  The  study  of 
chemistry  will  help  you  learn  these  things. 

CHEMISTR  Y — An  Experimental  Science  pre- 
sents chemistry  as  it  is  today.  It  does  so  with 
emphasis  upon  the  most  enjoyable  part  of  chem- 
istry: experimentation.  Unifying  principles  are 
developed,  as  is  appropriate  in  a  modern  chemis- 
try course,  with  the  laboratory  work  providing 
the  basis  for  this  development.  When  we  are 
familiar  with  these  widely  applicable  principles 
we  no  longer  have  need  for  endless  memorization 
of  innumerable  chemical  facts.  To  see  these  prin- 
ciples grow  out  of  observations  you  have  made 

January  1963 


in  the  laboratory  gives  you  a  valid  picture  of 
how  all  scientific  advances  begin.  It  permits  you 
to  engage  in  scientific  activity  and  thus,  to  some 
extent,  to  become  a  scientist  yourself. 

At  the  end  of  this  course  you  won't  know  all  of 
chemistry.  We  hope  that  you  will  know  enough 
chemistry  and  enough  about  science  to  feel  that 
the  part  you  don't  know  is  understandable,  not 
mysterious.  Perhaps  you  will  appreciate  the  great 
power  of  scientific  methods  and  appreciate  their 
limitations.  We  hope  that  you  will  have  become 
practiced  in  making  unexpected  observations,  in 
weighing  facts,  and  in  framing  valid  conclusions. 
We  hope  that  you  will  have  formed  the  habit  of 
questioning  and  of  seeking  understanding  rather 
than  being  satisfied  with  blind  acceptance  of 
dogmatic  assertions.  We  expect  that  you  will 
share  in  the  excitement  of  science  and  that  you 
will  feel  the  rich  pleasure  that  comes  with  dis- 
covery. If  most  of  these  hopes  are  fulfilled,  then 
you  have  had  an  optimum  introduction  to  sci- 
ence through  chemistry.  Nothing  could  be  a 
more  important  part  of  your  education  at  a  time 
when  science  is  molding  our  age. 

GEORGE    C.    PIMENTEL 

Editor  for  the  Chemical 

Education  Material  Study 


Foreword 


This  textbook  was  prepared  over  a  three  year 
period  by  a  group  of  university  and  high  school 
chemistry  teachers  under  a  grant  from  the  Na- 
tional Science  Foundation.  The  project,  called 
CHEM  Study,  was  organized  and  directed  on 
broad  policy  lines  by  a  Steering  Committee  of 
nationally  known  teachers  and  pre-eminent  sci- 
entists from  a  variety  of  chemical  fields.  The 
Steering  Committee,  headed  by  Nobel  Laureate 
Glenn  T.  Seaborg,  attempted  to  staff  the  study 
with  the  country's  most  able  university  scientists 
and  high  school  teachers.  The  university  profes- 
sors were  drawn  from  all  over  the  United  States 
on  the  basis  of  demonstrated  understanding  of 
science  and  recognized  leadership  in  teaching  it. 
The  names  of  the  contributors  to  this  text  al- 
ready appear  on  more  than  a  dozen  widely 
accepted  college  level  textbooks.  An  equal  num- 
ber of  outstanding  high  school  teachers  were 
named  as  contributors,  each  one  individually 
selected  on  the  basis  of  enthusiastic  recommen- 
dations by  his  peers.  These  teachers  participated 
in  every  phase  of  the  preparation  of  this  course. 
The  effort  of  these  highly  qualified  persons, 
totaling  over  fifteen  man-years,  is  summed  in  the 
CHEM  Study  course.  The  National  Science 
Foundation  deserves  commendation  for  making 
such  activities  possible;  never  before  has  such 
an  array  of  talent  been  assembled  to  construct 
a  high  school  chemistry  course. 


The  textbook,  CHEMISTR  Y—An  Experimen- 
tal Science,  is  designed  for  a  high  school  intro- 
ductory chemistry  course  and  it  is  meshed  closely 
with  an  accompanying  Laboratory  Manual  and  a 
set  of  pertinent  films.  A  comprehensive  Teachers 
Guide  is  available  to  aid  teachers  in  gaining 
familiarity  with  the  course.  The  first  editions  of 
the  textbook  and  laboratory  manual,  written 
during  the  summer  of  1960,  were  used  during 
1960-1961  in  23  high  schools  and  one  junior 
college  by  about  1300  students.  During  this  first 
year,  there  was  weekly  staff  contact  with  the 
pioneering  teachers.  On  the  basis  of  their  experi- 
ence, the  materials  were  revised  during  the 
summer  of  1961  and  the  Teachers  Guide  was 
written.  This  second  edition  was  used  in  123 
high  schools  and  3  junior  colleges  scattered  over 
the  country  and  involving  13,000  students.  Again 
the  closely  monitored  field  experience  founded 
the  third  and  final  revision.  The  course,  essen- 
tially in  the  form  presented  here,  was  used  during 
1962-1963  in  560  high  schools  in  46  states  by 
about  45,000  randomly  selected  students.  Its 
teachability  is  assured. 

The  title,  CHEMISTRY— An  Experimental 
Science,  states  the  theme  of  this  one  year  course. 
A  clear  and  valid  picture  of  the  steps  by  which 
scientists  proceed  is  carefully  presented  and  re- 
peatedly used.  Observations  and  measurements 
lead  to  the  development  of  unifying  principles 

vii 


VI11 


FOREWORD 


and  then  these  principles  are  used  to  interrelate 
diverse  phenomena.  Heavy  reliance  is  placed 
upon  laboratory  work  so  that  chemical  prin- 
ciples can  be  drawn  directly  from  student  ex- 
perience. Not  only  does  this  give  a  correct  and 
nonauthoritarian  view  of  the  origin  of  chemical 
principles,  but  it  gives  maximum  opportunity 
for  discovery,  the  most  exciting  part  of  scientific 
activity.  This  experimental  theme  is  supported 
by  a  number  of  films  to  provide  experimental 
evidence  that  is  needed  but  not  readily  available 
in  the  classroom  because  of  inherent  danger, 
rarity,  or  expense. 

The  initial  set  of  experiments  and  the  first  few 
textbook  chapters  lay  down  a  foundation  for  the 
course.  The  elements  of  scientific  activity  are 
immediately  displayed,  including  the  role  of  un- 
certainty. The  atomic  theory,  the  nature  of 
matter  in  its  various  phases,  and  the  mole  con- 
cept are  developed.  Then  an  extended  section  of 
the  course  is  devoted  to  the  extraction  of  im- 
portant chemical  principles  from  relevant  labo- 
ratory experience.  The  principles  considered 
include  energy,  rate  and  equilibrium  character- 
istics of  chemical  reactions,  chemical  periodicity, 
and  chemical  bonding  in  gases,  liquids,  and 
solids.  The  course  concludes  with  several  chap- 
ters of  descriptive  chemistry  in  which  the  ap- 
plicability and  worth  of  the  chemical  principles 
developed  earlier  are  seen  again  and  again. 

There  are  a  number  of  differences  from  more 
traditional  courses.  The  most  obvious  are,  of 
course,  the  shift  of  emphasis  from  descriptive 
chemistry  toward  chemical  principles  to  repre- 
sent properly  the  change  of  chemistry  over  the 
last  two  decades.  Naturally,  this  reconstruction 
of  the  entire  course  gives  a  unique  opportunity 
to  delete  obsolete  terminology  and  out-moded 
material.  Less  obvious  but  perhaps  more  im- 
portant is  the  systematic  development  of  the 
relationship  between  experiment  and  theory. 
Chemistry  is  gradually  and  logically  unfolded, 
not  presented  as  a  collection  of  facts,  dicta,  and 
dogma.  We  hope  to  convey  an  awareness  of  the 
significance  and  capabilities  of  scientific  activities 
that  will  help  the  future  citizen  assess  calmly  and 
wisely  the  growing  impact  of  technological  ad- 
vances on  his  social  environment.  Finally,  we 


have  striven  for  closer  continuity  of  subject 
matter  and  pedagogy  between  high  school  and 
modern  freshman  chemistry  courses  for  those 
students  who  will  continue  their  science  training. 

We  do  believe  that  the  CHEM  Study  course 
achieves  the  goals  we  have  set.  Experience  has 
shown  that  the  course  is  interesting  to  and  within 
the  grasp  of  the  average  high  school  chemistry 
student  and  that  it  challenges  and  stimulates 
the  gifted  student.  The  course  content  provides 
a  strong  foundation  for  the  college-bound  stu- 
dent. Inevitably  the  question  arises,  "Is  this 
course  better  than  (or,  as  good  as)  the  traditional 
one?"  An  answer  is  not  readily  found  in  com- 
parative tests.  A  CHEM  Study  student  might  be 
handicapped  in  a  test  that  has  little  emphasis 
upon  principles,  that  is  heavily  laden  with  de- 
scriptive "recall  questions,"  or  that  uses  obsolete 
terminology.  Conversely,  a  test  designed  specif- 
ically for  the  modern  CHEM  Study  course 
content  would  surely  prejudice  against  a  student 
with  a  traditional  preparation.  The  issue  can- 
not be  completely  resolved  "objectively"  be- 
cause value  judgments  are  ultimately  involved. 
Whether  the  CHEM  Study  goals  are  valid  and 
the  approach  is  reasonable  must  be  decided  with 
due  consideration  to  the  reported  experience  of 
teachers  and  to  the  credentials  of  those  who 
developed  the  materials. 

There  are  numberless  ways  in  which  CHEM 
Study  is  indebted  to  the  University  of  California 
and  to  Harvey  Mudd  College  for  contributions 
of  facilities,  personnel,  and  encouragement.  We 
acknowledge  with  thanks  the  stimulation  and 
support  we  have  received  from  the  National 
Science  Foundation.  Finally,  the  Staff  feels  a 
heavy  debt  of  gratitude  to  all  of  those  who 
participated  so  energetically  and  enthusiastically 
in  the  preparation  of  the  CHEM  Study  mate- 
rials. We  thank  the  Steering  Committee  for  their 
valued  and  helpful  guidance.  We  thank  the  con- 
tributors listed  on  the  title  page  for  their  dedica- 
tion of  time,  interest,  and  their  ample  talents  to 
this  effort.  We  acknowledge  especially  the  key 
roles  of  Mr.  Joseph  Davis,  Mr.  Saul  Geffner, 
Mr.  Keith  MacNab,  Miss  Margaret  Nicholson, 
and  Mr.  Harley  Sorensen.  These  individuals  not 
only  used  the  CHEM  Study  materials  in  the 


FOREWORD 


classroom  but  also  served  continuously  as  staff 
members.  Their  contributions  and  critiques  have 
greatly  increased  the  teachability  of  the  CHEM 
Study  course.  We  thank  the  many  teachers  who 
used  the  trial  editions  in  their  classrooms;  their 
careful  scrutiny  of  the  text  and  laboratory  man- 
ual and  their  many  valuable  suggestions  pro- 


vided a  firm  basis  for  revisions.  Finally,  we  thank 
the  many  students  who  labored  through  the  trial 
versions  of  CHEM  Study;  their  every  reaction — 
pain  or  pleasure,  enthusiasm  or  ennui,  spark  or 
sputter— was  noted  and  lent  to  the  improvement 
of  the  course. 


J.    ARTHUR    CAMPBELL, 

Director,  Chemical  Education  Material  Study 
Harvey  Mudd  College 

GEORGE   C.    PIMENTEL, 

Editor,  Textbook 
University  of  California 

Berkeley,  California 
January,  1963 


LLOYD    E.    MALM 

Editor,  Laboratory  Manual 
University  of  Utah 

A.    L.   MCCLELLAN 

Editor,  Teachers  Guide 

California  Research  Corporation 

DAVID   RIDGWAY 

Producer,  Films 


Acknowledgments 


Quotations  appearing  on  the  following  pages  are 
used,  with  permission,  from  the  indicated  sources. 

Page      1     History  of  Science,  W.   Dampier.  New 
York:  Cambridge  University  Press,  1949. 
17    Principia,  Isaac  Newton.  Mott's  transla- 
tion revised  by  F.  Cajori.  Berkeley:  Uni- 
versity of  California  Press,  1934,  p.  673. 
38     New   Systems   of  Chemical  Philosophy, 
John  Dalton.  Manchester,  England,  1810. 
49     Readings   in   the  Literature  of  Science, 
W.  C.  Dampier  and  M.  Dampier.  New 
York:  Harper  and  Row,  1959,  p.  100. 
65    Solutions,  W.  Ostwald.  London:  Long- 
mans, Green  and  Co.,  1891. 
Letter  by  J.  A.  R.  Newlands,  Chemical 
News,  Vol.  10,  1864,  p.  94. 
Chemical  Thermodynamics,  A  Course  of 
Study,  Frederick  T.  Wall.  San  Francisco: 
W.  H.  Freeman  and  Company,  1958,  p.  2. 

124  The  Drift  Toward  Equilibrium,  H.  Ey- 
ring,  from  Science  in  Progress,  Fourth 
Series,  edited  by  G.  A.  Baitsell,  New 
Haven:  Yale  University  Press,  1945,  p. 
169. 

142  Thermodynamics,  G.  N.  Lewis  and 
M.  Randall.  New  York:  McGraw-Hill 
Book  Co.,  Inc.,  1923,  p.  18. 

163  Solubility  of  Non-electrolytes,  J.  H.  Hilde- 
brand.  New  York:  Reinhold  Publishing 
Corp.,  1936,  p.  13. 


85 


108 


179  Elements  of  Chemistry,  A.  Lavoisier.  New 
York,  1806,  p.  14. 

199  Predictions  and  Speculation  in  Chemis- 
try, W.  M.  Latimer,  Chemical  and  Engi- 
neering News,  Vol.  31,  1953,  p.  3366. 

224  Textbook  of  Quantitative  Inorganic  Anal- 
ysis, I.  M.  Kolthoff  and  E.  B.  Sandell. 
New  York:  Macmillan,  1936,  p.  2. 

233  The  Rise  of  Scientific  Philosophy,  Hans 
Reichenbach.  Berkeley:  University  of 
California  Press,  1956,  p.  168. 

252  Valence,  C.  A.  Coulson.  New  York:  Ox- 
ford University  Press,  1961,  p.  3. 

274  Chemical  Analysis  by  Infrared,  Bryce 
Crawford,  Jr.,  New  York:  Scientific 
American,  Oct.  1953. 

300  The  Nature  of  the  Chemical  Bond, 
L.  Pauling.  Ithaca:  Cornell  University 
Press,  1939,  p.  422. 

321  Les  Prix  Nobel,  1947,  Nobel  lecture  by 
R.  Robinson.  Stockholm:  Norstedt  and 
Soner,  1947,  p.  110. 

421  From  Quantum  Chemistry  to  Quantum 
Biochemistry,  Alberte  Pullman  and  Ber- 
nard Pullman,  in  Albert  Szent-Gyoergyi 
and  Modern  Biochemistry,  edited  by  Rene 
Wurmser.  Paris:  Institute  of  Biology, 
Physics,  Chemistry,  1962. 

xi 


Xll 


ACKNOWLEDGMENTS 


436  Genesis  of  Life,  J.  B.  S.  Haldane,  in  The 
Earth  and  Its  Atmosphere,  edited  by  D.  R. 
Bates.  New  York:  Basic  Books,  Inc., 
1960. 

The  following  photographs  are  used  with  permission 
from  the  indicated  source. 

Frontispiece    The  Candle— Illuminating  Chemistry, 
by  Bernard  Abramson. 
Page      5    Ice  melting,  by  Ross  H.  McGregor. 
5     Aluminum  melting,  courtesy  Alumi- 
num Corporation  of  America. 
5    Solder  melting,  by  Charles  L.  Finance. 
48    G.  N.  Lewis,  courtesy  the  Hagemeyer 
Collection,  Bancroft  Library,  Univer- 
sity of  California. 
94    Cutting  potassium,  by  Charles  L.  Fi- 
nance. 
107     D.  Mendeleev,  courtesy  the  Univer- 
sity of  Leningrad. 
141     H.  Eyring,  courtesy  H.  Eyring. 
198    S.  Arrhenius,  courtesy  The  Bettmann 
Archive. 


299  L.  Pauling,  courtesy  The  California 
Institute  of  Technology. 

310  Network  silicates,  by  Charles  L.  Fi- 
nance. 

312  Sodium  chloride  crystals,  by  Charles 
L.  Finance. 

320  P.  Debye,  courtesy  Cornell  Univer- 
sity. 

351  R.  Robinson,  courtesy  Canadian  In- 
dustries Limited. 

386  A.  Stock,  courtesy  The  American 
Chemical  Society. 

420  G.  T.  Seaborg,  courtesy  California 
Research  Corporation,  Richmond 
Laboratory,  Richmond,  California. 

435  R.  B.  Woodward,  courtesy  The  Am- 
erican Chemical  Society. 

Color  plate     I    Elements  and  compounds,  by 
Charles  L.  Finance. 
II    Indicator   colors,    by    Charles   L. 

Finance. 
Ill    Spectrograph,  by  Charles  L. 
Finance. 


Contents 


Chapter  1.  Chemistry:  An  Experimental  Science  1 

2.  A  Scientific  Model:  The  Atomic  Theory  17 

3.  Chemical  Reactions  38 

4.  The  Gas  Phase:  Kinetic  Theory  49 

5.  Liquids  and  Solids:  Condensed  Phases  of  Matter  65 

6.  Structure  of  the  Atom  and  the  Periodic  Table  85 

7.  Energy  Effects  in  Chemical  Reactions  108 

8.  The  Rates  of  Chemical  Reactions  124 

9.  Equilibrium  in  Chemical  Reactions  142 

10.  Solubility  Equilibria  163 

11.  Aqueous  Acids  and  Bases  179 

12.  Oxidation-Reduction  Reactions  199 

13.  Chemical  Calculations  224 

14.  Why  We  Believe  in  Atoms  233 

15.  Electrons  and  the  Periodic  Table  252 

16.  Molecules  in  the  Gas  Phase  274 

17.  The  Bonding  in  Solids  and  Liquids  300 

18.  The  Chemistry  of  Carbon  Compounds  321 

19.  The  Halogens  352 

20.  The  Third  Row  of  the  Periodic  Table  364 

21.  The  Second  Column  of  the  Periodic  Table  377 

22.  The  Fourth-Row  Transition  Elements  387 

an 


XIV  CONTENTS 

23.  Some  Sixth-  and  Seventh-Row  Elements  41 1 

24.  Some  Aspects  of  Biochemistry:  An  Application  of  Chemistry  421 

25.  The  Chemistry  of  Earth,  the  Planets,  and  the  Stars  436 

Appendix  1.  A  Description  of  a  Burning  Candle  449 

2.  Relative  Strengths  of  Acids  in  Aqueous  Solution  451 

3.  Standard  Oxidation  Potentials  for  Half-Reactions  452 

4.  Names,  Formulas,  and  Charges  of  Some  Common  Ions  454 
Index  455 


THE    CANDLE  —  ILLUMINATING    CHEMISTRY 


CHAPTER 


l 


Chemistry: 

An  Experimental 

Science 


•    •    •    those  sciences  are  vain  and  full  of  errors  which  are  not  born  from 
experiment,  the  mother  of  all  certainty.    •    •    • 

LEONARDO   DA    VINCI,    1452-1519 


Many  words  have  been  spoken  and  written  in 
answer  to  the  questions: 

"What  is  the  nature  of  scientific  study?" 
"What  is  the  nature  of  chemistry?" 

We  shall  try  to  find  the  answers  in  this  course, 
not  through  words  alone,  but  through  experi- 
ence. No  one  can  completely  convey  through 
words  the  excitement  and  interest  of  scientific 


discovery.  Hence  we  shall  see  the  nature  of  sci- 
ence by  engaging  in  scientific  activity.  We  shall 
see  the  nature  of  chemistry  by  considering  prob- 
lems which  interest  chemists. 

Our  starting  point  will  be  based  on  examples 
of  the  activities  of  science,  rather  than  on  defini- 
tions. We  will  perform  these  activities,  beginning 
on  familiar  ground.  On  such  ground,  where  you 
know  the  answer,  you  will  best  see  the  steps  by 
which  science  advances. 


1-1    THE  ACTIVITIES  OF  SCIENCE 

Every  form  of  life  "feels"  its  surroundings  in  one 
way  or  another.  In  response  to  the  feel  of  the 
surroundings,  it  behaves  according  to  a  pattern 
which  tends  to  prolong  its  existence. 

A  tree  is  illuminated  by  the  morning  sunshine. 
In  response,  the  leaves  of  the  tree  turn  on  their 


stems  to  present  full  surface  to  the  light.  This 
movement  causes  the  leaves  to  intercept  more 
light,  and  light  is  the  source  of  energy  which  runs 
the  amazing  chemical  factory  operated  by  the 
tree.  The  tree  grows. 
A  bear  feels  that  summer  is  over — perhaps  by 

l 


chemistry:  an  experimental  science  I  CHAP.   1 


the  length  of  the  day  or  by  the  color  of  fall 
leaves,  perhaps  by  some  ursine  almanac  humans 
cannot  read.  In  response,  he  seeks  a  secluded 
spot  and  takes  a  winter-long  nap.  During  this 
hibernation,  his  blood  pressure  and  body  tem- 
perature drop,  his  digestion  closes  shop.  The 
bear  uses  the  minimum  energy  necessary  to  stay 
alive.  It  is  not  a  coincidence  that  this  occurs  dur- 
ing the  season  when  food  is  most  difficult  to  find 
and  the  weather  is  quite  unbearable. 

Of  all  living  things,  man  feels  his  surroundings 
and  responds  to  them  in  the  most  complex  way. 
He  is  more  curious  than  the  most  inquisitive 
kitten.  Through  his  intellect  he  uses  his  senses 
more  effectively  than  an  antelope  avoiding  a 
stalking  lion.  He  has  developed  communication 
far  beyond  the  warning  quack  of  a  sentry  duck 
or  the  mating  call  of  a  lonely  moose.  Man's  in- 
tellect, together  with  his  communicative  ability, 
permit  him  to  respond  to  his  environment  in 
uniquely  beneficial  ways.  He  accumulates  infor- 
mation about  his  surroundings,  he  organizes  this 
information  and  seeks  regularities  in  it,  he  won- 
ders why  the  regularities  exist,  and  he  transmits 
his  findings  to  the  next  generation.  These  are  the 
basic  activities  of  science: 

to  accumulate  information  through  observa- 
tion; 

to  organize  this  information  and  to  seek  regu- 
larities in  it; 

to  wonder  why  the  regularities  exist; 

to  communicate  the  findings  to  others. 

Fig.  1-1.  A  scientist  makes  careful  observations. 


So  the  activities  of  science  begin  with  observa- 
tion. Observation  is  most  useful  when  the  condi- 
tions which  affect  the  observation  are  controlled 
carefully.  A  condition  is  controlled  when  it  is 
fixed,  known,  and  can  be  varied  deliberately  if 
desired.  This  control  is  best  obtained  in  a  special 
locale — a  laboratory.  When  the  observation  is 
brought  under  careful  control,  it  is  dignified  by 
a  special  name — a  controlled  sequence  of  observa- 
tions is  called  an  experiment.  All  science  is 
built  upon  the  results  of  experiments. 


1-1.1    Observation  and  Description 

Everyone  thinks  of  himself  as  a  good  observer. 
Yet  there  is  much  more  to  it  than  meets  the  eye. 
It  takes  concentration,  alertness  to  detail,  in- 
genuity, and  often  just  plain  patience.  It  even 
takes  practice!  Consider  an  example  from  your 
own  experience.  Think  how  much  can  be  written 
about  an  object  as  familiar  as  a  burning  candle! 
Of  course,  it  takes  careful  observation — a  careful 
experiment.  This  means  the  candle  must  be  ob- 
served in  a  laboratory,  that  is,  in  a  place  where 
conditions  can  be  controlled.  But,  how  do  we 
know  which  conditions  need  be  controlled?  Be 
ready  for  surprises  here!  Sometimes  the  impor- 
tant conditions  are  difficult  to  discover.  Here  are 
some  conditions  that  are  important  in  some  ex- 
periments but  are  not  important  here. 

The  experiment  is  done  on  the  second  floor. 
The  experiment  is  done  in  the  daytime. 
The  room  lights  are  on. 

Here  are  some  conditions  that  might  be  impor- 
tant here. 

The  lab  bench  is  near  the  door. 
The  windows  are  open. 
You  are  standing  close  enough  to  the  candle 
to  breathe  on  it. 

Why  are  these  conditions  important?  Do  they 
have  something  in  common?  Yes,  there  is  the 
common  factor  that  a  candle  does  not  operate 
well  in  a  draft.  The  conditions  are  important 
because  they  influence  ihe  result  of  the  experi- 


SEC.     1-1    |    THE    ACTIVITIES    OF    SCIENCE 

ment.  Important  conditions  are  often  not  as 
easily  recognized  as  these.  A  good  experimen- 
talist pays  much  attention  to  the  discovery  of 
conditions  that  must  be  controlled.  His  success 
is  often  determined  by  his  ability  to  control  them. 
Review  your  own  description  of  a  burning 
candle  and  compare  your  essay  with  the  one  in 
Appendix  1.  How  many  of  your  observations 
are  included  there?  How  many  listed  in  the  ap- 
pendix are  not  in  your  description?  We  see  that 
the  burning  candle  is  a  complicated  and  fascinat- 
ing object  when  subjected  to  careful  observation 
and  detailed  description. 

1-1.2    The  Search  for  Regularities 

Observation  inevitably  leads  to  questions.  One  of 
the  first  questions  that  usually  arises  is  "What 
regularities  appear?"  The  discovery  of  regulari- 
ties permits  simplification  of  the  observations. 
Instead  of  each  observation  standing  alone,  sev- 
eral observations  can  be  classed  together  and, 
hence,  can  be  used  more  effectively. 

You  must  become  aware  of  the  pitfalls  that 
exist  in  the  search  for  regularities.  The  search  is 
a  meandering  one,  frequently  taking  wrong 
turns.  It  is  inherent  in  the  exploration  of  the 
unknown  that  not  every  step  is  an  advance.  Yet 
there  is  no  other  way  to  advance  than  by  taking 
steps.  How  the  search  proceeds  is  best  seen  in  a 
fable.  The  development  of  such  a  transparent 
example  may  help  you  see  how  a  scientist 
searches  for  regularities. 

Fable:  A  Lost  Child  Keeping  Warm 

Once  upon  a  time  a  small  child  became 
lost.  Because  the  weather  was  cold,  he  de- 
cided to  gather  materials  for  a  fire.  As  he 
brought  objects  back  to  his  campfire,  he 
discovered  that  some  of  them  burned  and 
some  of  them  didn't  burn.  To  avoid  collect- 
ing useless  substances,  the  child  began  to 
keep  track  of  those  objects  that  burned  and 
those  that  did  not.  (He  organized  his  infor- 
mation.) After  a  few  trips,  his  classification 
contained  the  information  that  is  shown 
in  Table  1-1. 


Table  1-1.  flammabilitv 

WILL  BURN  WON'T  BURN 


Tree  limbs 
Broom  handles 
Pencils 
Chair  legs 
Flagpoles 


Rocks 
Blackberries 
Marbles 
Paperweights 


This  organization  of  the  information  was 
quite  an  aid  in  his  quest  for  warmth.  How- 
ever, as  tree  limbs  and  broom  handles 
became  scarce,  the  child  tried  to  find  a 
regularity  that  would  guide  him  to  new 
burnable  materials.  Looking  at  the  pile  of 
objects  that  failed  to  burn  and  comparing 
it  with  the  pile  of  objects  that  would  burn, 
the  child  noticed  that  a  regularity  ap- 
peared. He  proposed  a  possible  "generali- 
zation." 

Perhaps:  "Cylindrical  objects  burn." 

This  procedure  is  one  of  the  elementary  logical 
thought  processes  by  which  information  is  sys- 
tematized. It  is  called  inductive  reasoning,  and 
it  means  that  a  general  rule  is  framed  on  the  basis 
of  a  collection  of  individual  observations  (or 
"facts").  Of  what  use  is  the  inductive  process? 
It  is  an  efficient  way  of  remembering. 

The  next  day  the  child  went  looking  for 
burnable  materials,  but  he  forgot  to  bring 
along  his  list.  However,  he  remembered  his 
generalization.  So,  he  returned  to  his 
hearthside  hauling  a  tree  limb,  an  old  cane, 
and  three  baseball  bats  (successful  predic- 
tions!). What's  more,  he  reflected  with 
pleasure  that  he  hadn't  bothered  to  carry 
back  some  other  objects:  an  automobile 
radiator,  a  piece  of  chain,  and  a  large  door. 
Since  these  objects  weren't  cylindrical  there 
was  no  reason  to  expect  them  to  burn. 

No  doubt  you  are  ready  to  complain  that  this 
generalization  isn't  really  true!  Quite  the  oppo- 
site! The  generalization  states  a  regularity  dis- 
covered among  all  the  observations  available, 
and  as  long  as  observations  are  restricted  to 
objects  in  the  list,  the  generalization  is  applicable. 
A  generalization  is  reliable  within  the  bounds 
defined  by  the  experiments  that  led  to  the  rule. 


chemistry:   an  experimental  SCIENCE  I  CHAP.   1 


As  long  as  we  restrict  ourselves  to  the  objects 
in  Table  1-1  (together  with  canes  and  baseball 
bats)  it  is  surely  true  that  all  of  the  cylindrical 
objects  burn! 


Fig.  1-2.  "Cylindrical  objects  burn." 


Because  of  his  successful  predictions,  the 
child  became  confident  of  his  generaliza- 
tion. The  next  day  he  deliberately  left  the 
list  at  his  campsite.  This  time,  with  the  aid 
of  his  rule,  he  came  back  heavily  laden  with 
three  pieces  of  pipe,  two  ginger  ale  bottles, 
and  the  axle  from  an  old  car,  while  spurn- 
ing a  huge  cardboard  box  full  of  news- 
papers. 

During  the  long  cold  night  that  followed 
he  drew  these  conclusions: 

(1)  The  cylindrical  shape  of  a  burnable  ob- 
ject may  not  be  intimately  associated 
with  its  fiammability  after  all. 

(2)  Even  though  the  "cylindrical"  rule  is 
no  longer  useful,  tree  limbs,  broom 
handles,  pencils,  and  the  other  burn- 
ables  in  Table  1-1  still  burn. 

(3)  He'd  better  bring  the  list  along  tomor- 
row. 

But,  thinking  over  the  longer  list,  he  saw 
a  new  regularity  that  fitted  Table  1-1  and 
the  newly  acquired  information  as  well: 

Perhaps:  "Wooden  objects  burn." 


What  good  is  this  rule  in  the  light  of  the  earlier 
disappointment?  Well,  it  caused  the  child  to  go 
back  and  get  that  door  he  had  passed  up  two 
days  earlier,  but  it  didn't  lead  him  to  go  after 
the  chain,  the  automobile  radiator,  or  the  card- 
board box  full  of  newspapers. 

Don't  think  this  is  facetious — it  is  exactly  what 
science  is  all  about!  We  make  some  observations, 
organize  them,  and  seek  regularities  to  aid  us  in 
the  effective  use  of  our  knowledge.  The  regulari- 
ties are  stated  as  generalizations  that  are  called 
theories.  A  theory  is  retained  as  long  as  it  is 
consistent  with  the  known  facts  of  nature  or  as 
long  as  it  is  an  aid  in  systematizing  our  knowl- 
edge. We  can  be  sure  that  some  day  a  number  of 
our  present  scientific  views  will  seem  as  absurd 
as  "Cylindrical  objects  burn."  But  on  that  day 
we  will  be  proud  of  better  views  that  have  been 
substituted.  If  you  are  discouraged  by  the  child's 
faltering  progress — he  hasn't  yet  decided  that  the 
box  of  newspapers  will  burn — be  reassured.  This 
child  is  a  scientist  and  his  faltering  steps  will  lead 
him  to  the  newspapers.  They  are  the  same  steps 
that  led  us  to  our  present  understanding  of  rela- 
tivity, to  our  discovery  of  polio  vaccine,  and  to 
our  propulsion  of  rockets  to  the  moon. 

A    GENERALIZATION    ABOUT    THE 
MELTING    OF    SOLIDS 

Through  experiment,  you  yourself  have  discov- 
ered an  important  regularity  in  the  behaviors  of 
solid  substances. 

A  solid  melts  to  a  liquid  when  the  temperature 
is  raised  sufficiently.  The  temperature  at  which 
a  solid  melts  is  characteristic.  When  the  warm 
liquid  is  recooled,  it  solidifies  at  this  same 
temperature. 

This  generalization  is  of  great  value.  It  is  based 
upon  exactly  the  type  of  experiment  you  have 
performed.  We  have  confidence  in  the  rule  be- 
cause this  type  of  experiment  has  been  conducted 
successfully  on  hundreds  of  thousands  of  sub- 
stances. The  melting  behavior  is  one  of  the  most 
commonly  used  methods  of  characterizing  a  sub- 
stance. It  leads  us  to  wonder  if  every  solid  can 
be  converted  to  a  liquid  if  the  temperature  is 
raised  sufficiently.  Further,  it  leads  us  to  wonder 


SEC.     1-1    |    THE    ACTIVITIES    OF    SCIENCE 


Fig.  1-3.  A  solid  melts  to  a  liquid  at  a  characteristic 
temperature. 


if  every  liquid  can  be  converted  to  a  solid  if  the 
temperature  is  lowered  sufficiently. 

SOMB    TERMINOLOGY 

We  have  discovered  that  a  solid  can  be  converted 
to  a  liquid  by  warming  it  at  or  above  its  melting 
point.  Then  the  solid  can  be  restored  merely  by 
recooling.  The  solid  and  the  liquid  are  similar 
in  many  respects  and  one  is  easily  obtained  from 
the  other.  Hence  they  are  called  different  phases 
of  the  same  substance.  Ice  is  the  solid  phase  of 
water  and,  at  room  temperature,  water  is  in  the 
liquid  phase.  The  change  that  occurs  when  a 
solid  melts  or  a  liquid  freezes  is  called  a  phase 
change. 

1-1.3    Wondering  Why 

We  have  already  experienced  some  of  the  activi- 
ties of  science.  First  came  careful  observation 
under  controlled  conditions,  then  organization 
of  the  information  and  the  search  for  regularities 
of  behavior.  There  is  one  more  activity  that,  like 
dessert,  fittingly  comes  last.  This  activity  may  be 


called  "wondering  why,"  and  it  arises  from  our 
irresistible  urge  to  know  more  than  merely 
"What  happens?"  We  must  also  seek  the  answer 
to  "Why  does  it  happen?"  This  activity  is  prob- 
ably the  most  creative  and  the  most  rewarding 
part  of  a  science.  What  is  the  process?  What  does 
it  mean  to  answer  a  question  beginning  "Why"? 

EXPLANATIONS 

Let  us  see  what  it  means  to  search  for  an  expla- 
nation. Consider  a  child  blowing  up  a  balloon. 
As  he  blows  into  the  balloon  again  and  again  it 
expands  and  it  becomes  "harder."  Evidently  the 
gas  is  "pushing"  on  the  inside  of  the  balloon, 
stretching  its  elastic  walls.  Why  does  the  gas  push 


chemistry:   an  experimental  SCIENCE  I  CHAP.    1 


outward  more  and  more  on  the  walls  of  the 
balloon  as  it  is  inflated?  Why  does  the  gas  con- 
tinue to  push  outward  without  "tiring"  or  "run- 
ning down"?  These  are  "wondering  why"  ques- 
tions. 

There  are  two  ways  to  proceed  in  trying  to 
answer  these  questions.  We  have  already  ex- 
amined one  of  these  ways— to  look  more  closely 
at  the  balloon,  to  record  carefully  what  we  see, 
and  to  seek  regularities  in  what  we  observe.  The 
second  way  is  to  look  away  from  the  balloon  and 
to  seek  similar  behavior  in  another  situation  that 
we  understand  better.  Maybe  this  will  enable  us 
to  frame  an  explanation  of  the  gas  pressure  in 
terms  of  the  better-understood  situation.  Some- 
times a  useful  explanation  turns  up  in  a  quite 
unexpected  direction. 

Consider  the  motion  of  a  billiard  ball.  After 
it  is  struck  by  the  cue,  it  moves  until  it  strikes  a 
cushion,  from  which  it  bounces,  apparently  with 
undiminished  velocity.  It  rolls  along  in  a  new 
direction  until  it  strikes  another  cushion,  chang- 
ing its  direction  again.  It  may  continue  to  roll 
until  it  has  hit  the  cushions  six  or  seven  times. 
The  billiard  ball  seems  almost  tireless  as  it  re- 
bounds time  and  again  from  the  "walls"  of  the 
billiard  table.  Could  there  be  a  connection  be- 
tween the  "untiring"  motion  of  a  billiard  ball 
and  the  "untiring"  pressure  of  a  gas  in  a  balloon? 

Billiard  balls  have  long  fascinated  both  idle 
and  curious  men.  The  latter  group  has  found 
that  the  motion  of  a  billiard  ball  can  be  described 


Fig.  1-4.  A  rebounding  billiard  ball  suggests  a  possi- 
ble explanation  of  gas  pressure. 


by  assuming  each  collision  with  a  cushion  is 
perfectly  elastic.  When  the  ball  strikes  the  cush- 
ion, pushing  on  it,  the  cilshion  pushes  back,  and 
the  ball  leaves  again  without  any  loss  of  velocity. 
Its  movement  can  be  reasonably  foretold  on  the 
basis  of  this  elastic  collision  assumption.  Perhaps 
the  behavior  of  a  gas  can  be  explained  in  these 
same  terms.  Suppose  we  picture  a  gas  as  a  col- 
lection of  particles  bouncing  around  in  a  con- 
tainer, endlessly,  making  elastic  collisions  with 
the  walls,  just  like  miniature  billiard  balls.  Every 
time  one  of  these  particles  strikes  a  wall,  it 
"pushes"  on  the  wall  and  bounces  away  again. 
If  there  are  many  particles,  there  will  be  many 
such  collisions  per  second,  which  accounts  for 
the  pressure  of  the  gas.  If  gas  is  added  to  the 
balloon,  there  will  be  even  more  particles,  hence 
more  wall  collisions  per  second,  hence  higher 
pressure.  Thus  the  billiard  ball  model  does  offer 
a  possible  answer  to  our  question. 

With  this  example,  we  can  now  see  the  mean- 
ing of  an  explanation.  It  began  with  a  "wonder- 
ing why"  question: 

Question:  Why  does  a  balloon  expand  as  it  is 
inflated? 

Possible  Answer:  Perhaps  the  gas  put  in  the 
balloon  consists  of  a  collection  of  small  par- 
ticles that  rebound  from  the  wall  of  the  bal- 
loon just  as  billiard  balls  rebound  from  the 
cushions  of  a  billiard  table.  As  the  gas  particles 
rebound  from  the  balloon  wall,  they  push  on 
it.  When  more  gas  particles  are  added,  the 
number  of  such  wall  collisions  per  second  in- 
creases, hence  the  outward  push  on  the  bal- 
loon wall  increases.  The  balloon  expands. 


SEC.     1-1    |    THE    ACTIVITIES    OF    SCIENCE 


This  is  the  characteristic  pattern  of  an  explana- 
tion. It  begins  with  a  "Why?"  question  that  asks 
about  a  process  that  is  not  well  understood.  An 
answer  is  framed  in  terms  of  a  process  that  is 
well  understood.  In  our  example,  the  origin. of 
gas  pressure  in  the  balloon  is  the  process  we  wish 
to  clarify.  It  is  difficult  even  to  sense  the  presence 
of  a  gas.  The  air  around  us  usually  cannot  be 
seen,  tasted,  nor  smelled  (take  away  smog);  it 
cannot  be  heard  or  felt  if  there  is  no  wind.  So  we 
attempt  to  explain  the  properties  of  a  gas  in 
terms  of  the  behavior  of  billiard  balls.  These 
objects  are  readily  seen  and  felt;  their  behavior 
has  been  thoroughly  studied  and  is  well  under- 
stood. 

The  search  for  explanation  is,  then,  the  search 
for  likenesses  that  connect  the  system  under 
study  with  a  model  system  which  has  been 
studied  earlier.  The  explanation  is  considered  to 
be  "good"  when: 

(1)  the  model  system  is  well  understood  (that  is, 
when  the  regularities  in  the  behavior  of  the 
model  system  have  been  thoroughly  ex- 
plored); and 

(2)  the  connection  is  a  strong  one  (that  is,  when 
there  are  close  similarities  between  the  stud- 
ied system  and  the  model  system). 

Our  example  constitutes  a  good  explanation, 
because: 

(1)  how  a  billiard  ball  rebounds  is  well  under- 
stood; we  can  calculate  in  mathematical  de- 
tail just  how  much  push  the  billiard  ball 
exerts  on  the  cushion  at  each  bounce;  and 

(2)  the  connection  to  gas  pressure  is  close;  ex- 
actly the  same  mathematics  describes  the 
pressure  behavior  if  the  gas  is  pictured  as  a 
collection  of  many  small  particles,  endlessly 
in  motion,  bouncing  elastically  against  the 
walls  of  the  container. 

Therefore,  the  particle  explanation  of  gas  pres- 
sure is  a  good  one. 

Now,  perhaps,  you  can  see  that  answering  the 
question  "Why?"  is  merely  a  highly  sophisti- 
cated form  of  seeking  regularities.  It  is  indeed  a 
regularity  of  nature  that  gases  and  billiard  balls 
have  properties  in  common.  The  special  creativ- 


ity shown  in  the  discovery  of  this  regularity  is 
that  the  likeness  is  not  readily  apparent.  Fit- 
tingly, there  is  a  special  reward  for  the  discovery 
of  such  hidden  likeness.  The  discoverer  can  bring 
to  bear  on  the  studied  system  all  of  the  experience 
and  knowledge  accumulated  from  the  well  un- 
derstood system. 

HEATING    STEEL    WOOL 

These  ideas  are  best  learned  by  using  them  your- 
self. Let  us  take  an  example  from  your  own  labo- 
ratory work.  You  have  observed  the  heating  of  a 
variety  of  solid  materials:  sulfur,  wax,  tin,  lead, 
silver  chloride,  and  copper.  Each  melts  at  a  char- 
acteristic temperature.  This  fact  led  us  to  the 
generalization  "A  solid  melts  to  a  liquid  at  a 
characteristic  temperature."  Your  confidence  in 
the  generalization  was  bolstered  by  the  addi- 
tional information  (communicated  to  you,  not 
experienced)  that  this  applies  to  "hundreds  of 
thousands  of  substances." 

Your  teacher  demonstrated  the  effect  of  heat- 
ing steel  wool,  and  it  proved  to  be  a  spectacular 
exception  to  the  generalization  about  melting. 
An  inexperienced  observer  might  note  this  spe- 
cial behavior  in  his  notebook,  cross  reference  it 
under  "Sparklers,"  and  proceed  to  the  next  sub- 
stance. A  curious  person,  however,  cannot  resist 
wondering,  "Why  does  steel  wool  behave  in  this 
special  manner?" 

In  quite  another  direction  of  thought  you  in- 
vestigated the  burning  of  a  candle.  You  dis- 
covered that  air  plays  a  role  and  that  the  prod- 
ucts are  different  from  the  starting  materials. 

Here  we  have  two  areas  that  have  been  in- 
vestigated thoroughly  and  that  are  well  under- 
stood (and  that  you  have  examined  to  some  ex- 
tent yourself).  Stated  briefly,  our  starting  point 
of  knowledge  is: 

(1)  Solids  melt  on  heating. 

(2)  A  candle  burns,  consuming  oxygen  from  the 
air. 

We  are  wondering  why  steel  wool  makes  such 
a  brilliant  display  on  heating.  Perhaps  the  ex- 
planation is  to  be  found  in  2.  Though  there  are 
striking  visual  differences,  perhaps  we  can  ex- 
plain the  behavior  of  steel  wool  by  likening  it  to 


chemistry:   an  experimental  science     chap.   1 


t 


Steel  wool   hums 
■when    heated  in  air- 


Fig.  1-5.  The  behavior  of  steel  wool  on  heating. 


a  candle.  Can  we  substitute  the  words  "steel 
wool"  for  "candle"  in  2? 

Perhaps:  (3)  Steel  wool  burns,  consuming 
oxygen  from  the  air. 

If  3  is  a  useful  connection  with  the  behavior 
of  a  burning  candle,  then  we  should  enjoy  the 
special  reward  for  this  discovery.  The  connection 
we  propose  implies  that  the  knowledge  accumu- 
lated on  the  candle  can  be  brought  to  bear  on 
the  new  system. 

Since:  A  candle  does  not  burn  if  it  is  denied 
oxygen. 

then:  steel  wool  should  not  sparkle  if  it  is 
denied  oxygen. 


Steel  wool  melts  to 

a    liquid    -when 
heated  in  absence 
of  oxygen. 


dioutd& 


Here  is  a  proposal  subject  to  direct  test.  We 
can  place  steel  wool  under  an  atmosphere  that 
excludes  oxygen  gas,  and  look  for  a  change  in 
the  heating  behavior.  We  go  to  the  laboratory, 
heat  steel  wool  under  carbon  dioxide  and,  lo,  the 
steel  wool  melts! 

The  special  behavior  of  steel  wool  on  heating 
can  now  be  said  to  be  explained: 


'Steel  wool,  like  a  candle,  burns  when  heated 


in  air. 

"Steel  wool,  like  other  solids,  melts  to  a  liquid 
when  heated  under  conditions  that  prevent  burn- 
ing." 

This  type  of  understanding  makes  possible  the 
metallurgical  processing  of  steel  (and  other  met- 
als). This  type  of  reasoning  makes  possible  an 
increase  in  our  perception  of  the  regularities  of 
nature.  //  begins  with  wondering  why. 


1-2     UNCERTAINTY  IN  SCIENCE 

Here  are  three  statements  concerning  the  melting 
behavior  of  /wadichlorobenzene: 

(1)  The  melting  point  is  53°C. 

(2)  The  melting  point  is  53.2°C. 

(3)  The  melting  point  is  53.203°C. 

Apparently  the  third  statement  says  more  than 
the  second,  and  the  second  statement  says  more 
than  the  first.  Would  it  surprise  you  to  learn  that 


the  second  statement  could  be  the  most  informa- 
tive of  the  three?  It  is  so.  To  understand  why.  we 
must  consider  uncertainty  in  measurement. 

1-2.1     Uncertainty  in  a  Measurement 

A  scientific  statement  conveys  knowledge  about 
the  environment.  The  statement  is  careless  if  it 
says  less  than  is  known.  It  is  misleading  if  it  says 


SEC.    1-2    I    UNCERTAINTY    IN    SCIENCE 


more  than  is  known.  The  most  accurate  statement 
clearly  conveys  just  what  is  known  and  no  more. 
Thus,  a  scientist  will  decide  whether  to  list  the 
melting  point  as  53°C,  53.2°C  or  53.203°C  by 
considering  which  value  tells  just  what  is  known 
about  the  melting  behavior  and  no  more. 

Consider  your  own  laboratory  measurement 
of  the  melting  point  of  paradichlorobenzene 
(Experiment  3).  Do  your  temperature  measure- 
ments permit  you  to  say  that  the  melting  point 
is  53°C,  not  54°C?  Probably,  yes.  It  isn't  too 
difficult  to  read  the  thermometer  with  this  cer- 
tainty. Can  you  read  the  thermometer  so  as  to 
distinguish  53.0°C  and  53.2°C?  This  is  more  dif- 
ficult. It  depends  upon  the  thermometer  and 
your  skill  in  its  use.  It  also  depends  upon  whether 
the  temperature  of  the  solid  during  melting  is 
uniform  throughout.  Still,  a  magnifying  glass 
permits  more  certainty  in  the  reading  of  the 
scale,  and  slower  heating  would  increase  the  uni- 
formity of  the  sample  temperature.  With  this 
type  of  extra  care,  it  is  possible  to  tell  that  the 
melting  point  of  /raradichlorobenzene  is  53.2°C, 


Fig.  1-6.  Every  measurement  involves  some  uncer- 
tainty. The  thermometer  on  the  left  can  be 
read  to  ±0.2°C.  The  one  on  the  right  can 
be  read  to  ±0.002°C. 


not  53.0°C.  Consider,  however,  the  chance  of  re- 
fining your  thermometry  sufficiently  to  distin- 
guish between  53.200°C  and  53.203°C.  With  the 
equipment  available  to  you,  it  just  isn't  possible. 
We  conclude,  then,  that  a  careful  measurement 
could  establish  the  melting  point  to  be  53.2°C. 
Then  the  second  statement  (m.p.  =  53.2°C) 
would  tell  just  what  is  known.  The  first  state- 
ment (m.p.  =  53°C)  does  not  tell  all  that  is 
known  since  only  two  figures  are  given,  though 
three  were  measured.  The  third  statement 
(m.p.  =  53.203°C)  tells  far  more  than  is  known 
since  the  last  two  figures  were  not  learned  ex- 
perimentally. We  see  that  the  measurement  gives 
us  three  figures  which  are  meaningful  and  sig- 
nificant. The  number  53.2°C  is  said  to  have  three 
significant  figures. 


oat 


10 


chemistry:    an  experimental  SCIENCE  I  CHAP.    1 


1-2.2    Uncertainty  in  a  Derived  Quantity: 
Addition  and  Subtraction 

Results  of  scientific  observations  are  often  com- 
bined. For  example,  in  Experiment  5  you  will 
determine  the  change  of  water  temperature  dur- 
ing the  combustion  of  a  candle  (or  during  the 
solidification  of  candle  wax).  The  change  of  tem- 
perature, which  we  called  At,  is  the  result  of  two 
measurements,  not  just  one — it  is  a  derived  quan- 
tity: 

Temperature  after  heating  =  38.5°C 

Temperature  before  heating  =    9.3°C 

Difference  (temperature  change),  At  =  29.2°C 

In  accordance  with  good  scientific  practice,  we 
would  like  to  express  the  temperature  change  so 
as  to  include  just  what  is  known  but  no  more. 
To  do  this,  we  must  investigate  how  the  uncer- 
tainties in  the  two  temperature  readings  fix  the 
uncertainty  in  the  difference,  At.  Suppose  the 
final  temperature  is  measured  by  a  second  stu- 
dent who  finds  the  temperature  to  be  38.3°C. 
Then  a  third  student  finds  it  to  be  38.7°C.  Ap- 
parently, different  students  making  the  same 
measurement  may  differ  by  a  few  tenths  of  a 
degree.  Apparently,  the  measurement  recorded 
by  any  one  student  could  be  in  error  by  this 
amount,  about  0.2°C.  Perhaps  the  temperature 
recorded  as  38.5°C  could  be  as  high  as  38.7°C 
(0.2°  higher)  or  as  low  as  38.3°C  (0.2°  lower)! 
This  can  be  expressed  briefly  as  follows: 

Temperature  after  heating  =  38.5  ±  0.2°C* 

Presumably,  the  same  uncertainty  is  present  in 
the  first  temperature  measurement,  so  our  cal- 
culation becomes: 


Temperature  after  heating 
Temperature  before  heating 


=  38.5  ±  0.2°C 
=    9.3  ±  0.2°C 


Difference  (temperature  change),  At  =  29.2  ±  ??°C 

To  decide  what  uncertainty  to  ascribe  to  29.2, 
consider  the  worst  possible  combination  of  the 
uncertainties.  The  first  temperature  could  be  as 
low  as  9.1°C  and  the  final  temperature  could  be 
as  high  as  38.7°C.  Then  the  difference  would  be 
29.6°C.  Thus  the  worst  possible  combination  of 
errors  would  place  an  error  in  the  difference 

*  The  symbol  ±  is  read  "plus  or  minus." 


equal  to  the  sum  of  the  uncertainties  in  the  parts, 
0.2  +  0.2  =  0.4.  Hence  our  derived  result  can 
be  written 

Difference  (temperature  change)  =  29.2  ±  0.4°C 

We  see  that  the  uncertainty  in  a  derived  quantity 
is  fixed  by  the  uncertainties  in  the  measurements 
that  must  be  combined.  For  an  addition  or  a 
subtraction,  the  maximum  uncertainty  is  simply 
the  sum  of  the  uncertainties  in  the  components: 
0.2  +  0.2  =  0.4. 


EXERCISE  1-1 

In  Experiment  5,  the  weight  of  a  sample  of  water 
is  determined  by  subtracting  the  weight  of  the 
empty  can  from  the  weight  of  the  can  containing 
the  water. 

(wt.  water)  =  (wt.  can  +  water)  —  (wt.  empty  can) 

Suppose  the  weight  of  the  can  is  61  ±  1  grams 
and  the  weight  of  the  can  plus  water  is  406  ±  1 
grams.  Calculate  the  weight  of  the  water  and  the 
maximum  uncertainty  in  the  weight  caused  by 
the  uncertainties  in  each  of  the  two  weighings. 


1-2.3     Uncertainty  in  a  Derived  Quantity: 
Multiplication  and  Division 

The  temperature  measurements  made  in  Experi- 
ment 5  enable  you  to  calculate  the  quantity  of 
heat  liberated  when  a  known  weight  of  candle  is 
burned.  Heat  is  measured  in  units  called  calories; 
one  calorie  is  the  heat  needed  to  raise  the  tem- 
perature of  one  gram  of  water  one  degree  Centi- 
grade.^ To  raise  the  temperature  of  two  grams 
of  water  one  degree  would  require  two  calories — 
10  grams  would  require  10  calories.  In  general, 

quantity  of  heat  necessary  to  raise  w  grams 
of  water  one  degree 

=  w  calories 

But  in  the  experiment,  the  temperature  of  the 

t  The  heat  necessary  to  raise  the  temperature  of  one 
gram  of  water  one  degree  is  constant  within  ±0.2% 
between  8  and  80°C.  The  calorie  has  been,  in  the  past, 
defined  to  be  the  heat  necessary  to  raise  the  temperature 
of  one  gram  of  water  from  14.5  to  15.5°C. 


SEC.     1-2    I    UNCERTAINTY    IN    SCIENCE 


11 


water  rises  several  degrees — we  call  the  tempera- 
ture change  At.  If  it  takes  one  calorie  to  warm 
one  gram  of  water  one  degree  Centigrade,  it 
takes  five  calories  to  warm  it  five  degrees.  In 
general, 

quantity  of  heat  necessary  to  raise  w  grams 
of  water  At  degrees 

=  w  X  At  calories    (/) 

Again  we  are  faced  with  the  dual  problem:  first, 
to  calculate  the  quantity  of  heat,  q;  and,  second, 
to  decide  the  uncertainty  in  q. 

The  quantity  of  heat,  q,  is  calculated  with  the 
aid  of  equation  (7);  it  is  simply  the  product  of  the 
weight  of  water  times  the  temperature  difference. 
Referring  to  our  data,  we  find, 

wt.  of  water  =  345  ±  2  grams 
At  =  29.2  ±  0.4°C 

The  quantity  of  heat  is  equal  to  the  product, 

345 
X  29.2 
690 
3105 
690 


10,074.0  calories 
q  =  10,074.0  ±  ?  calories 

As  in  the  last  section,  we  can  estimate  the 
uncertainty  in  q  by  considering  the  worst  possible 
combination  of  uncertainties.  Suppose  the  weight 
is  actually  343  grams  and  At  is  actually  28.8°. 
Then  the  product  would  be  below  10,074.0  calo- 
ries. But  perhaps  the  weight  is  actually  347  grams 
and  At  is  actually  29.6°.  Then  the  product  would 
be  above  10,074.0  calories.  These  extremes  deter- 
mine the  uncertainty  in  the  product: 


Minimum  value 

Maximum  value 

343 

347 

X  28.8 

X29.6 

2744 

2082 

2744 

3123 

686 

694 

9878.4  calories 

10,271.2  calories 

We  see  that  the  product  10,074.0  could  be  in 
error  by  almost  200  calories.  Now  we  can  express 
our  result  together  with  its  uncertainty: 

q  =  10,074.0  ±  about  196  calories 


In  view  of  this  large  uncertainty,  we  can  round 
off  the  answer  sensibly: 

q  =  10,100  ±  200  calories 

Once  again  the  uncertainty  in  the  product,  a 
derived  quantity,  is  fixed  by  the  uncertainties  in 
the  measurements  that  must  be  combined. 

EXERCISE  1-2 

Calculate  the  uncertainty  in  the  product  w  X  At 
caused  by  the  temperature  measurement  alone 
(assuming  the  uncertainty  in  the  345  gram  weight 
of  water  has  been  made  negligible  by  more  care- 
ful weighings).  Calculate  the  uncertainty  caused 
by  the  ±2  gram  weighing  uncertainty  alone 
(assuming  the  uncertainty  in  the  temperature 
change,  29.2°C,  has  been  made  negligible  by  use 
of  a  more  sensitive  thermometer).  Compare  these 
two  contributions  to  the  total  uncertainty,  about 
200  calories. 


The  uncertainty  in  the  product,  ±200  calo- 
ries, is  not  simply  the  sum  of  the  uncertainties 
in  the  factors,  ±0.4°C  and  ±2  grams.  Instead, 
the  sum  of  the  percentage  uncertainties  in  the 
factors  determines  the  uncertainty  in  a  product 
or  a  quotient.*  Fortunately,  there  is  an  easy 
method  for  estimating  it  roughly  without  cal- 
culating percentages.  This  method,  based  upon 
the  number  of  figures  written,  is  described  in 
Section  1-2.5. 

1-2.4    The  Absence  off  Certainty  in  Science 

Each  measuring  device  has  limitations  that  fix  its 
accuracy.  Hence  every  individual  observation 
has  some  uncertainty  associated  with  it.  Since 
every  regularity  of  nature  is  discovered  through 
observations,  every  regularity  (law,  rule,  theory) 
has  uncertainty  attached  to  it. 

Every  scientific  statement  involves  some 
uncertainty. 

A  corollary: 

No  scientific  statement  is  absolutely  certain. 

*  The  calculation  based  upon  percentage  uncertainty  is 
presented  in  Appendix  4  of  the  Laboratory  Manual. 


12 


chemistry:   an  experimental  SCIENCE  I  CHAP.   1 


1-2.5    How  Uncertainty  Is  Indicated 

We  now  have  seen  two  methods  for  indicating 
uncertainty  in  a  number.  The  most  informative 
is  to  follow  the  number  by  the  symbol  db  and 
then  the  best  estimate  available  of  the  uncer- 
tainty. The  less  informative  but  more  widely  used 
method  is  to  indicate  crudely  the  uncertainty  by 
the  number  of  figures  shown.  The  last  figure 
shown  is  generally  the  one  in  which  there  is  some 
uncertainty.  Thus,  the  number  53.2°C  indicates 
there  may  be  uncertainty  in  the  figure  2,  but 
there  is  none  in  either  of  the  figures  5  or  3.  The 
digits  that  are  certain  and  one  more  are  called 
significant  figures.  The  correct  number  of  sig- 
nificant figures  should  always  be  used  and,  wher- 
ever possible,  the  more  definite  indication  ± 
should  be  added. 

We  need  convenient  rules  for  estimating  the 
maximum  uncertainty  in  derived  quantities.  This 
is  rather  easy  for  a  sum  or  a  difference.  In  either 
case,  merely  add  up  the  uncertainties  in  the 
components.  Fortunately  there  is  an  equally  sim- 


ple rule  for  estimating  roughly  the  uncertainty 
in  a  product  or  quotient.  A  product  (a  X  b)  or 
quotient  (a/b)  has  the  same  number  of  significant 
figures  as  the  less  precise  component  (a  or  b).* 


EXERCISE  1-3 

In  Section  1-2.3  we  multiplied  345  X  29.2  to 
obtain  10,074.0  calories. 

(a)  How  many  significant  figures  are  there  in  the 
factor  345?  In  the  factor  29.2? 

(b)  How  many  significant  figures  should  be  re- 
tained in  the  product,  10,074.0? 

(c)  Six  figures  are  specified  in  the  number 
10,074.0 — more  than  are  warranted.  "Round 
off"  this  number  in  accordance  with  your 
answer  to  (b).  Compare  your  answer  with 
the  final  result  derived  in  Section  1-2.3, 
q  =  10,100  ±  200  calories. 


1-3    COMMUNICATING  SCIENTIFIC  INFORMATION 


One  of  the  most  important  reasons  for  man's 
progress  in  understanding  and  controlling  his  en- 
vironment is  his  ability  to  communicate  knowl- 
edge to  the  next  generation.  It  isn't  necessary  for 
each  twentieth  century  scientist  to  invent  the 
atomic  description  of  matter.  This  was  invented 
by  John  Dalton  in  the  nineteenth  century,  and 
Dalton  recorded  his  ideas  in  the  scientific  litera- 
ture together  with  the  observations  that  led  him 
to  the  model.  By  study  of  this  and  subsequent 
literature  a  modern  scientist  can  appraise  the 
nature  of  the  description,  the  facts  it  will  explain, 
and  the  limitations.  He  is  quickly  able  to  ap- 
proach the  frontiers  of  knowledge — the  frontiers 
defined  by  the  limitations  in  our  accepted  models 
of  the  behavior  of  matter. 

One  can  almost  say  that  a  scientific  advance  is 
important  only  if  it  is  communicated  to  others. 
If  Dalton  had  not  told  others  of  his  ideas  nor 
tried  to  convince  them  (through  his  supporting 


arguments),  then  someone  else  would  have  had 
to  do  it  all  again. 

Communication  of  knowledge  is,  then,  an  im- 
portant part  of  scientific  activity.  The  first  re- 
quirement is  good  use  of  language.  If  an  idea  is 
not  well  expressed,  whether  orally  or  in  writing, 
it  is  not  likely  to  be  clearly  understood.  An  argu- 
ment loses  its  force  if  it  is  stated  in  an  ambiguous 
way.  An  essential  thought  can  be  lost  in  a  maze 
of  excess  words.  Choose  and  use  your  language 
carefully. 

The  manner  in  which  you  present  an  idea  de- 
pends to  some  extent  upon  the  intended  use,  to 
some  extent  upon  the  type  of  information  avail- 
able. Usually,  the  more  precise  the  statement  of 
the  regularity,  the  more  valuable  it  is.  In  general, 
there  will  be  more  than  one  way  to  express  a 
generalization,  and  judgment  is  needed  in  mak- 

*  The  use  of  significant  figures  is  discussed  in  Appendix 
4  in  the  Laboratory  Manual. 


SEC.    1-3    I    COMMUNICATING    SCIENTIFIC    INFORMATION 


13 


Fig.  1-7.  A  regularity  in  the  behavior  of  a  fixed 
amount  of  gas.  As  pressure  rises,  volume 
decreases. 


ing  a  choice  among  them.  This  is  best  seen  in 
terms  of  an  example. 

Figure  1-7  shows  a  tire  pump  with  a  pressure 
gauge  attached  to  the  hose  so  that  the  gauge 
prevents  escape  of  the  gas  in  the  pump.  Pushing 
on  the  handle  of  the  pump  causes  the  plunger  to 
go  down,  and  reduces  the  volume  occupied  by 
the  gas.  The  gauge  shows  that  the  pressure  is 
higher.  Pushing  still  harder  on  the  handle  in- 
creases the  pressure  still  more.  Again  the  increase 
in  pressure  causes  a  reduction  in  volume.  We  see 
that  as  the  pressure  rises,  the  volume  decreases. 
This  qualitative  statement  describes  a  regularity 
in  the  behavior  of  a  fixed  amount  of  gas.  Such  a 
qualitative  statement  is  the  crudest  form  in  which 
a  regularity  can  be  expressed. 

A  curious  person,  attempting  to  understand 
this  regularity,  might  see  a  necessity  for  more 
careful  measurements.  He  might  build  a  new 
piece  of  apparatus,  one  that  would  be  more  suit- 
able for  measuring  volumes  and  pressures  over 
a  wide  range.  After  carrying  out  a  series  of 


measurements,  he  would  conclude  with  a  table 
of  data,  such  as  Table  1  -II.  A  table  of  data  is  a 
second  way  in  which  the  regularity  can  be  ex- 
pressed. In  a  third  mode  of  expression,  the  meas- 
urements can  be  presented  graphically  in  a  plot 
of  pressure  against  volume,  as  in  Figure  1-8. 

With  these  careful  measurements,  we  might 
also  seek  a  mathematical  statement  of  the  behav- 
ior. Sometimes  inspection  of  the  data  suggests  a 
relationship.  Sometimes  the  appearance  of  a 
plot,  such  as  Figure  1-8,  reveals  a  mathematical 
expression.  In  the  example  we're  studying,  the 
curve  resembles  a  hyperbola,  a  curve  described 
by  the  simple  equation,  xy  =  a  constant.  This 
similarity  leads  us  to  multiply  each  pressure  and 
volume  pair,  as  shown  in  the  third  column  of 
Table  l-II.  We  find  that  during  a  ten-fold  in- 
crease in  pressure,  the  product  P  X  V  is  fairly 
constant.  There  are  some  variations  in  the  prod- 
uct, as  seen  both  in  Table  l-II  and,  as  well,  in 
the  scatter  of  the  points  around  the  smooth  curve 
in  Figure  1-8.  The  random  nature  of  the  devia- 
tions from  constancy  suggests  that  they  measure 
the  uncertainty  due  to  experimental  technique. 
We  can  use  these  deviations  to  provide  an  esti- 
mate of  the  uncertainty  in  the  average,  ±0.6. 
(How  this  is  done  is  shown  in  Exercise  1-4.) 


14 


chemistry:   an  experimental  SCIENCE  I  CHAP.    1 


Hence,  with  reasonable  confidence  we  can  state 
the  regularity  mathematically: 

PXV=  22.4  ±  0.6 

Thus  we  have  found  four  ways  to  express  the 
regularity  between  the  pressure  and  volume  of 
oxygen  gas: 

(a)  Qualitatively:  As  the  pressure  rises,  the  vol- 
ume decreases. 

(b)  Quantitatively:  List  the  original  data  that 
show  how  pressure  and  volume  are  related, 
as  in  Table  1  -II. 

(c)  Graphically:  Plot  the  relationship  between 
pressure  and  volume  of  32.0  grams  of  oxygen 
gas  at  0°C,  as  in  Figure  1-8. 

(d)  Mathematically: 

PXV=  22.4  ±  0.6 

P  =  pressure  (in  atmospheres) 
V  =  volume  (in  liters)  of  32.0  grams 
of  oxygen  gas  at  0°C 

Obviously  the  regularity  expressed  in  the  qual- 
itative form  (a)  is  far  less  informative  than  any 
one  of  the  quantitative  presentations,  (b),  (c),  or 
(d).  The  relative  merits  of  the  expressions  (b), 
(c),  and  (d)  depend  upon  the  use.  Table  l-II  tells 
in  most  detail  exactly  how  much  is  known  about 
the  pressure-volume  behavior  of  oxygen  gas 
(from  this  experiment).  In  the  graphical  presen- 
tation of  Figure  1-8  the  trend  of  the  data  is 
shown  by  the  smooth  curve  drawn  to  pass  near 
as  many  points  as  possible.  Uncertainties  caused 


© 

zoo 

150 

- 

100 

- 

- 

50 

- 

- 

2  4  .6  .8 

Pressure,  atmospheres 


1.0 


Table  l-II. 

PRESSURE    AND    VOLUME    OF    32.0    GRAMS 
OF    OXYGEN    GAS    t  =  0  C 


PRESSURE 

(in  units  called 
atmospheres) 


VOLUME 

(in  units  called 
liters) 


PXV 


0.100 
0.200 
0.400 
0.600 
0.800 
1.00 


224 

109 
60.0 
35.7 
27.7 
22.4 


22.4 
21.8 
24.0 
21.4 
22.2 
22.4 


Average        22.4  ±  0.6 


Fig.  1-8.  A   plot  of  the  pressure  versus  volume  of 

32.0  grams  of  oxygen  gas;  t  =  0°C. 


by  experimental  errors  cause  the  data  points  to 
fall  above  and  below  this  smooth  curve.  Hence 
the  graphical  presentation  reveals  reliability  of 
the  measurements.  The  smooth  curve  "irons 
out"  these  uncertainties  and  provides  a  con- 
venient basis  for  predicting  volumes  at  inter- 
mediate pressures  (that  is,  for  interpolating). 
However,  from  the  standpoint  of  usefulness,  the 
mathematical  expression  (d)  is  often  the  best. 
It  is  the  most  compact  way  of  stating  the  regu- 
larity together  with  its  uncertainty.  Mathematics 
is  one  of  the  most  important  tools  of  chemistry. 
No  matter  how  expressed,  all  scientific  "rules," 
"laws,"  and  "theories"  are  statements  of  regu- 
larities of  nature.  Their  usefulness  depends  upon 
the  amount  of  experimental  evidence  that  shows 
that  the  "rule,"  "law,"  or  "theory"  corresponds 
to  experimental  reality.  Within  the  bounds  that 
it  is  known  to  correspond  to  experimental  reality, 
the  relation  can  be  used  for  prediction. 


EXERCISE  1-4 

(a)  Add  the  six  values  of  P  X  V  in  Table  l-II 
and  divide  by  6  to  obtain  the  average. 
(P  X  V)„  or,  {PV)„. 


SEC.     1-4    I    REVIEW 


15 


(b)  Now  add  a  fourth  column  to  Table  I  —I I 
showing  the  deviation  of  each  P  X  V  prod- 
uct from  (PV)av.  Head  this  column  with  the 
word  "Deviation,"  and  calculate  each  entry 
by  subtracting  (PV)av  from  the  measured 
value.  For  example,  the  second  entry  will  be 
-0.6  (since,  21.8  -  22.4  =  -0.6). 

(c)  When  you  have  completed  the  column  of 


deviations,  add  the  column  (disregarding  al- 
gebraic signs)  and  divide  by  6  to  obtain  an 
average  deviation, 
(d)  Compare  your  calculations  in  (a)  and  (c) 
with  the  result  given  in  Table  I -II, 

Average  =  22.4  db  0.6 


1-4    REVIEW 


This  chapter  began  with  the  statement  that  we 
would  find  through  experience  what  science  is 
all  about.  Already  you  have  had  opportunities 
to  do  so  in  the  laboratory.  We  see  that  science 
is  man's  systematic  investigation  of  his  en- 
vironment. Chapter  1  has  told  how  this  in- 
vestigation proceeds.  The  remainder  of  the  book 
is  concerned  with  those  parts  of  this  investigation 
that  are  carried  out  by  chemists.  Before  going 
on  to  see  what  chemistry  is,  let  us  review  your 
laboratory  accomplishments  so  far  with  empha- 
sis on  the  activities  of  science. 

1-4.1    Accumulating  Information  Through 
Observation 

Observation  of  a  burning  candle  reveals  an  as- 
tonishing complexity.  It  also  reveals  the  impor- 

Fig.  1-9.  A  good  experimentalist  is  a  good  observer. 

Record  in  your  notebook  at  the  moment  of 
observation.  Prepare  tables  in  advance. 


tance  and  value  of  careful  study  and  attention 
to  detail. 

In  your  experimentation,  be  alert  and  ready 
for  unexpected  developments.  Record  in  your 
notebook  at  the  moment  of  observation  a  de- 
scription of  everything  you  see.  The  time  of  the 
observation  frequently  has  importance.  Com- 
pleteness is,  by  far,  the  most  important  property 
of  a  good  notebook.  Next  in  importance,  legi- 
bility, neatness,  and  organization  make  your 
notebook  a  more  valuable  record.  Whenever 
possible,  prepare  tables  in  advance  for  the  results 
of  measurements  you  can  anticipate.  This  guar- 
antees that  you  won't  forget  to  note  important 
information,  and  it  frees  you  from  clerical  work 
during  an  experiment. 

Remember,  chemistry  is  built  upon  the  results 
of  experiments.  An  experiment  is  a  controlled 
sequence  of  observations.  A  good  experimentalist 
is  a  good  observer. 

1-4.2    Organizing  Information  and  Seeking 
Regularities  in  It 

The  mere  cataloging  of  observations  is  not  sci- 
ence. In  fact,  the  advance  of  our  knowledge  of 
nature  would  long  ago  have  ground  to  a  halt  if 
we  merely  made  observations.  The  multitude  of 
known  facts  can  be  dealt  with  only  if  it  is  pack- 
aged efficiently.  This  packaging  we  have  called 
"organizing  the  information"  and  "seeking  regu- 
larities." 

There  is  no  single  recipe  for  seeking  regulari- 
ties. That  is  probably  why  the  search  is  so 
interesting  and  why  the  scientist  receives  so  much 
personal  satisfaction  from  his  work.  Here  is  op- 


16 


chemistry:   an  experimental  SCIENCE  I  CHAP.    1 


portunity  for  originality,  opportunity  for  testing 
one's  wit  and  cleverness.  You  can  experience  the 
pleasure  a  scientist  derives  from  clarifying  a  pre- 
viously mystifying  behavior  by  careful  experi- 
ments of  your  own. 

In  our  study  of  the  candle  the  presence  of 
liquid  at  the  top  of  the  candle  caught  our  atten- 
tion. It  led  us  to  wonder  about  the  behavior  of 
other  familiar  solids  given  a  similar  treatment. 
In  this  case,  we  went  looking  for  regularity— we 
sought,  through  experiment,  to  discover  how 
other  solids  behave  on  heating.  Our  first  studies, 
when  organized,  led  us  to  the  generalization  that 
solids  melt  at  a  characteristic  temperature  when 
heated.  We  made  two  gains  thereby:  we  found 
an  efficient  expression  of  the  results  of  a  number 
of  experiments,  and  we  provided  a  basis  for  ex- 
pectation on  the  effect  of  heating  solids  which 
we  have  not  studied  before.  The  confidence  this 
expectation  deserves  is  fixed  by  the  amount  of 
evidence  supporting  the  generalization. 

1-4.3    Wondering  Why 

The  culmination  of  the  investigation  of  our  en- 


vironment we  have  dubbed  "wondering  why." 
We  seek  explanations.  Through  an  example,  we 
have  seen  that  an  explanation  is  the  discovery  of 
likenesses  connecting  a  process  we  do  not  under- 
stand with  processes  we  do  understand.  This  is 
the  most  rewarding  activity  of  science.  It  leads 
to  exploration.  Learn  to  ask  yourself  questions 
beginning  with  "Why"  when  you  observe — both 
in  and  out  of  the  chemistry  laboratory.  It  is  a 
good  habit  to  have,  and  it  frequently  makes  life 
more  interesting. 

You  have  had  opportunities  to  ask  many 
"Why"  questions  already  from  your  work  in  the 
laboratory.  In  fact,  there  are  enough  such  ques- 
tions to  provide  the  basis  for  the  rest  of  the 
course.  Some  of  the  questions  that  have  been 
raised  in  your  experiments  are  listed  at  the  end 
of  the  chapter.  Can  you  add  to  this  list?  How 
many  of  the  questions  can  you  answer  now?  We 
will  find  the  answers  to  many  of  them  in  our 
subsequent  study.  Some  may  not  yet  have  satis- 
factory answers.  These  are  the  most  interesting 
questions  because  they  point  into  the  future — 
your  future. 


SOME    QUESTIONS    RAISED    DURING    THE    STUDY    OF    A    BURNING    CANDLE 


Why  does  a  solid  absorb  heat  when  it  melts? 

Why  is  heat  liberated  in  the  burning  of  a  candle? 

Why  is  the  heat  effect  so  much  larger  in  the  chemical  reaction  than  in  the  phase  change  you  studied  ? 

Why  does  the  candle  react  with  air  to  give  carbon  dioxide  and  water  rather  than  the  reverse,  carbon  dioxide  and 

water  reacting  to  give  candle  and  air? 
Why  didn't  the  candle  react  with  air  (that  is,  burn)  while  the  candle  was  stored  in  your  desk  drawer  ?  Why  did  it 

wait  until  you  wanted  it  to  burn  ?  What  is  the  role  of  the  match  you  used  to  light  the  candle  ? 
Why  does  a  candle  burn  slowly  when  you  light  the  wick,  in  contrast  to  what  happens  when  you  light  the  "wick" 

of  a  firecracker  ? 
What  is  the  role  of  the  wick  of  the  candle? 

How  much  water  and  carbon  dioxide  are  produced  from  a  burning  candle? 
Why  does  carbon  dioxide  cause  limewater  to  become  cloudy  ? 

Why  does  the  burning  of  sulfur  produce  a  bad  smell  while  the  burning  of  steel  wool  produces  sparks? 
Why  does  a  flame  emit  colored  light? 
Why  is  the  base  of  the  flame  blue? 
What  is  the  dark  zone  in  the  candle  flame? 
Why  does  the  candle  flame  smoke  more  in  a  breeze  ? 
Why  don't  we  run  out  of  questions? 


CHAPTER 


2 


A  Scientific  Model: 
The  Atomic  Theory 


•  •  •  hypotheses  ought  to  be  fitted  merely  to  explain  the  properties  of 
things  and  not  attempt  to  predetermine  them  except  in  so  far  as  they 
can  be  an  aid  to  experiments. 

ISAAC    NEWTON,    1689 


One  of  the  activities  of  science  is  the  search  for 
regularity.  Regularities  that  directly  correlate  ex- 
perimental results  are  generally  called  rules  or 
laws.  A  more  abstract  regularity,  expressing  a 
hidden  likeness,  is  generally  called  a  model, 
theory,  or  principle.  Thus,  the  behavior  of  oxy- 
gen gas  summarized  in  the  equation  P  X  V  =  a 
constant  is  called  a  law.*  The  explanation  of  this 
same  regular  gas  behavior  in  terms  of  the  motion 
of  particles  is  called  a  theory.  It  is  a  greater 
abstraction  to  connect  the  PV  product  with  the 
mathematical  equations  that  describe  rebound- 
ing billiard  balls.  Nevertheless,  rules,  laws,  mod- 
els, theories,  and  principles  all  have  a  common 
aim — they  all  systematize  our  experimental 
knowledge.  They  all  state  regularities  among 
known  facts. 

The  more  abstract  regularities  come  from  the 
discovery  of  a  hidden  likeness.  When  the  likeness 


*  It  is  called  Boyle's  Law,  after  Robert  Boyle,  the 
scientist  who  first  discovered  this  particular  regularity. 


involves  a  real  physical  system  (such  as  rebound- 
ing billiard  balls),  the  explanation  is  usually 
called  a  model.  When  the  likeness  involves  an 
abstract  idea  (such  as  a  mathematical  equation), 
the  explanation  is  usually  called  a  theory.  There 
is  no  real  distinction  to  be  made,  however,  and 
we  shall  use  the  words  model  and  theory  inter- 
changeably. 

When  seeking  an  explanation,  we  sometimes 
find  more  than  one  explanation.  When  this  hap- 
pens the  model  (or  theory)  that  is  most  useful 
is  most  used.  A  model  that  proves  to  be  useful 
generally  points  to  new  directions  of  thought. 
The  new  directions  guide  us  to  new  experiments 
and,  thereby,  new  facts  come  to  light.  Often,  the 
new  facts  will  require  growth  of  the  model.  Oc- 
casionally the  new  facts  forcibly  contradict  the 
model  and  it  must  be  abandoned  in  favor  of 
another.  Both  the  growth  and  abandonment  of 
models  or  theories  reflect  increase  in  our  under- 
standing of  the  environment. 

Let  us  see  how  a  model  grows. 

17 


18 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    |    CHAP.    2 


2-1    IMPLICATIONS  AND  GROWTH  OF  A  SCIENTIFIC  MODEL 


As  an  example,  we  can  explore  the  implications 
of  our  explanation  of  the  behavior  of  gases. 

Question:  Why  does  a  balloon  expand  as  it  is 
inflated? 

Possible  Answer:  Perhaps  the  gas  put  in  the 
balloon  consists  of  a  collection  of  small  par- 
ticles that  rebound  from  the  wall  of  the  bal- 
loon just  as  billiard  balls  rebound  from  the 
cushions  of  a  billiard  table.  As  the  gas  particles 
rebound  from  the  balloon  wall,  they  push 
on  it. 

This  model  is  useful,  first,  because  we  can 
calculate  in  mathematical  detail  just  how  much 
push  a  billiard  ball  exerts  on  a  cushion  at  each 
rebound,  and,  second,  because  exactly  the  same 
mathematics  describes  the  pressure  behavior  of 
gas  in  a  balloon.  The  success  of  the  model  leads 
to  new  directions  of  thought.  For  example,  we 
might  now  wonder  whether  the  pressure-volume 
behavior  of  oxygen,  as  shown  in  Table  l-II 
(p.  14),  can  be  explained  in  terms  of  the  particle 
model  of  a  gas. 


2-1.1    The  Pressure-Volume  Behavior 
of  Oxygen  Gas 

The  experimental  data  in  Table  l-II  show  that 
decreasing  the  volume  by  one-half  doubles  the 
pressure  (within  the  uncertainty  of  the  measure- 
ments). How  does  the  particle  model  correlate 
with  this  observation?  We  picture  particles  of 
oxygen  bounding  back  and  forth  between  the 
walls  of  the  container.  The  pressure  is  deter- 
mined by  the  push  each  collision  gives  to  the 
wall  and  by  the  frequency  of  collisions.  If  the 
volume  is  halved  without  changing  the  number 
of  particles,  then  there  must  be  twice  as  many 
particles  per  liter.  With  twice  as  many  particles 
per  liter,  the  frequency  of  wall  collisions  will  be 
doubled.  Doubling  the  wall  collisions  will  double 
the  pressure.  Hence,  our  model  is  consistent  with 
observation:  Halving  the  volume  doubles  the 
pressure. 


Fig.  2-1.  In  the  particle  model,  wall  collisions  deter- 
mine pressure.  Halving  the  volume  doubles 
the  pressure. 


SEC.    2-1    I    IMPLICATIONS    AND    GROWTH    OF    A    SCIENTIFIC    MODEL 


19 


2-1.2    The  Pressure-Volume  Behavior 
of  Other  Gases 

Having  gained  this  understanding  of  the  pres- 
sure-volume behavior  of  oxygen,  it  is  natural  to 
wonder  whether  the  same  model  is  applicable 
to  other  gases.  Thus,  the  development  of  the 
theory  leads  us  to  perform  new  experiments. 
Such  experiments  provide  a  systematic  growth  of 
our  knowledge  of  our  environment.  They  are 
generally  much  more  effective  than  random, 
"shot-in-the-dark"  experiments. 

Two  other  gases  on  the  chemist's  shelf  are 
ammonia  and  hydrogen  chloride.  Is  the  particle 
model  applicable  to  them  as  well  as  to  oxygen? 
To  find  out,  we  must  perform  experiments  that 
duplicate  the  conditions  used  in  the  study  of 
oxygen.  Table  2-1  shows  pressure-volume  meas- 
urements for  32.0  grams  of  gaseous  ammonia  at 
0°C.  Table  2-II  shows  the  same  type  of  data  for 
32.0  grams  of  gaseous  hydrogen  chloride  at  this 
same  temperature. 

Table  2-1 

PRESSURE  AND  VOLUME  OF  32.0  GRAMS 
OF    AMMONIA    GAS    t  =  0  C 


PRESSURE 

(atmospheres) 


VOLUME 

(liters)         P  X  V 


0.100 

421 

42.1 

0.500 

84.2 

42.1 

1.00 

42.1 

42.1 

We  see  that  these  two  gases  also  show  the 
behavior  at  a  fixed  temperature,  PV  =  a  con- 
stant. The  particle  model  should  be  useful  for 
these  gases  as  well  as  for  oxygen.  On  the  other 
hand,  the  numerical  value  of  the  constant  varies 
from  one  gas  to  another,  if  the  same  weight  of 
gas  is  considered.  Thus  32.0  grams  of  oxygen  at 
0°C  and  one  atmosphere  occupy  22.4  liters.  The 
same  weight  of  ammonia  at  this  temperature  and 
pressure  occupies  42.1  liters.  The  same  weight  of 
hydrogen  chloride  occupies  only  19.6  liters.  The 
particle  model  of  gases  must  be  modified  to  ex- 
plain these  differences. 


Table  2-11 

PRESSURE    AND    VOLUME    OF   32.0    GRAMS 
OF    HYDROGEN    CHLORIDE    GAS    t  =  0°C 


PRESSUR£ 

(atmospheres) 


VOLUME 

(liters)         P  X  V 


0.100 

196 

19.6 

0.500 

39.2 

19.6 

1.00 

19.6 

19.6 

To  explain  this  behavior,  chemists  have  found 
it  convenient  to  consider  a  different  weight  of 
each  gas;  they  select  that  amount  that  gives  the 
same  PV  product  as  32.0  grams  of  oxygen  gas. 
Consider,  first,  ammonia  gas.  At  0°C  and  a  pres- 
sure of  one  atmosphere,  32.0  grams  of  ammonia 
occupies  42.1  liters.  We  have  taken  too  large  a 
weight  of  ammonia.  The  weight  of  ammonia 
needed  to  occupy  only  22.4  liters  at  this  pressure 
is  smaller  by  a  factor  22.4/42.1 : 

22.4 
wt.  ammonia  =  32.0  g  X  ttt  =  17.0  g 

Pressure-volume  data  for  this  weight  of  ammonia 
are  shown  in  Table  2-III. 


Table  2-III 

PRESSURE    AND    VOLUME    OF    17.0   GRAMS 
OF    AMMONIA    GAS 


PRESSURE 

(atmospheres) 


VOLUME 

Giters)         PX  V 


0.100 

224 

22.4 

0.500 

44.8 

22.4 

1.00 

22.4 

22.4 

EXERCISE  2-1 

If  32.00  grams  of  hydrogen  chloride  gas  (at  0°C 
and  one  atmosphere)  occupy  19.65  liters,  then  a 
larger  weight  of  hydrogen  chloride  is  needed  to 
occupy  the  larger  volume,  22.4  liters.  Show  that 
the  weight  needed  is  36.5  grams. 


20 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    I    CHAP.    2 


Now  the  regularity  between  pressure  and  vol- 
ume of  these  three  gases  can  be  expressed  as 
follows: 

For  32.0  grams  of  oxygen  at  0°C, 

P  X  V  =  22.4; 
for  17.0  grams  of  ammonia  at  0°C, 

PX  V  =  22.4; 
for  36.5  grams  of  hydrogen  chloride  at  0°C, 

P  X  V  =  22.4. 

Each  of  the  gases  has  a  behavior  consistent 
with  the  particle  model  of  a  gas  (PV  =  a  con- 
stant). However,  the  particles  of  the  gas  called 
oxygen  must  differ  from  the  particles  of  gas 
called  ammonia.  These,  in  turn,  must  differ  from 
the  particles  of  the  gas  called  hydrogen  chloride. 
How  do  the  particles  differ?  Why  is  it  that  32.0 
grams  of  oxygen  give  the  same  PV  product  as 
17.0  grams  of  ammonia  and  36.5  grams  of  hy- 
drogen chloride  (all  at  0°Q?  Do  the  particles 
have  different  weights?  Again,  we  are  led  to  new 
questions  and  new  questions  lead  to  new  experi- 
ments. 


2-1.3    Some  Properties  of  Gases 

What  gases  do  we  find  in  the  chemical  stock- 
room? How  do  they  compare  in  other  proper- 
ties? Looking  down  the  row  of  tanks  we  find  the 
names  ammonia,  chlorine,  hydrogen,  hydrogen 
chloride,  nitric  oxide,  nitrogen  dioxide,  and  oxy- 
gen, among  others.  Two  of  these  gases  are 
colored — chlorine  is  yellow-green  and  nitrogen 
dioxide  is  reddish-brown — and  the  other  five  are 
colorless.  The  colorless  gases  can  be  further 
sorted  according  to  their  solubilities  in  water. 
Figure  2-2  shows  what  happens  if  a  stoppered 
test  tube  full  of  each  gas  is  opened  with  the 
mouth  of  the  test  tube  under  water.  In  the  tubes 
containing  ammonia  and  hydrogen  chloride  the 
water  rises  rapidly,  filling  the  tubes.  These  two 
gases  dissolve  readily  in  water.  In  each  of  the 
other  three  test  tubes  the  liquid  level  rises  very 
little,  showing  that  little  gas  dissolves. 


Fig.  2-2.  Gases  have  different  solubilities  in  water. 


Immediately 
after  gases 
placed  in 
contact 
with  water 


Gases  with  high  solubility  in  water 
Ammonia  Hydrogen  chloride 


Gases  with  low  solubility  in  water 
Nitric  oxide  Oxygen  Hydrogen 


After  30 
minutes 
in  contact 
with  water 


SBC.    2-2   I   MOLECULES    AND    ATOMS 


21 


Though  ammonia  and  hydrogen  chloride  both 
dissolve  in  water,  these  two  gases  are  very-  dif- 
ferent in  other  properties.  For  example,  they 
behave  differently  when  placed  in  contact  with 
the  dye,  litmus.  This  dye,  when  moistened,  turns 
red  if  it  is  placed  in  hydrogen  chloride.  However, 
if  it  is  placed  in  ammonia,  it  turns  blue. 

We  have  not  yet  distinguished  the  gases  nitric 
oxide,  hydrogen,  and  oxygen.  Nitric  oxide  has 
its  own  personality.  Immediately  upon  exposure 
to  air,  the  colorless  nitric  oxide  becomes  reddish- 
brown — exactly  the  color  of  nitrogen  dioxide. 
Neither  oxygen  nor  hydrogen  behaves  this  way. 

The  gases  oxygen  and  hydrogen  are  readily 
distinguished  by  their  combustion  properties. 
When  a  glowing  splint  is  plunged  into  oxygen, 


the  splint  bursts  into  flame.  When  a  brightly 
glowing  splint  is  plunged  into  hydrogen,  the  glow 
is  either  extinguished  or,  if  air  has  mixed  with 
hydrogen,  it  produces  a  small  explosion. 

Thus  we  find  that  each  of  these  gases  has  dis- 
tinctive properties.  If  these  gases  are  made  up  of 
particles,  then  the  particles  must  be  distinctive. 
The  particles  that  are  present  in  ammonia  cannot 
be  like  the  particles  in  hydrogen  chloride,  or  like 
those  in  the  other  gases.  The  nature  of  the  am- 
monia particles,  then,  is  the  key  to  the  properties 
of  ammonia.  The  particles  that  make  up  a  gas 
determine  its  chemistry.  They  are  so  important 
to  the  chemist  that  they  are  given  a  special  name. 
A  gas  is  described  as  a  collection  of  particles 
called  molecules. 


2-2    MOLECULES  AND  ATOMS 


The  particles,  or  molecules,  of  the  gas  nitric 
oxide  cannot  be  exactly  like  those  of  nitrogen 
dioxide.  There  must  be  differences  that  account 
for  the  fact  that  one  gas  is  colorless  and  the  other 
reddish-brown.  Yet,  when  nitric  oxide  and  air 
are  mixed,  color  appears,  suggesting  that  nitro- 
gen dioxide  has  been  formed.  Apparently  mole- 
cules present  in  air  somehow  combine  with  the 
molecules  of  nitric  oxide  to  form  molecules  of 
nitrogen  dioxide.  We  would  like  to  develop  our 
picture  of  molecules  so  it  will  aid  us  in  discussing 
these  changes. 

To  explain  how  molecules  can  rearrange  and 
change,  we  assume  they  must  be  built  of  smaller 
fragments.  These  smaller  fragments,  or  building 


blocks,  are  called  atoms.  With  this  assumption, 
we  can  explain  differences  between  two  mole- 
cules in  terms  of  the  atoms  present  in  each 
molecule.  Nitric  oxide  is  different  from  hydrogen 
chloride  because  it  contains  different  atoms. 
Nitric  oxide  exposed  to  air  forms  nitrogen  di- 
oxide by  some  rearrangement  of  the  available 
atoms.  In  general,  the  properties  of  a  gas  are 
fixed  by  the  number  and  types  of  atoms  it  con- 
tains. 


Fig.  2-3.  Models  of  molecules.  The  properties  of  a 
molecule  are  fixed  by  the  number  and  types 
of  atoms  it  contains. 


A  model  of 
AMMONIA 


A  model  of 
HYDROGEN 
CHLORIDE 


22 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    I    CHAP.    2 


SEARCH  FOR  AN  EXPLANATION : 

Why  do  equal  volumes  of  ammonia  and  hydrogen  chloride  have  different  weights? 


At  room  temperature 

and  one 
atmosphere  pressure. 


and 


1     2.45  liters  of 
hydrogen  chloride 
weigh  3.  65 grams 


Chemists  construct  models  of  molecules  to 
show  how  many  atoms  they  contain.  Figure  2-3 
shows  some  examples.  The  model  of  ammonia 
represents  three  atoms  of  one  kind  attached  to 
one  atom  of  another  kind.  The  model  of  hydro- 
gen chloride  contains  only  two  atoms,  and  of 
different  kind.  We  shall  spend  the  entire  year 
discussing  why  these  and  other  models  are  con- 
structed as  they  are.  We  shall  see  that  a  molecule 
of  ammonia  is  pictured  as  shown  in  Figure  2-3 
because  this  model  helps  us  explain  the  proper- 
ties of  ammonia.  Throughout  the  course,  we 
shall  investigate  properties  of  substances  found 
in  nature  or  prepared  in  the  laboratory  and  then 
we  will  seek  explanations  in  terms  of  the  num- 
bers, types,  and  arrangements  of  the  atoms  pres- 
ent. These  explanations  are  called  the  atomic 
theory.  The  atomic  theory  is  regarded  as  the 
cornerstone  of  chemistry. 

2-2.1    The  Weights  of  Molecules 

We  have  discovered  that  32.0  grams  of  oxygen, 
17.0  grams  of  ammonia,  and  36.5  grams  of  hy- 
drogen chloride  each  exhibits  the  regular  be- 
havior, 

PV  =  22.4        at  0°C  (/) 

For  any  of  the  three  gases,  if  the  pressure  is 
given,  the  volume  occupied  can  be  calculated. 
At  one  atmosphere,  each  of  the  specified  weights 
of  gas  occupies  22.4  liters.  At  two  atmospheres, 
each  gas  is  compressed  into  a  smaller  volume, 
11.2  liters.  We  wonder,  Why  do  22.4  liters  of 


WHY? 

Fig.  2-4.  Why  do  equal  volume  of  ammonia  and  hy- 
drogen chloride  have  different  weights? 


ammonia  weigh  17.0  grams  when  the  same  vol- 
ume of  hydrogen  chloride  weighs  36.5  grams? 
There  are  two  factors  to  consider.  These  are 
the  same  two  factors  we  would  be  concerned  with 
if  we  were  to  ask  why  a  bag  full  of  beans  weighs 
17.0  grams  whereas  the  same  bag  full  of  marbles 
weighs  36.5  grams.  The  explanation  would  be 
found  by  comparing  the  number  of  beans  in  the 
bag  and  the  weight  per  bean  to  the  number  and 
individual  weights  of  marbles  in  the  same  bag. 
In  our  gas  problem  we  make  the  same  kind  of 
comparison;  the  weight  per  molecule  and  num- 
ber of  ammonia  molecules  in  22.4  liters  must  be 
compared  to  the  weight  per  molecule  and  num- 
ber of  hydrogen  chloride  molecules  in  22.4  liters. 
There  are  two  particularly  simple  possibilities: 

(A)  Perhaps: 

(1)  Equal  volumes  of  these  two  gases  con- 
tain the  same  number  of  molecules,  ai.d 

(2)  ammonia  molecules  weigh  less,  per  mol- 
ecule, than  hydrogen  chloride  molecules 
by  the  factor  17.0/36.5. 

(B)  Perhaps: 

(1)  Equal  volumes  of  these  two  gases  con- 
tain different  numbers  of  molecules,  am- 
monia containing  fewer  by  the  factor 
17.0/36.5,  and 

(2)  ammonia  molecules  weigh  exactly  the 
same  as  hydrogen  chloride  molecules. 


SEC.    2-2    I    MOLECULES    AND    ATOMS 


23 


SEARCH  FOR  AN EXPLANATION 7 

Why  do  equal  volumes  of  ammonia  and  hydrogen  chloride  have  different  weights? 


At  room  temperature 

and  one 
atmosphere  pressure. 


'  ^2.45  titers  of 
<l  III 

ammonia,  weigh 
ii 
/.  70  grams 


and 


v.. 

n  in 
hydrogen  chloride 


WHY? 


MODEL  A 

Perhaps: 

1.  2.  45  liters  of  ammonia 
contain  the  same 
number  of  molecules 
as  2.45  titers  of 
hydrogen,  chloride. 

and 

2.  Afnmonia  molecules 
weigh  less  than, 
hydrogen  chloride 
molecules,  (less  by  a 
factor,  J.  70/3. 65). 


MODEL  B 

Perhaps: 

1.  2.45  titers  of  ammonia 
contain  fewer  molecules 
than  do  2.45  liters  of 
hydrogen,  chloride, 
(lest  by  a  factor, 
1.70/3.65). 

and 

2.  Ammonia  molecules 
■weigh  the  same  as 
hydrogen  chloride 
molecules 


Fig.  2-5.  Two  simple  models  to  explain  the  weights  of  equal  volumes  of  gases. 


24 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    I    CHAP.    2 


30  ml 
ammonia 


Fig.  2-6.  An  apparatus  suited  to  the  measurement  of 

volumes  of  gases. 


These  two  possibilities  are  attractive  because 
they  are  simple — one  factor  alone  is  held  respon- 
sible for  the  weight  difference.  We  must  be 
prepared,  however,  for  disappointment.  There  is 
the  third  possibility  that  neither  of  these  pro- 
posals, A  or  B,  accounts  for  the  properties  of 
gases.  After  all,  neither  A  nor  B  applies  to  the 
beans  and  marbles  example.  The  bag  probably 
wouldn't  contain  the  same  number  of  beans  as 
marbles  (as  in  B)  but,  in  addition,  beans  and 
marbles  don't  weigh  the  same  (as  in  A).  We  need 
more  information  to  decide  if  either  proposal  A 
or  B  applies  to  gases.  More  information  is  ob- 
tained by  observing  how  some  gases  behave 
when  mixed. 

2-2.2     Mixtures  off  Ammonia  and 
Hydrogen  Chloride 

When  hydrogen  chloride  and  ammonia  gases  are 
mixed,  a  white  powder  is  formed.  When  hydro- 
gen chloride  molecules  and  ammonia  molecules 
are  mixed,  the  atoms  are  rearranged  and  an  en- 


tirely different  substance,  a  solid,  results.  A  quan- 
titative study  of  this  process  is  informative. 

Figure  2-6  shows  an  apparatus  suited  to  the 
measurement  of  volumes  of  gases.  Thirty  milli- 
liters of  ammonia  have  been  admitted  to  the  left 
tube  from  the  ammonia  storage  tank.  Next,  50 
ml  of  hydrogen  chloride  were  admitted  to  the 
right  tube.  The  leveling  bulbs  were  used  to  adjust 
the  pressure  of  each  gas  to  one  atmosphere.  The 
apparatus  is  ready.  The  hydrogen  chloride  sam- 
ple can  be  transferred  slowly  into  the  tube  con- 
taining the  ammonia.  Figure  2-7  shows  the  prog- 
ress of  the  experiment. 

In  Figure  2-7A  we  see  the  situation  after  20  ml 
of  hydrogen  chloride  gas  have  been  transferred. 
A  cloud  of  the  white  solid  fills  the  left  tube  where 
the  gases  mix  and,  after  the  leveling  bulbs  are 
adjusted,  we  find  that  just  10  ml  of  ammonia 
remain.  Twenty  milliliters  of  hydrogen  chloride 
combined  with  just  20  ml  of  ammonia,  forming 
the  white  solid.  Ten  milliliters  more  hydrogen 
chloride  are  just  enough  to  consume  the  last  of 
the  ammonia,  forming  the  solid  of  insignificant 
volume,  as  shown  in  Figure  2-7B.  Continued 
addition  of  hydrogen  chloride  causes  no  further 
solid  formation  but  merely  leaves  an  excess  of 
hydrogen  chloride  in  the  left  tube  (Figure  2-7C). 

This  is  a  significant  and  simple  result.  Thirty 
milliliters  of  hydrogen  chloride  combine  with 
just  30  ml  of  ammonia,  measured  at  the  same 
temperature  and  pressure.  Therefore,  one  liter  of 
hydrogen  chloride  would  combine  with  just  one 
liter  of  ammonia.  Though  a  given  volume  of 
ammonia  weighs  less  than  the  same  volume  of 

Fig.  2-7.  Mixing  measured  volumes  of  hydrogen  chlo- 
ride gas  and  ammonia  gas. 


SEC.    2-2    I    MOLECULES    AND    ATOMS 


25 


hydrogen  chloride  (less  by  the  factor  17.0/36.5), 
these  equal  volumes  combine.  This  simple  situa- 
tion suggests  that  we  should  seek  a  simple  ex- 
planation. Our  proposal  A  in  Section  2-2.1  fits 
nicely.  If  we  propose  that  equal  volumes  contain 
equal  numbers  of  molecules,  then  30  ml  of  am- 
monia contain  the  same  number  of  molecules  as 
do  30  ml  of  hydrogen  chloride.  Proposal  A  leads 
us  to  conclude  that  one  molecule  of  ammonia 
combines  with  one  molecule  of  hydrogen  chlo- 
ride to  form  the  white  solid.  Through  proposal 
A,  the  combining  volumes  tell  us  the  numbers 
of  molecules  that  combine.  In  contrast,  there  is 
no  correspondingly  simple  way  to  explain  the 
new  data  with  proposal  B. 

Of  course,  a  single  example  hardly  furnishes 
compelling  evidence  that  equal  volumes  of  any 
pair  of  gases  (at  the  same  temperature  and  pres- 
sure) contain  equal  numbers  of  molecules.  Nei- 
ther can  we  be  convinced  on  the  basis  of 
simplicity  when  we  have  but  one  example.  How- 
ever, many  gases  behave  as  simply  as  a  mixture 
of  hydrogen  chloride  and  ammonia.  For  ex- 
ample, 


2-2.3    The  Relative  Weights  of  Molecules 

The  importance  of  Avogadro's  Hypothesis  is 
that  it  furnishes  a  basis  for  weighing  molecules. 
Two  equal  volumes  of  gas  (at  the  same  tempera- 
ture and  pressure)  are  weighed.  If  we  assume 
these  two  volumes  contain  identical  numbers  of 
molecules,  then  we  must  also  conclude  that  the 
gas  that  weighs  more  must  have  heavier  mole- 
cules. Furthermore,  the  ratio  of  the  weights  of 
the  molecules  must  be  exactly  the  ratio  of  the 
weights  of  the  two  gas  samples. 

For  example,  in  Table  1  -I I  (p.  14)  data  were 
given  that  show  that  32.0  grams  of  oxygen  at  0°C 
and  one  atmosphere  pressure  occupy  22.4  liters. 
This  same  volume  of  ammonia  (also  at  0°C  and 
one  atmosphere  pressure)  weighs  17.0  grams.  By 
Avogadro's  Hypothesis,  these  two  volumes  con- 
tain equal  numbers  of  molecules.  Hence  each 
ammonia  molecule  must  weigh  less  than  an  oxy- 
gen molecule  by  the  factor  17.0/32.0.  By  the 
same  argument,  each  hydrogen  chloride  mole- 
cule must  weigh  more  than  an  oxygen  molecule 
by  the  factor  36.5/32.0.  We  say  "must  weigh" 
but,  of  course,  this  is  a  valid  statement  only  if 


two  liters  of 
the  gas 
nitric  oxide 


combine 
with 


one  liter  of 
the  gas 
oxygen 


to  form 


nitrogen 
dioxide 


one  liter  of 
the  gas 
oxygen 


will  burn 
coal  to 
form 


one  liter  of  the 
gas  carbon 
dioxide 


two  liters  of 
the  gas 
hydrogen 


combine 
with 


one  liter  of 
the  gas 
oxygen 


These  simple,  integer  volume  ratios  confirm 
the  usefulness  of  the  interpretation  that  equal 
volumes  contain  equal  numbers  of  molecules. 
This  proposal  was  first  made  in  1811  by  an 
Italian  scientist,  Amadeo  Avogadro;  hence  it  is 
called  Avogadro's  Hypothesis.  It  has  been  used 
successfully  in  explaining  the  properties  of  gases 
for  a  century  and  a  half. 

Avogadro's  Hypothesis:  Equal  volumes  of 
gases,  measured  at  the  same  temperature  and 
pressure,  contain  equal  numbers  of  molecules. 


to  form 


water 


Avogadro's  Hypothesis  is  applicable. 

By  many  such  weighings,  scientists  have 
learned  the  relative  weights  of  many  gases.  The 
experiment  is  fairly  simple.  A  carefully  meas- 
ured volume  of  oxygen  is  weighed  at  a  fixed 
pressure  and  temperature.  Then  the  same  volume 
of  another  gas  is  weighed  at  this  same  pressure 
and  temperature.  The  relative  weights  of  the 
gases  indicate  the  relative  weights  of  the  mole- 
cules, provided  Avogadro's  Hypothesis  is  ap- 
plicable. Neither  the  pressure  nor  the  tempera- 


26 


A    SCIENTIFIC    MODEL!    THE    ATOMIC    THEORY    I    CHAP.    2 


ture  need  be  measured,  provided  they  are  held 
constant. 

Table  2-IV 

WEIGHTS  OF  EQUAL  VOLUMES  OF 
GASES  UNDER  FIXED  TEMPERATURE 
AND  PRESSURE  (BASED  ON  THE  VOLUME 
OCCUPIED    BY    32.0    GRAMS    OF    OXYGEN) 

WEIGHT 

gas  (grams) 


oxygen 
ammonia 
chlorine 
hydrogen 
hydrogen  chloride 
nitric  oxide 
nitrogen  dioxide 


(32.0) 
17.0 
71.0 
2.02 
36.5 
30.0 
46.0 


Table  2-IV  shows  for  some  other  gases  the 
weights  that  have  the  same  volume  as  32.0  grams 


been  formed.  How  can  this  change  be  discussed 
in  terms  of  our  molecular  model  of  a  gas? 

First,  we  explain  the  differences  between  nitric 
oxide,  oxygen,  and  nitrogen  dioxide  by  asserting 
that  the  molecules  of  nitric  oxide,  oxygen,  and 
nitrogen  dioxide  are  somehow  different.  They 
must  be  composed  of  smaller  components  that 
we  call  atoms.  The  numbers  and  kinds  of  atoms 
in  a  molecule  of  nitric  oxide  must  be  different 
from  the  numbers  and  kinds  of  atoms  in  a  mole- 
cule of  oxygen. 

Now  we  find  that  nitrogen  dioxide  can  be 
formed  from  a  mixture  of  nitric  oxide  and  oxy- 
gen. This  means  that  the  atoms  in  nitrogen  di- 
oxide must  have  come  from  those  in  nitric  oxide 
together  with  those  in  oxygen. 

Finally,  we  discover  that  exactly  two  volumes 
of  nitric  oxide  combine  with  one  volume  of  oxy- 
gen and  that  exactly  two  volumes  of  nitrogen 
dioxide  are  formed.  According  to  Avogadro's 
Hypothesis,  this  indicates  that 


two  molecules 
of  nitric 
oxide 


combine 
with 


one  molecule 
of  oxygen 


to  form 


two  molecules 
of  nitrogen 
dioxide 


of  oxygen  (at  the  same  pressure  and  tempera- 
ture). 

2-2.4    The  Number  of  Atoms  in  a  Molecule 

Figure  2-3  shows  a  model  of  an  ammonia  mole- 
cule and  a  model  of  a  hydrogen  chloride  mole- 
cule. These  models  show  how  chemists  picture 
the  molecule  of  ammonia:  it  contains  four 
atoms.  A  hydrogen  chloride  molecule  contains 
only  two  atoms.  Chemists  decide  how  to  con- 
struct these  molecules  from  the  same  type  of 
information  described  in  Section  2-2.2,  by  the 
volumes  of  gases  that  combine. 

Consider  the  combination  of  nitric  oxide  and 
oxygen.  Nitric  oxide  (a  colorless  gas)  when  mixed 
with  oxygen  gas  (also  colorless)  becomes  reddish- 
brown.  The  color  is  identical  to  that  of  another 
gas,  nitrogen  dioxide.  All  the  properties  of  the 
nitric  oxide-oxygen  mixture  are  consistent  with 
the  conclusion  that  the  gas  nitrogen  dioxide  has 


All  of  the  atoms  in  the  two  molecules  of  nitrogen 
dioxide  came  from  two  molecules  of  nitric  oxide 
and  one  molecule  of  oxygen.  Of  course,  the  two 
molecules  of  nitrogen  dioxide  have  twice  as  many 
atoms  as  does  a  single  molecule  of  nitrogen  di- 
oxide. Hence,  no  matter  how  many  atoms  one 
molecule  of  nitrogen  dioxide  might  contain  (for 
example,  one,  two,  three,  four,  .  .  .),  two  mole- 
cules of  nitrogen  dioxide  must  contain  an  even 
number  of  atoms  (for  example,  two,  four,  six, 
eight,  .  .  .).  The  same  statement  is  applicable  to 
the  two  molecules  of  nitric  oxide  that  were  com- 
bined. No  matter  how  many  atoms  one  molecule 
of  nitric  oxide  contains,  two  molecules  must  con- 
tain an  even  number  of  atoms. 

Thus  we  see  that  after  the  even  number  of 
atoms  in  two  molecules  of  nitric  oxide  have 
combined  with  the  atoms  in  one  molecule  of 
oxygen,  there  is  still  an  even  number  of  atoms. 
This  can  be  so  only  if  a  molecule  of  oxygen  also 
contains  an  even  number  of  atoms.  We  are  led 


SEC.    2-2    I    MOLECULES    AND    ATOMS 


27 


to  the  conclusion  that  a  molecule  of  oxygen  con- 
tains an  even  number  of  atoms. 

This  can  be  demonstrated  clearly  in  algebraic 
language. 

Suppose: 
one  molecule  of  nitric  oxide  contains 

X  atoms, 
one  molecule  of  oxygen  contains 

Y  atoms,  and 
one  molecule  of  nitrogen  dioxide  contains 

Z  atoms, 
where  X,  Y,  and  Z  are  integers. 

Then: 
two  molecules  ot  nitric  oxide  contain 

2X  atoms,  and 
two  molecules  of  nitrogen  dioxide  contain 

2Z  atoms. 

Also: 
all  of  the  atoms  in  two  molecules  of  nitrogen 
dioxide  (2Z  atoms)  came  from  two  mole- 
cules of  nitric  oxide  (2X  atoms)  plus  one 
molecule  of  oxygen  ( Y  atoms),  or 


IX  +  Y  =  2Z 


So: 


we  can  solve  for  Y  by  subtracting  2X  from 
each  side  of  this  equation: 

Y  =2Z-2X  (3) 

Y=2(Z-  X)  (4) 

Thus  no  matter  what  the  integer  values  of  Z 
and  X  are,  their  difference  (Z  —  X)\s  an  integer. 
Since  doubling  any  integer  produces  an  even 
number,  Y  =  2(Z  —  X)  must  be  an  even  num- 
ber. We  have  proved  that  a  molecule  of  oxygen 
must  contain  an  even  number  of  atoms. 

The  simplest  acceptable  structure  we  can  pic- 
ture for  oxygen  is  that  it  contains  two  atoms. 
More  experiments  are  needed  before  we  can 
eliminate  the  possibility  that  oxygen  contains 
four,  six,  or  a  higher  (but  even)  number  of  atoms. 


EXERCISE  2-2 

In  Section  2-2.2  (p.  24)  it  was  noted  that  two 
volumes  of  hydrogen  gas  combine  with  one  vol- 
ume of  oxygen  gas  and  two  volumes  of  gaseous 


water  are  produced.  According  to  Avogadro's 
Hypothesis,  this  means  that  two  molecules  of 
hydrogen  combine  with  one  molecule  of  oxygen 
to  form  two  molecules  of  water.  If  we  define 

X  =  the  number  of  atoms  in  a  molecule  of 

hydrogen, 
Y  =  the  number  of  atoms  in  a  molecule  of 

oxygen, 
Z  =  the  number  of  atoms  in  a  molecule  of 

water, 

then  we  have  the  algebraic  relationship 
2X  +  Y  =  2Z 

(a)  Convince  yourself  that  from  these  data  alone 
we  can  conclude  that  Y  must  be  even  (that  is, 
oxygen  molecules  must  contain  an  even  num- 
ber of  atoms). 

(b)  By  solving  for  X  in  terms  of  Y  and  Z,  con- 
vince yourself  that  from  the  above  data 
alone,  X  could  be  odd  or  even. 


C?)      2-2.5    Atoms  in  Liquids  and  Solids 


When  a  candle  is  burned,  a  gas  is  produced — a 
gas  containing  carbon  dioxide  and  water  vapor. 
It  is  useful  to  describe  such  a  gas  as  a  collection 
of  molecules,  each  molecule  containing  smaller 
units  called  atoms.  Each  carbon  dioxide  mole- 
cule contains  one  carbon  atom  and  two  oxygen 
atoms.  Each  water  molecule  contains  one  oxygen 
atom  and  two  hydrogen  atoms.  Where  did  these 
atoms  come  from?  Were  they  present  in  the 
candle  before  it  burned? 

Similar  questions  are  raised  if  we  consider  the 
effect  of  cooling  the  gases  produced  from  the 
candle.  Cooling  these  gases  results  in  condensa- 
tion— drops  of  liquid  water  appear.  If  the  water 
vapor  contains  molecules,  made  up  of  atoms, 
what  happens  to  these  molecules  (and  atoms) 
when  the  gas  condenses?  Are  they  still  present 
in  the  liquid? 

Scientists  always  seek  the  simplest  explanation 
that  fits  the  known  facts.  Since  we  find  it  con- 
venient to  describe  water  vapor  as  a  collection 
of  groups  of  atoms  (called  molecules),  the  sim- 
plest assumption  we  can  make  about  the  con- 


28 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    |    CHAP.    2 


5-»    '      .<—,-- 


Solid  carbon    dioxide 


Liquid    carbon    dioxi 


G-aseous   carbon  dioxide 


Fig.  2-8.  All  matter  consists  of  particles.  In  a  gas,  the 
particles  are  far  apart;  in  a  liquid  or  solid, 
they  are  close  together. 


densation  of  this  vapor  is  that  the  liquid  still 
contains  the  atoms.  Since  a  candle  burns  to 
produce  gases — collections  of  molecules — the 
simplest  assumption  we  can  make  is  that  the 
candle  already  contained  the  atoms  that  formed 
the  gaseous  molecules  during  combustion.  These 
simplifying  assumptions — that  liquids  and  solids 
are  made  up  of  atoms — are  acceptable  and  con- 
venient as  long  as  they  prove  to  be  consistent 
with  all  that  is  known  about  liquids  and  solids. 
Thus,  we  are  led  to  the  view  that  all  matter 


consists  of  particles.  We  can  state  this  as  a  pro- 
posal. 

Proposal:  All  matter,  whether  solid,  liquid,  or  gas, 
consists  of  particles.  In  a  gas  these  particles  are 
far  apart;  in  a  liquid  or  a  solid,  they  are  packed 
close  together. 

This  proposal  is  called  the  atomic  theory. 
As  with  any  theory,  its  value  depends  upon  its 
ability  to  aid  us  in  explaining  facts  of  nature. 
There  is  no  more  valuable  theory  in  science  than 
the  atomic  theory.  We  shall  use  it  throughout 
this  course.  Later,  in  Chapter  14,  we  shall  review 
many  of  the  types  of  experiments  which  cause 
chemists  to  regard  the  atomic  theory  as  the 
cornerstone  of  their  science. 


2-3     SUBSTANCES:  ELEMENTS  AND  COMPOUNDS 


Molecules  are  clusters  of  atoms.  Two  types  of 
molecules  are  possible.  Some  molecules  are  clus- 
ters of  atoms  in  which  all  the  atoms  in  a  cluster 
are  identical;  some  molecules  contain  two  or 
more  different  kinds  of  atoms.  These  two  kinds 
of  molecules  are  given  different  names. 

An  element  or  elementary  substance  contains 
only  one  kind  of  atom. 

A  compound  or  compound  substance  contains 
two  or  more  kinds  of  atoms. 


Usually  a  good  deal  of  experimentation  is 
needed  before  a  substance  can  be  considered  to 
be  pure.  Even  then,  much  more  work  and  study 
are  needed  before  one  can  decide  with  confidence 
that  a  given  pure  substance  is  an  element  or  a 
compound.  Consider  the  substance  water.  Water 
is  probably  the  most  familiar  substance  in  our 
environment  and  all  of  us  recognize  it  easily.  We 
are  familiar  with  its  appearance  and  feel,  its 
density  (weight  per  unit  volume),  the  way  in 


sec.  2-3  I  substances:  elements  and  compounds 


29 


which  it  flows,  the  temperature  at  which  it 
freezes  and  boils,  and  the  way  in  which  it  dis- 
solves sugar  and  salt.  Because  water  is  identified 
by  constant  and  characteristic  properties,  it  is  a 
pure  substance.  In  a  later  experiment  you  will 
see  how  we  can  change  the  pure  substance  water 
into  two  other  substances,  hydrogen  gas  and 
oxygen  gas.  The  hydrogen  and  oxygen  are  pro- 
duced in  definite  amounts.  Since  water  can  be 
decomposed  into  two  other  substances,  it  must 
contain  at  least  two  kinds  of  atoms.  Hence  water 
is  a  compound. 

Notice  the  pattern  here.  First,  we  established 
the  characteristic  properties  of  water  that  cause 
us  to  identify  it  as  a  pure  substance.  Second,  we 
found  a  change  in  which  two  other  substances 
were  formed  in  definite  amounts  from  water 
alone.  This  second  piece  of  information  shows 
that  water  contains  more  than  one  kind  of  atom 
and  that,  hence,  water  is  a  compound. 

Common  sugar  is  another  example  of  a  sub- 
stance. Most  commercial  samples  of  white  sugar 
are  rather  pure;  that  is,  they  contain  only  very 
small  amounts  of  substances  other  than  sugar. 
One  characteristic  property  of  sugar  is  its  sweet- 
ness. Another  characteristic  property  is  the  way 
it  dissolves  in  water.  Still  another  is  the  way  it 
behaves  when  heated.  At  a  definite  temperature 
sugar  not  only  begins  to  melt  to  a  liquid  but  it 
also  begins  to  decompose.  The  liquid  darkens 
and  gaseous  water  bubbles  off.  Finally,  a  black 
solid  (charcoal)  remains  in  the  container.  We 
recognize  the  black  solid  as  a  form  of  carbon. 
Thus,  pure  sugar,  identified  by  its  characteristic 
properties,  can  be  decomposed  to  form  water 
and  charcoal  in  definite  amounts.  Sugar  is  a 
compound. 

Water  and  sugar  are  compounds.  What  about 
hydrogen  and  oxygen?  Hydrogen,  for  example, 
is  a  gas  at  normal  conditions.  It  can  be  liquefied 
at  a  characteristic  temperature  by  cooling.  By 
further  cooling  it  can  be  solidified  at  a  second 
characteristic  temperature.  It  is  a  pure  substance. 
No  treatment,  however,  causes  it  to  form  two 
other  substances.  Hydrogen  must  contain  only 
one  kind  of  atom,  hence  hydrogen  is  an  element. 
We  call  this  kind  of  atom,  the  hydrogen  atom. 
Oxygen,  too,  has  characteristic  properties  but 


cannot  be  caused  to  form  two  other  substances. 
Oxygen,  then,  is  an  element — it  contains  only  one 
kind  of  atom,  called  the  oxygen  atom. 

Now  we  can  return  to  the  decomposition  of 
water.  Water  can  be  decomposed  to  give  hydro- 
gen and  oxygen.  Since  hydrogen  contains  only 
hydrogen  atoms  and  oxygen  contains  only  oxy- 
gen atoms,  water  molecules  must  contain  some 
hydrogen  atoms  and  some  oxygen  atoms,  but  no 
other  kind  of  atom. 

This  type  of  problem  is  one  of  the  most  im- 
portant in  chemistry — deciding  what  atoms  are 
present  in  a  given  substance.  How  important  this 
is  can  be  seen  by  comparing  the  three  substances 
water,  oxygen,  and  hydrogen.  Both  water  and 
oxygen  contain  oxygen  atoms  but  these  sub- 
stances are  very  different  in  their  properties.  Both 
water  and  hydrogen  contain  hydrogen  atoms  but 
these  substances  are  no  more  alike  than  are 
water  and  oxygen.  The  properties  of  water  are 
fixed  by  the  combination  of  the  two  kinds  of 
atoms  and  these  properties  are  distinctive. 


EXERCISE  2-3 

What  differences  between  water,  oxygen,  and 
hydrogen  can  you  point  out  from  your  own  ex- 
perience? For  example,  you  might  consider 

(a)  boiling  and  melting  points, 

(b)  role  in  combustion, 

(c)  role  in  supporting  life. 


Sugar  is  another  substance  that  contains  both 
oxygen  and  hydrogen  atoms  but  it  contains  car- 
bon atoms  as  well.  Sugar  does  not  resemble 
water,  oxygen,  or  hydrogen.  The  presence  of  the 
various  types  of  atoms  and  their  arrangement 
accounts  for  the  distinctive  properties  which 
identify  sugar.  In  any  substance,  the  atoms  pres- 
ent, their  numbers,  and  their  arrangement  fix  the 
properties  of  that  substance. 

2-3.1    The  Elements 

An  element  is  a  pure  substance  that  contains 
only  one  kind  of  atom.  There  are  about  one  hun- 


30 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY   I    CHAP.    2 


dred  different  elements  known  today — hence 
there  are  about  one  hundred  kinds  of  atoms  that 
are  chemically  different.  Some  of  these  elements 
occur  pure  in  nature  and  hence  have  been  known 
for  thousands  of  years.  Such  elements  as  iron, 
silver,  gold,  mercury,  and  sulfur  were  known  to 
the  ancients  and  were  given  Latin  names  by  the 
alchemists.  For  example,  iron  was  called  ferrum, 
silver  was  called  argentum,  and  gold  was  called 
aurum. 

During  the  nineteenth  century  the  discovery  of 
elements  increased  as  chemists  began  to  adopt 
quantitative  methods.  At  the  beginning  of  the 
nineteenth  century,  perhaps  26  elements  were 
known.  One  hundred  years  later,  at  the  begin- 
ning of  the  present  century,  over  81  elements 
were  known.  Over  twice  as  many  elements  were 
discovered  in  that  one  century  as  were  discovered 
in  all  of  time  before. 

Figure  2-9  shows,  as  a  function  of  time,  our 


Fig.  2-9.  The  discovery  of  the  elements.  A.  The  total 
number  of  elements  known  as  a  function 
of  time.  B.  The  number  of  elements  discov- 
ered in  each  half-century  since  1700. 


t 

a  * 

~l 


100 


80- 


60 


40 


V 


20- 


1700 


A 


H3I 


1800 
Year 


1900 


30 


K 

Hi 

5, 


20 


10 


m 
% 


m 


13  before 
1700 


# 


■ 


■,^wii 


1700 


1800 
Year 


1900 


knowledge  of  the  elements.  In  Figure  2-9A  we 
see  how  the  total  number  of  elements  known  has 
increased  since  1700.  Figure  2-9B  re-expresses 
this  same  information  in  terms  of  the  number  of 
elements  discovered  in  each  half-century  since 
1700.  Both  graphs  show  that  the  rate  of  discovery 
of  new  elements  is  declining.  The  plots  suggest 
that  there  is  a  limited  number  of  elements  to  be 
found  in  nature. 

Each  element  has  been  named  and,  for  con- 
venience, has  been  given  a  nickname — a  short- 
hand symbol  of  one  or  two  letters.  Thus  the 
element  carbon  is  symbolized  by  the  letter  C,  the 
element  neon  by  the  letters  Ne.  The  symbols  are 
adopted  by  international  agreement  among 
chemists.  Eleven  of  the  elements  have  names 
derived  from  the  capitalized  first  letter  of  the 
Latin  name  of  the  element  and,  if  necessary,  by 
a  second  letter  (uncapitalized).*  These  eleven  in- 
clude seven  common  metals  known  to  the  an- 
cients. (See  Table  2-V.) 

The  elements  discovered  more  recently  have 
the  same  names  in  all  languages,  again  by  inter- 
national agreement.  Except  for  the  eleven  ele- 
ments listed  in  Table  2-V,  all  of  the  elements 
have  symbols  that  can  be  derived  from  their 
English  names.  For  example,  the  symbols  for 
hydrogen  (H),  helium  (He),  carbon  (C),  nitrogen 
(N),  oxygen  (O),  calcium  (Ca),  and  chlorine  (CI) 
are  easily  obtained  from  the  names.  Notice  that 
He  is  used  for  helium  to  distinguish  it  from  H 
for  hydrogen.  Again,  since  C  is  used  for  carbon, 
the  symbols  for  calcium  and  chlorine  each  have 
a  second  letter  added  to  the  first.  The  table  of 
elements  (inside  the  back  cover  of  the  book)  con- 
tains a  complete  list  of  the  chemical  symbols. 

2-3.2    Chemical  Formulas 

Molecules  are  made  up  of  atoms  in  definite  num- 
bers and  definite  arrangements.  Models  and  sym- 
bols for  the  elements  aid  us  in  showing  the 


*  These  symbols  were  adopted  as  a  result  of  the  urging 
of  an  outstanding  Swedish  chemist,  Jons  Berzelius,  during 
the  first  half  of  the  nineteenth  century.  Until  Berzelius 
lent  his  prestige  as  a  chemist  to  this  system,  many  of  the 
elements  had  several  symbols  in  accordance  with  their 
different  names  in  different  languages. 


sec.  2-3  I  substances:  elements  and  compounds 


31 


A  model  of 
WATtR 


Fig.  2-10.  A  model  of  a  molecule  of  water. 


composition  of  a  molecule.  Figure  2-10  shows  a 
model  of  a  water  molecule.  Experiments  have 
shown  that  the  model  should  contain  two  atoms 
of  hydrogen  and  one  atom  of  oxygen.  The  ad- 
vantage of  such  a  model  is  that  it  shows  also 
the  spatial  arrangement  of  the  atoms.  In  a  mole- 
cule of  water,  each  of  the  two  hydrogen  atoms 
is  connected  to  the  oxygen  atom  in  a  triangular 
arrangement.  How  the  shape  is  determined  and 
how  important  it  is  in  chemistry  will  be  treated 
later  in  the  course. 

The  number  and  kinds  of  atoms  in  a  molecule 
can  also  be  shown  in  a  molecular  formula.  For 
example,  the  water  molecule  is  symbolized 
"H20."  In  this  molecular  formula,  "H"  means 
"hydrogen  atom,"  "O"  means  "oxygen  atom," 
and  the  subscript  "2"  following  "H"  indicates 
there  are  two  hydrogen  atoms  bound  to  the 
single  oxygen  atom.  The  molecular  formula  of 
ammonia,  NH3,  indicates  that  one  molecule  of 
ammonia  contains  one  atom  of  nitrogen  (N)  and 


three  atoms  of  hydrogen  (H).  Experiments  show 
that  oxygen  is  diatomic  (each  molecule  contains 
two  atoms),  hence  its  molecular  formula  is  02. 
Hydrogen  gas  is  diatomic;  its  formula  is  H2. 

Both  the  numbers  and  the  arrangement  of  the 
atoms  in  the  molecule  are  shown  by  a  structural 
formula.  The  structural  formulas,  like  the  models 
we  have  seen,  show  which  atoms  are  attached  to 
each  other.  Thus,  H20  has  the  structural  formula 


not 


or 


H— O— H 
H— H— O 


A 

H H 

In  the  structural  formula  H— O— H,  the  dashes 
indicate  the  connections  between  the  atoms.  The 
connections  between  atoms  are  called  chemical 
bonds.  We  see  that  each  of  the  two  hydrogen 
atoms  is  bound  to  the  oxygen  atom.  Both  of 
the  alternate  arrangements, 


and 


H— H— O 


H H 


agree  with  the  molecular  formula  H20  but  the 
properties  of  water  show  that  the  atoms  are  not 
so  bonded. 

No  written  formula  is  quite  as  effective  as  a 
molecular  model  to  help  us  visualize  molecular 
shape.  Since  chemists  find  that  the  shape  of  a 
molecule  strongly  influences  its  chemical  behav- 
ior, pictures  and  models  of  molecules  are  im- 
portant aids.  A  variety  of  types  of  models  are 


Table  2-V.     chemical  symbols  that  are  not  derivable  from  the  common 

ENGLISH    NAME    OF    THE    ELEMENT 


COMMON  NAME 

SYMBOL 

SYMBOL  SOURCE 

COMMON  NAME 

SYMBOL 

SYMBOL  SOURCE 

antimony 

Sb 

stibnum 

potassium 

K 

kalium 

copper 

Cu 

cuprum 

silver 

Ag 

argentum 

gold 

Au 

aurum 

sodium 

Na 

natrium 

iron 

Fe 

ferrum 

tin 

Sn 

stannum 

lead 

Pb 

plumbum 

tungsten 

W 

wolfram 

mercury 

Hg 

hydrargyrum 

32 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    I    CHAP.    2 


Name         Molecular      Structural 
Formula         Formula 


Models 


Ball-  and- Stick  Ball- and- Spring  Space-Filling 


Hydrogen  H2 


.H 


H' 


Water  H20 


H 


-H 


& 


-AT- 


Ammonia       NH* 


"  / 

H 


-H 


Fig.  2-11.  Different  representations  of  molecules  Hi, 
H,0,  and  NHS. 


commonly  used,  depending  upon  the  emphasis 
needed.  Figure  2-1 1  shows  some  representations 
of  molecules  of  hydrogen,  water,  and  ammonia. 
The  ball-and-stick  and  ball-and-spring  models 
display  clearly  the  bonds  and  their  orientations. 
The  ball-and-spring  models  indicate  molecular 
flexibility.  The  space-filling  models  provide  a 
more  realistic  view  of  the  spatial  relationships 
and  crowding  among  nonbonded  atoms. 

EXERCISE  2-4 

Carbon  dioxide  has  the  formula  C02.  Remem- 
bering that  the  prefix  "di"  means  two,  and  "tri" 
means  three,  write  the  molecular  formula  for 
each  of  the  following  substances:  carbon  disul- 
fide, sulfur  dioxide,  sulfur  trioxide.  (If  you  don't 
know  the  symbol  for  an  element,  use  the  table 
inside  the  back  cover  of  the  book.) 


2-3.3    The  Mole 

Any  sample  of  matter  we  examine  contains  a 
very  large  number  of  atoms.  We  never  work  with 


individual  atoms  or  molecules  but  always  with 
collections  of  these  particles.  Chemists,  there- 
fore, have  selected  a  unit  larger  than  a  single 
atom  or  molecule  for  comparing  amounts  of 
different  materials.  This  unit,  called  the  mole, 
contains  a  very  large  number  of  particles, 
6.02  X  1023.  A  mole  of  oxygen  atoms,  or  a  mole 
of  hydrogen  atoms,  or  a  mole  of  copper  atoms 
contains  6.02  X  1023  atoms  of  the  specified  kind. 
A  mole  of  oxygen  molecules,  02,  contains  2 
moles  of  oxygen  atoms  (2  X  6.02  X  1023  oxygen 
atoms)  because  each  oxygen  molecule  contains 
two  atoms.  A  mole  of  P4  molecules  contains 
4  X  6.02  X  1023  phosphorus  atoms,  that  is,  four 
moles  of  phosphorus  atoms.* 

A  baker  counts  biscuits  in  dozens — a  con- 
venient number.  Money  is  counted  in  dollars — 
one  hundred  cents  is,  again,  a  convenient  num- 
ber. How  did  chemists  choose  to  count  in  terms 
of  moles — the  number  6.02  X  1023  seems  an  odd 
choice.  Why,  for  instance,  didn't  they  settle  on 
some  simpler  number,  such  as  exactly  one  billion 
particles?  There  is  a  reason.  Chemists  prefer  a 
definition  in  terms  of  a  quantity  that  can  be 
measured  readily  and  with  high  accuracy.  Weigh- 

*  If  you  have  difficulty  expressing  numbers  in  terms  of 
powers  of  ten,  refer  to  Appendix  5  in  the  Laboratory 
Manual. 


sec.  2-3  I  substances:  elements  and  compounds 


33 


ing  is  easier  than  counting  when  the  number  of 
particles  to  be  counted  is  so  very  large.  Conse- 
quently, chemists  based  the  definition  of  the  mole 
upon  a  chosen  weight  rather  than  a  chosen  num- 
ber of  particles.  During  the  nineteenth  century, 
chemists  decided  that  the  number  of  molecules 
in  a  sample  of  oxygen  weighing  exactly  32  grams 
would  be  taken  as  a  standard  number.  Thus  de- 
fined, a  mole  is  the  number  of  oxygen  mole- 
cules in  exactly  thirty-two  grams  of  oxygen* 
The  significance  of  a  mole  is  most  usefully  con- 
nected with  this  number  of  particles,  rather  than 
the  weight.  The  number,  found  later  to  be 
6.02  X  W23,  is  called  Avogadro's  number. 
(Avogadro  was  the  first  to  propose  how  to  ob- 
tain equal  numbers  of  molecules  of  different  sub- 
stances.) 


weighed  on  an  ordinary  balance.  For  practical 
purposes,  the  weight  of  a  mole  of  atoms  is  a 
valuable  number.  This  weight  is  called  the  atomic 
weight.  The  atomic  weight  of  an  element  is  the 
weight  in  grams  of  A  vogadro's  number  of  atoms. 

Now  consider  compounds.  Again,  a  useful 
number  to  the  chemist  is  the  weight  of  a  mole  of 
molecules.  This  weight  is  called  the  molecular 
weight  (or  the  molar  weight).  The  molecular 
aeight  of  a  compound  is  the  weight  in  grams  oj 
A  vogadro's  number  of  molecules. 

Consider  the  substance  hydrogen  chloride. 
This  compound  has  the  molecular  formula  HC1. 
A  chemist  working  with  hydrogen  chloride,  HC1, 
must  often  know  the  weight  of  a  mole  of  mole- 
cules (the  molecular  weight).  This  weight  is 
readily  calculated  from  the  atomic  weights  of 
the  two  kinds  of  atoms,  H  and  CI: 


one  molecule 
of  HC1 


contains 


one  atom  of 
hydrogen 


and 


one  atom  of 
chlorine 


one  mole  of  HC1 
molecules 


contains 


one  mole  of 
hydrogen 
atoms 


and 


one  mole  of 
chlorine 
atoms 


the  weight  of 
one  mole  of 
HC1  molecules 


is  the 
same  as 


the  weight  of 
one  mole  of 
hydrogen  atoms 


+ 


the  weight  of 
one  mole  of 
chlorine  atoms 


molecular 
weight  of  HC1 

molecular 
weight  of  HC1 


atomic  weight 
of  hydrogen 

1.01  g 


+ 


+ 


atomic  weight 
of  chlorine 

35.5  g 


molecular 
weight  of  HC1 


36.5  g 


2-3.4    Atomic  and  Molecular  Weights 

Chemists  deal  with  amounts  of  substances  that 
are  readily  measured.  Although  a  chemist  is 
aware  that  the  mass  of  a  single  oxygen  atom  is 
2.656  X  10-23  gram,  he  finds  it  much  more  useful 
to  know  that  a  mole  of  oxygen  atoms  weighs 
16.00  grams.  This  is  an  amount  that  can  be 

*  Recently  the  definition  of  a  mole  has  been  altered  to 
put  it  in  terms  of  measurements  made  with  higher  accu- 
racy. The  change  implied,  —  .0045  %,  is  unimportant  from 
a  chemist's  point  of  view. 


In  a  similar  way,  the  weight  of  a  mole  of  H20 
molecules  is  the  weight  of  two  moles  of  hydrogen 
atoms  plus  the  weight  of  one  mole  of  oxygen 
atoms.  Hence 

molecular  wt.  of  H^O  =  2  X  (atomic  wt.  of  H) 

+  1  X  (atomic  wt.  of  O) 
-  2(1.01)  + (16.00)  g 
=  18.02  g 

Atomic  weights  have  everyday  importance  to 
a  chemist.  Therefore,  the  atomic  weights  must 
be  readily  available.  They  are  listed  both  in  the 


34 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY   I    CHAP.    2 


periodic  table  (inside  the  front  cover)  and  in  the 
table  of  atomic  weights  (inside  the  back  cover). 


EXERCISE  2-5 

Show  that  the  weight  of  a  mole  (the  molecular 
weight)  of  COi  is  44.0  grams  and  that  the  weight 
of  a  mole  of  S02  is  64. 1  grams. 

EXERCISE  2-€ 

In  Experiment  6  you  calculated  the  ratio  of  the 
weight  of  carbon  dioxide  to  the  weight  of  the 
same  volume  of  oxygen.  Oxygen  has  been  as- 
signed a  molecular  weight  of  32.0.  From  the 
molecular  weight  of  oxygen  and  your  measured 
ratio,  calculate  the  molecular  weight  of  carbon 


dioxide.  Estimate  the  uncertainty  in  your  result. 
Compare  to  the  value  obtained  in  Exercise  2-5. 


EXERCISE  2-7 


What  is  the  molecular  weight  of  each  of  the  sub- 
stances sulfur  (formula,  S8),  ammonia  (formula, 
NH3),  and  nitrogen  (a  diatomic  molecule)? 


EXERCISE  2-8 


Calculate  the  weight  of  6.02  X  1023  molecules  of 
carbon  monoxide,  CO. 


EXERCISE  2-9 


How  many  moles  of  iron  atoms  are  present  in 
1.73  grams  of  iron? 


2-4    REVIEW:  THE  ATOMIC  THEORY 

"Gases  are  composed  of  particles"  was  proposed 
in  Chapter  1  as  a  useful  model  to  aid  us  in  dis- 
cussing certain  properties  of  oxygen  gas.  We 
have  continued  to  use  this  model  in  discussing 
other  types  of  information.  First  it  was  tested 
for  applicability  to  other  gases,  ammonia  and 
hydrogen  chloride.  The  assumption  of  molecular 
particles  turned  out  to  fit  the  properties  of  these 
two  gases  as  well,  but  it  was  discovered  that  the 
weights  of  the  particles  of  different  gases  must 
be  different.  Further  studies  of  a  variety  of  prop- 
erties of  gases  (color,  solubility  in  water,  behav- 
ior toward  litmus,  etc.)  caused  us  to  propose  that 
the  molecules  differ  in  their  structures.  We  were 
thus  led  to  the  view  that  molecules  contain 
smaller  building  blocks,  called  atoms. 

Having  proposed  the  existence  of  atoms,  we 
began  representing  the  structure  of  molecules 
through  molecular  models  and  molecular  for- 
mulas. These  models  and  formulas  picture  what 


is  known  about  the  numbers,  types,  and  arrange- 
ments of  atoms  in  the  molecules  represented. 
Our  success  in  treating  gases,  using  this  atomic 
theory,  led  us  to  propose  that  the  atoms  are 
present,  as  well,  in  solids  and  liquids.  Now  we 
postulate  that  atoms  are  present  in  all  matter. 
Thus  a  model  (a  theory)  grows.  As  it  is  tested 
in  an  ever  widening  range  of  experience,  the 
model  often  grows  more  complex.  This  is  offset 
by  the  advantage  of  developing  interrelationships 
among  diverse  phenomena  (that  is,  by  discover- 
ing hidden  likenesses).  The  atomic  theory,  as 
developed  to  correlate  chemical  behavior,  is 
much  more  complicated  than  is  needed  to  ex- 
plain the  simple  gas  behavior  mentioned  in 
Chapter  1.  Nevertheless,  connections  developed 
between  billiard  balls  rebounding  from  a  cushion 
and  the  pressure  in  a  balloon  have  provided  us 
with  a  substantial  start  in  understanding  chem- 
istry. 


QUESTIONS    AND    PROBLEMS 


35 


QUESTIONS  AND  PROBLEMS 

1.  Hydrogen,  helium,  and  carbon  dioxide  are  all 
gases  at  normal  temperatures.  What  differences 
among  the  properties  of  these  gases  account  for 
the  following? 

(a)  Hydrogen  and  helium  are  used  in  balloons 
whereas  carbon  dioxide  is  not. 

(b)  Helium  is  less  dangerous  to  use  in  balloons 
than  is  hydrogen. 

2.  Four  differences  between  helium  gas  and  nitro- 
gen gas  are  listed  below. 

(a)  Dry  air  contains  80%  nitrogen  but  only 
0.0005  %  helium  (by  volume). 

(b)  Helium  is  much  less  dense  than  nitrogen. 

(c)  Helium  has  much  lower  solubility  in  water 
than  nitrogen  has. 

(d)  Helium  is  much  more  expensive  than  nitro- 
gen. 

Which  difference  could  account  for  the  fact  that 
a  diver  is  much  less  likely  to  suffer  from  the 
bends  if  he  breathes  a  mixture  of  80%  helium 
and  20%  oxygen  than  if  he  breathes  air?  (The 
bends  is  a  painful,  sometimes  fatal,  disease 
caused  by  the  formation  of  gas  bubbles  in  the 
veins  and  consequent  interruption  of  blood  flow. 
The  bubbles  form  from  gas  dissolved  in  the 
blood  at  high  pressure.) 

3.  The  most  important  step  in  the  process  for  the 
conversion  of  atmospheric  nitrogen  into  impor- 
tant commercial  compounds  such  as  fertilizers 
and  explosives  involves  the  combination  of  one 
volume  of  nitrogen  gas  with  three  volumes  of 
hydrogen  gas  to  form  two  volumes  of  ammonia 
gas. 

From  these  data  alone  and  Avogadro's  Hy- 
pothesis, how  many  molecules  of  hydrogen  com- 
bine with  one  molecule  of  nitrogen  ?  How  many 
molecules  of  ammonia  are  produced  from  one 
molecule  of  nitrogen  ? 


4.  Gaseous  uranium  hexafluoride  is  important  in 
the  preparation  of  uranium  as  a  source  of 
"atomic  energy." 

A  flask  filled  with  this  gas  is  weighed  under 
certain  laboratory  conditions  (temperature  and 
pressure),  and  the  weight  of  the  gas  is  found  to 
be  3.52  grams.  The  same  flask  is  filled  with  oxy- 
gen gas  and  is  weighed  under  the  same  labora 
tory  conditions.  The  weight  of  the  oxygen  in  the 
flask  is  found  to  be  0.32  gram. 

What  is  the  ratio  of  the  weight  of  one  uranium 
hexafluoride  molecule  to  the  weight  of  an  oxygen 
molecule?  State  any  guiding  principles  needed 
in  answering  this  question. 

5.  Two  volumes  of  hydrogen  fluoride  gas  combine 
with  one  volume  of  the  gas  dinitrogen  difluoride 
to  form  two  volumes  of  a  gas  G. 

(a)  According  to  Avogadro's  Hypothesis,  how 
many  molecules  of  G  are  produced  from  one 
molecule  of  dinitrogen  difluoride? 

(b)  If  X  =  number  of  atoms  in  a  molecule  of 

hydrogen  fluoride, 
Y  =  number  of  atoms  in  a  molecule  of 

dinitrogen  difluoride, 
Z  =  number  of  atoms  in  a  molecule  of  G, 
write  the  relation  among  X,  Y,  and  Z  appro- 
priate to  the  combining  volumes  given. 

(c)  For  each  of  the  following  possible  values  of 
X  and  y,  calculate  the  required  value  of  Z. 

If  X  is      and       Y  is      then      Z  must  be 


(d)  No  odd  value  of  Y  is  suggested  in  question 
(c).  Prove  that  Y  must  be  an  even  integer. 


one  volume  of 

combines 

Ill    111C 

one  volume  of 

gdlClclI    JJ< 

to 

3UCI  11 

two  volumes  of 

gas  A 

with 

gas  B 

form 

gas  C 

For  example, 

one  volume  of 

combines 

one  volume  of 

to 

two  volumes  of 

carbon  dioxide 

with 

carbon  disulfide 

form 

carbon  oxysulfide 

one  volume  of 

combines 

one  volume  of 

to 

two  volumes  of 

hydrogen 

with 

chlorine 

form 

hydrogen  chloride 

36 


A    SCIENTIFIC    MODEL:    THE    ATOMIC    THEORY    I    CHAP.    2 


If  X  =  number  of  atoms  in  a  molecule  of  A, 
Y  =  number  of  atoms  in  a  molecule  of  B,  and 
Z  =  number  of  atoms  in  a  molecule  of  C, 

(a)  Show  that  if  X  is  even,  Y  must  be  even. 

(b)  Show  that  if  X  is  odd,  Y  must  be  odd. 

7.  A  pure  white  substance,  on  heating,  forms  a 
colorless  gas  and  a  purple  solid.  Is  the  substance 
an  element  or  a  compound? 

8.  What  do  the  following  symbols  represent?  K, 
Ca,  Co,  CO,  Pb. 

9.  Write  formulas  for 

silicon  dioxide  (common  sand,  or  silica), 
sulfur  dichloride, 
nitrogen  trifluoride, 
aluminum  trifluoride, 
dinitrogen  difluoride. 

10.  (a)  Write  formulas  for 

hydrogen  chloride, 
hydrogen  bromide, 
hydrogen  iodide, 
boron  trichloride, 
carbon  tetrachloride, 
nitrogen  trichloride, 
oxygen  dichloride. 
(b)  Locate  in  the  table  inside  the  front  cover  the 
symbol  for  each  element  involved  in  these, 
compounds. 


11. 


For  each  of  the  following  substances  give  the 
name  of  each  kind  of  atom  present  and  the  total 
number  of  atoms  represented  in  the  formula 
shown. 


Name 


Formula 


(a)  graphite  (pencil  lead) 

C 

(b)  diamond 

C 

(c)  sodium  chloride  (table  salt) 

NaCl 

(d)  sodium  hydroxide 

NaOH 

(e)  calcium  hydroxide 

Ca(OH)> 

(f)  potassium  nitrate 

KNOs 

(g)  magnesium  nitrate 

Mg(N03)2 

(h)  sodium  sulfate 

Na2S04 

(i)   calcium  sulfate 

CaS04 

12.  All  of  the  following  substances  are  called 
"acids."  What  element  do  they  have  in  com- 
mon? 


(b)  hydrochloric  acid  (or, 
hydrogen  chloride) 

(c)  hydrofluoric  acid  (or, 
hydrogen  fluoride) 

(d)  sulfuric  acid 

(e)  phosphoric  acid 


HC1 

HF 

H2S04 

H3P04 


(a)  nitric  acid 


HNO3 


13.  Here  are  the  names  of  some  common  chemicals 
and  their  formulas.  What  elements  does  each 
contain? 

(a)  hydrogen  peroxide  H202 

(b)  jeweler's  rouge  Fe203 

(c)  light  bulb  filament  W 

(d)  tetraethyl  lead  Pb(C2H5)4 

(e)  baking  soda  NaHC03 
(0  octane  C8Hi8 
(g)  household  gas  CH4 

14.  (a)  What   does   the   molecular    formula   CBr4 

mean? 
(b)  What  information  is  added  by  the  following 
structural  formula? 

Br 

I 
Br— C— Br 

I 
Br 

15.  How  many  particles  are  there  in  a  mole? 

16.  A  stone  about  the  size  of  a  softbail  weighs 
roughly  a  kilogram.  How  many  moles  of  such 
stones  would  be  needed  to  account  for  the  entire 
mass  of  the  Earth,  about  6X10"  grams? 

17.  If  we  had  one  mole  of  dollars  to  divide  among 
all  the  people  in  the  world,  how  much  would 
each  of  the  three  billion  inhabitants  receive? 

18.  How  many  moles  of  atoms  are  in 

(a)  9.0  grams  of  aluminum 

(b)  0.83  gram  of  iron 

Answer,  (a)  ^  mole. 

19.  The  most  delicate  balance  can  detect  a  change 
of  about  10-8  gram.  How  many  atoms  of  gold 
would  be  in  a  sample  of  that  weight  ? 

20.  How  many  moles  of  oxygen  atoms  are  in  one 
mole  of  nitric  acid  molecules?  Of  sulfuric  acid 
molecules  ? 

21.  Determine  the  weight,  in  grams,  of  one  silver 
atom. 

Answer.  1.79  X  10-22  gram. 


QUESTIONS    AND    PROBLEMS 


37 


22.  Write  the  formulas  for  the  following  compounds 
and  give  the  weight  of  one  mole  of  each:  carbon 
disulfide,  sulfur  hexafluoride,  nitrogen  trichlo- 
ride, osmium  tetroxide. 

23.  Consider  the  following  data 

Element  Atomic  Weight 


12.01 

35.5 


A  and  B  combine  to  form  a  new  substance,  X. 
If  four  moles  of  B  atoms  combine  with  one  mole 
of  A  to  give  one  mole  of  X,  then  the  weight  of 
one  mole  of  X  is 

(a)  47.5  grams 

(b)  74.0  grams 

(c)  83.0  grams 

(d)  154.0  grams 

(e)  166.0  grams 

24.  Calculate  the  molecular  weight  for  each  of  the 
following:  SiF4,  HF,  Cl2,  Xe,  N02. 

25.  If  1^  moles  of  hydrogen  gas  (H2)  react  in  a  given 
experiment,  how  many  grams  of  H2  does  this 
represent  ? 

26.  How  many  moles  are  contained  in  49  grams  of 
pure  H2S04? 

Answer.  0.50  mole. 

27.  A  chemist  weighs  out  10.0  grams  of  water,  10.0 
grams  of  ammonia,  and  10.0  grams  of  hydrogen 
chloride  (hydrochloric  acid).  How  many  moles 
of  each  substance  does  he  have? 

28.  (a)  The  ratio  of  the  weight  of  a  liter  of  chlorine 

gas  to  the  weight  of  a  liter  of  oxygen  gas, 
both  measured  at  room  temperature  and 
pressure,  is  2.22.  Calculate  the  molecular 
weight  of  chlorine, 
(b)  How  does  this  value  compare  with  the 
atomic  weight  of  chlorine  found  in  the  chart 
of  atomic  weights?  What  is  the  formula  of  a 
molecule  of  chlorine? 


29.  A  flask  of  gaseous  CCL,  was  weighed  at  a  meas- 
ured temperature  and  pressure.  The  flask  was 
flushed  and  then  filled  with  oxygen  at  the  same 
temperature  and  pressure.  The  weight  of  the 
CCI4  vapor  will  be  about 

(a)  the  same  as  that  of  oxygen, 

(b)  one-fifth  as  heavy  as  the  oxygen, 

(c)  five  times  as  heavy  as  the  oxygen, 

(d)  twice  as  heavy  as  the  oxygen, 

(e)  one-half  as  heavy  as  the  oxygen. 

30.  Suppose  chemists  had  chosen  a  billion  billion 
(1018)  as  the  number  of  particles  in  one  mole. 
What  would  the  molecular  weight  of  oxygen 
gas  be? 

31.  One  volume  of  hydrogen  gas  combines  with  one 
volume  of  chlorine  gas  to  produce  two  volumes 
of  hydrogen  chloride  gas  (all  measured  at  the 
same  temperature  and  pressure).  A  variety  of 
other  types  of  evidence  suggests  that  hydrogen 
is  an  element  and  that  its  molecules  are  diatomic. 

(a)  Which  one  of  the  following  possible  molecu- 
lar formulas  for  the  substance  chlorine  is  not 
consistent  with  the  volumes  that  combine? 
(Use  only  the  data  given  here;  do  not  pre- 
sume the  molecular  formula  of  hydrogen 
chloride.) 

(i)  CI, 

(ii)  CU 

(iii)  H2C12 

(iv)  H3C12 

(v)  H4C12 

(b)  For  each  formula  in  part  (a)  that  is  con- 
sistent with  the  combining  volumes  data  (and 
the  formula  H2  for  hydrogen),  calculate  the 
molecular  weight  indicated  by  that  formula. 

(c)  For  each  acceptable  formula  in  part  (a)  pre- 
dict the  molecular  formula  for  the  substance 
hydrogen  chloride. 


CHAPTER 


3 


Chemical 
Reactions 


We  might  as  well  attempt  to  introduce  a  new  planet  into  the  solar  sys- 
tem, or  to  annihilate  one  already  in  existence,  as  to  create  or  destroy  a 
particle  of  hydrogen.  All  the  changes  we  can  produce  consist  in  sep- 
arating particles  that  are  in  a  state  of  •  •  •  combination,  and  joining 
those  that  were  previously  at  a  distance. 

JOHN    DAL  TON,    1810 


In  Experiment  5  you  compared  the  magnitude  of 
the  heat  released  when  melted  wax  solidifies  to 
the  heat  released  when  the  same  amount  of  wax 
is  burned.  In  each  case  an  obvious  change  occurs 
(liquid  changes  to  solid  or  solid  burns  to  a  gas) 
and  it  is  accompanied  by  a  measurable  heat  re- 
lease. These  are  likenesses  between  solidification 
and  combustion.  More  apparent,  however,  are 
striking  differences.  Through  experiments  on 
warming  and  cooling  wax,  you  know  that  after 
wax  melts  to  a  liquid  the  solid  wax  can  be  re- 
covered merely  by  recooling.  After  combustion 
of  the  wax,  however,  cooling  the  gases  produced 
does  not  restore  the  wax.  Instead,  the  products, 


carbon  dioxide  and  liquid  water  (and  perhaps 
some  soot),  bear  no  resemblance  to  wax.  Equally 
dramatic  is  the  contrast  of  the  heat  effects.  The 
heat  released  during  combustion  is  well  over  two 
hundred  times  greater  than  the  heat  released 
during  solidification. 

Because  of  these  differences,  chemists  differ- 
entiate these  two  kinds  of  change.  We  have 
already  named  the  solidification  of  wax — in  Sec- 
tion 1-1.2  we  called  this  type  of  change  a  phase 
change.  A  change  like  combustion,  with  its  much 
larger  heat  effects,  is  called  a  chemical  change  or 
a  chemical  reaction. 


3-1    PRINCIPLES  OF  CHEMICAL  REACTIONS 


The  central  activity  of  the  chemist  is  to  explore 
and  exploit  chemical  changes.  Sometimes  his 
38 


wish  is  to  cause  a  change,  sometimes  to  prevent 
it,  in  a  system  of  interest.  Always  he  wishes  to 


SEC.    3-1    I    PRINCIPLES    OF    CHEMICAL    REACTIONS 


39 


understand  and  control  the  chemical  changes 
that  might  occur. 

We  shall  begin  our  study  of  chemical  changes 
with  a  simple  chemical  reaction  that  forms  a 
familiar  substance — water. 

3-1.1    Formation  of  Water  from  Hydrogen 
and  Oxygen 

In  Section  2-2.2  it  was  mentioned  that  two  vol- 
umes of  hydrogen  combine  with  one  volume  of 
oxygen.  Water  is  produced.  The  reaction  gives 
off  heat — a  large  amount  of  heat,  as  in  the  com- 
bustion of  a  candle.  The  product,  water,  is  not 
at  all  like  the  starting  materials,  hydrogen  and 
oxygen.  Hence,  the  change  that  occurs  when 
hydrogen  and  oxygen  combine  should  be  clas- 
sified as  a  chemical  reaction. 

In  terms  of  the  atomic  theory,  we  begin  with 
molecules  of  hydrogen  and  molecules  of  oxygen. 
After  reaction,  we  find  molecules  of  water.  The 
bonds  between  atoms  in  the  reacting  substances 
are  broken,  and  the  atoms  rearrange  to  form  the 
new  bonds  in  the  product  molecules.  These 
changes  are  readily  pictured  with  the  aid  of  our 
molecular  models.  In  Figure  3-1  two  hydrogen 
molecules  (four  atoms)  and  one  oxygen  molecule 
(two  atoms)  are  represented  on  the  left.  If  these 
molecules  are  to  react  to  form  water,  the  bonds 
between  the  atoms  in  the  oxygen  molecule  and 
in  the  hydrogen  molecules  must  be  broken.  Then, 
the  atoms  can  rearrange  themselves  to  form  two 
water  molecules.  Notice  that  the  atoms  are  re- 
arranged as  a  result  of  the  reaction  but  that  the 
total  number  of  atoms  does  not  change. 


water  could  be  formed?  What  would  be  left 


over 


EXERCISE  3-2 


One  million  oxygen  molecules  react  with  suffi- 
cient hydrogen  molecules  to  form  water  mole- 
cules. How  many  water  molecules  are  formed? 
How  many  hydrogen  molecules  are  consumed? 


Extending  Figure  3-1,  you  can  reason  that  the 
production  of  100  molecules  of  water  requires 
100  molecules  of  hydrogen  and  50  molecules  of 
oxygen.  Also,  to  produce  one  mole  of  water 
(6.02  X  1023  molecules)  you  need  one  mole  of 
hydrogen  gas  (6.02  X  1023  molecules)  and  one- 
half  mole  of  oxygen  (3.01  X  1023  molecules). 
These  results  are  summarized  in  Table  3-1. 


Table  3-1. 

REACTING  AMOUNTS  OF  HYDROGEN 
AND  OXYGEN  TO  FORM  WATER 

Amounts  Amounts 

of  Reacting  of 

Substances  Product 

HYDROGEN    OXYGEN      WATER 


(a)  In  numbers 

2 

1 

2 

of  molecules 

4 

2 

4 

100 

50 

100 

6.02  X  10» 

3.01  X  10a 

6.02  X  10a 

(b)  In  numbers 

1 

* 

1 

of  moles 

2 

1 

2 

10 

5 

10 

EXERCISE  3-1 

Suppose  ten  hydrogen  molecules  and  ten  oxygen 
molecules  are  mixed.  How  many  molecules  of 


Fig.  3- J.  Formation  of  water  molecules  from  hydro- 
gen molecules  and  oxygen  molecules. 


reacts    with 


The  reaction  between  hydrogen  gas  and  oxy- 
gen gas  proceeds  more  quickly  if  we  mix  the 
gases  and  then  ignite  the  mixture  with  a  spark. 
A  violent  explosion  results.  Even  so,  the  quan- 
tity of  product,  water,  and  the  heat  evolved  are 
the  same  per  mole  of  hydrogen  reacting  as  in 
controlled  burning. 


# 


to  form 


40 


CHEMICAL  REACTIONS  I  CHAP.  3 


If  we  react  one  mole  of  pure  hydrogen  and 
one-half  mole  of  pure  oxygen,  one  mole  of  water 
is  produced.  The  quantity  of  heat  produced  when 
one  mole  of  water  is  formed  is  68,000  calories. 
If  we  mix  only  0.025  mole  of  pure  hydrogen,  only 
(2)  X  (0.025)  mole  of  oxygen  is  needed.  The 
amount  of  water  produced  is  0.025  mole.  If  only 
0.025  mole  of  water  is  produced,  only  (0.025)  X 
(68,000)  calories  or  about  1700  calories  of  heat 
are  released. 

The  source  of  this  heat  energy  must  be  the  re- 
actants  (hydrogen  and  oxygen)  themselves  since 
no  heat  was  supplied  externally  other  than  that 
needed  to  ignite  the  mixture.  We  may  conclude 
that  the  water  has  less  energy  than  did  the  re- 
actants  used  to  make  it.  Such  a  reaction  in  which 
energy  is  released  is  called  an  exothermic  re- 
action. The  quantity  of  energy  produced  when 
one  mole  of  hydrogen  is  burned,  68,000  calories 
(68  kcal*),  is  called  the  molar  heat  of  com- 
bustion of  hydrogen. 


EXERCISE  3-3 

How  much  heat  is  released  when  two  moles  of 
hydrogen  burn?  One-half  mole? 


3-1.2    Decomposition  of  Water 

We  can  decompose  the  water  in  a  solution  of 
water  and  sulfuric  acid  by  passing  an  electric 
current  through  the  solution  in  an  electrolysis 
apparatus,  as  shown  in  Figure  3-2.  In  this  ap- 
paratus, two  conductors  (called  electrodes)  are 
immersed  in  the  liquid.  When  the  electrodes  are 
connected  to  a  source  of  electrical  energy  hydro- 
gen gas  appears  at  one  electrode  and  oxygen  gas 
appears  at  the  other.  If  we  operate  the  apparatus 
until  one  mole  of  water  has  decomposed,  one 
mole  of  hydrogen  gas  and  one-half  mole  of  oxy- 
gen gas  are  produced.  We  observe  also  that 
energy  (electrical  energy  in  this  case)  is  needed 
to  cause  the  water  to  decompose.  If  energy  is 


*  1  kilocalorie  (1  kcal)  =  1000  calories.  The  prefix 
"kilo-"  always  means  1000.  Thus,  1  kilogram  =  1000 
grams;  1  kilometer  =  1000  meters. 


Fig.  3-2.     Electrolytic  decomposition  of  water. 

absorbed  in  a  reaction,  the  reaction  is  called 
endothermic. 

Now  we  can  compare  the  formation  and  de- 
composition of  water.  As  shown  graphically  in 
Figure  3-3,  the  chemical  change  involved  in  the 
formation  of  water  is  exactly  the  reverse  of  that 
involved  in  the  decomposition  of  water.  These 
changes  are  governed  by  simple  rules.  On  the 
left  we  find  two  atoms  of  oxygen,  and  on  the 
right  we  find  two  atoms  of  oxygen.  On  the  left 
we  find  four  atoms  of  hydrogen,  and  on  the 
right  we  find  four  atoms  of  hydrogen.  We  see 
that  atoms  are  neither  gained,  nor  are  they  lost. 
In  chemical  reactions,  atoms  are  conserved. 

The  number  of  molecules  of  H2  needed  to 
react  with  one  molecule  of  02  is  the  number 
needed  to  produce  two  molecules  of  H20.  If  two 
molecules  of  H20  are  formed,  four  atoms  of  hy- 
drogen are  needed.  Two  molecules  of  H2  contain 
four  atoms  of  hydrogen.  Remember,  in  chemical 
reactions,  atoms  are  conserved. 

3-1.3    Conservation  of  Mass 

Belief  in  the  conservation  of  atoms  is  based  upon 
a  generalization  that  has  stood  the  test  of  many 


SEC.    3-2    |    EQUATIONS    FOR    CHEMICAL    REACTIONS 


41 


decades.  Matter  can  be  neither  created  nor  de- 
stroyed. Since  we  often  measure  a  quantity  of 
matter  in  terms  of  its  mass  (by  weighing,  for 
example)  we  may  say  that  mass  is  conserved. 
Thus,  one  mole  of  liquid  water  weighs  18.0 
grams;  in  the  decomposition  of  one  mole  of 
water,  2.0  grams  of  hydrogen  and  16.0  grams  of 
oxygen  are  produced: 

Fig.  3-3.  The  reactions  of  formation  and  decomposi- 
tion of  water  shown  with  molecular  models. 


1  mole  of 

liquid  water, 

H20 

18.0  g 
18.0  g 


1  mole  of  ,   5  mole  of 
hydrogen,        oxygen, 


H2 


rO. 


2.0  g     +     16.0  g 
18.0  g 


(0 


Since  the  mass  of  a  mole  of  water  is  the  sum 
of  the  masses  of  the  atoms  in  the  mole  of  water, 
the  conservation  of  mass  implies  conservation  of 
atoms. 


reacts    with 


# 


to  form 


is   decomposed    to 


# 


and 


3-2  EQUATIONS  FOR  CHEMICAL  REACTIONS 


The  graphic  representations  we  have  used  for 
reactions  help  us  visualize  the  rearrangement  of 
atoms  in  the  reactions.  By  a  slight  change  we  can 
show  the  results  in  a  less  detailed  but  simpler 
way.  Chemical  formulas  can  be  used  rather  than 
drawings  of  the  atoms  and  molecules.  Thus,  the 
formula  of  elementary  hydrogen  is  H2,  the  for- 
mula of  elementary  oxygen  is  02,  and  the  formula 
of  water  is  H20.  By  using  the  formulas  to  repre- 
sent the  molecules,  we  can  replace  the  diagram 
of  Figure  3-3  by  the  following  expressions:* 


2  H2  +  1  02 

4  H2  +  2  02 
2H20 


2H20 
4H.0 
1  02  +  2  H2 


(2) 
(i) 
(4) 


Such  expressions  are  called  chemical  equations. 
Notice  that  we  show  two  molecules  of  a  sub- 
stance by  writing  the  coefficient  2  before  the 

*  Many  chemists  use  an  "equals"  sign  in  place  of  the 
arrow.  Thus,  equation  (2)  would  be  written: 
2  H2  +  1  02  =  2  HjO 


formula.  The  coefficient  2  before  the  formula 
H20  means  two  molecules — it  applies  to  every 
symbol  in  the  formula.  In  two  molecules  of 
water,  there  are  six  atoms,  four  of  hydrogen  and 
two  of  oxygen.  Notice  also  that  we  can  change 
equation  (2)  to  equation  (5)  merely  by  doubling 
all  the  coefficients.  We  can  change  equation  (2) 
to  equation  (4)  by  reversing  it.  Equation  (2)  rep- 
resents the  formation  of  water;  equation  (4)  rep- 
resents the  decomposition  of  water. 

All  equations  are  based  upon  the  conservation 
of  atoms.  Every  symbol,  when  multiplied  by  the 
subscript  after  it  and  the  coefficient  before  the 
formula,  must  appear  as  often  on  the  left  side  of 
the  equation  as  on  the  right. 

Natural  gas  contains  mainly  the  substance 
methane,  with  the  formula  CH4.  The  chemical 
equation  for  the  burning  of  methane  is 

1  CH4  +  2  02  — >-  1  C02  +  2  H20  (5) 

The  number  of  atoms  in  the  reactants  equals  the 


42 


CHEMICAL    REACTIONS    I    CHAP.    3 


number  of  atoms  in  the  products.  One  carbon 
atom,  four  hydrogen  atoms,  and  four  oxygen 
atoms  are  represented  on  each  side  of  the  equa- 
tion. 

Because  of  the  relation  between  a  molecule 
and  a  mole,  we  can  read  equation  (5)  in  either 
of  two  ways:  (1)  "one  molecule  of  methane  reacts 
with  two  molecules  of  oxygen  to  form  one  mole- 
cule of  carbon  dioxide  and  two  molecules  of 
water,"  or  (2)  "one  mole  of  methane  reacts  with 
two  moles  of  oxygen  to  form  one  mole  of  carbon 
dioxide  and  two  moles  of  water." 


EXERCISE  3-4 

Write  an  equation  containing  the  information 
expressed  in  your  answer  to  Question  1  of  Ex- 
periment 7. 


3-2.1    Writing  Equations  for  Reactions 

How  can  you  write  the  equations  for  a  reaction 
without  first  drawing  a  picture?  Remember  that 
to  do  either  you  must: 

(1)  know  what  reactants  are  consumed  and  what 
products  form; 

(2)  know  the  correct  formula  of  each  reactant 
and  each  product; 

(3)  satisfy  the  law  of  conservation  of  atoms. 

Consider  the  reaction  for  the  burning  of  mag- 
nesium to  form  magnesium  oxide.  Magnesium 
metal  and  magnesium  oxide  are  solids.  They 
have  the  formulas  Mg  and  MgO,  respectively. 
In  preparation  for  writing  the  equation,  we  write 
the  formulas  for  the  reactants  on  the  left  and  the 
formula  for  the  product  on  the  right: 

Mg  +  02  — >-  MgO  (a) 

Expression  (a)  does  not  yet  conserve  atoms.  We 
must  find  numerical  coefficients  to  place  before 
each  formula  so  that  there  are  the  same  number 
of  atoms  of  each  element  on  the  left  side  of  the 
equation  as  there  are  on  the  right.  The  process  of 
finding  these  coefficients  is  called  balancing  the 
equation.  For  simple  reactions,  it  is  an  easy  and 
logical  process. 


First,  we  may  begin  by  choosing  one  mole  of 
oxygen  as  the  amount  of  this  reactant  consumed 
in  (a): 

Mg  +  102  — >-  MgO        (b) 

But  one  mole  of  02,  with  its  two  moles  of  oxygen 
atoms,  will  form  two  moles  of  magnesium  oxide: 

Mg  +  102  — >-  2MgO        (c) 

Two  moles  of  magnesium  oxide  require  two 
moles  of  magnesium  metal.  Thus, 

2Mg  +  102  — >-  2MgO        (d)        (<5) 

Equation  (6)  is  a  chemical  equation — since  atoms 
are  conserved,  it  is  said  to  be  balanced. 

We  could  have  decided  to  begin  by  choosing 
one  mole  of  magnesium  metal  as  the  amount  of 
reactant  consumed  in  (a): 

lMg  +  02  — >-  MgO        (bf) 

One  mole  of  magnesium  contains  a  mole  of 
atoms,  hence  will  form  one  mole  of  magnesium 
oxide: 

lMg  +  02  — >-  lMgO       (c') 

One  mole  of  magnesium  oxide  contains  one  mole 
of  oxygen  atoms,  the  number  contained  in  one- 
half  mole  of  oxygen  molecules.  Thus, 

lMg  +  i02  — ^  MgO       {d')       (7) 

Equation  (7)  is  also  a  chemical  equation — again 
atoms  are  conserved.  It  is  just  as  correct  an  ex- 
pression for  the  burning  of  magnesium  as  is  (5). 
To  show  this,  we  can  multiply  (7)  by  2  to  obtain 
equation  (<5).  We  can  always  multiply  all  the  co- 
efficients by  a  common  factor  or  divide  by  a 
common  factor  and  obtain  equally  valid  equa- 
tions. 

In  equation  (<5)  the  coefficient  1  may  be 
dropped  but  it  is  never  wrong  to  retain  it. 

3-2.2    Other  Examples  of  Chemical  Equations 

Hydrogen  gas,  H2,  and  chlorine  gas,  Cl2,  react 
to  form  hydrogen  chloride  gas,  HC1.*  The  re- 
actants are  H2  and  Cl2;  the  product  is  HC1: 

1H2  +  Cl2  — +■  HC1 


*  Hydrogen  chloride,  HC1,  dissolved  in  water,  is  com- 
monly called  hydrochloric  acid. 


SEC.    3-2    I    EQUATIONS    FOR    CHEMICAL    REACTIONS 


43 


If  we  base  the  reaction  upon  one  mole  of  H2,  we 
see  that  conservation  of  hydrogen  atoms  requires 
that  two  moles  of  HC1  be  produced: 


1H2  +  C12 


2HC1 


Now  the  product,  2HC1,  contains  two  moles  of 
chlorine  atoms.  This  is  just  the  number  of  chlo- 
rine atoms  in  one  mole  of  chlorine.  The  reaction 
is  balanced.  We  may  write 


or 

or 


1H2  +  1C12  — ►-  2HC1 
H2  +  Cl2  — ■+-  2HC1 
H2+    Cl2    =    2HC1 


(5) 


EXERCISE  3-5 


Ammonia  gas,  NH3,  can  be  burned  with  oxygen 
gas,  02,  to  give  nitrogen  gas,  N2,  and  water,  H20. 
See  if  you  can  follow  the  logic  of  the  following 
steps  in  balancing  this  reaction. 


NH3+    02 

NH3+    02 

2NH3+    02 

2NH3+    02 

2NH3  +  |02 


N2+  H20 
1N2+  H20 
1N2+  H20 
1N2  +  3H20 
1N2  +  3H20 


(9) 


State  briefly  what  was  done  in  each  step. 


The  molecular  formula  for  the  substance  form- 
aldehyde is  H2CO.  Formaldehyde  burns  to  form 
carbon  dioxide  and  water.  What  equation  rep- 
resents this  reaction? 

Again  we  begin  by  writing  the  formulas  for 
reactants  and  products: 


lH2CO  +  02 


C02  +  H20 


Suppose  we  burn  one  molecule  of  formalde- 
hyde. The  one  carbon  atom,  the  two  hydrogen 
atoms,  and  one  oxygen  atom  in  the  H2CO  mole- 
cule must  appear  in  the  products.  Since,  among 
the  products,  carbon  atoms  appear  only  in  car- 
bon dioxide,  there  must  be  one  molecule  of  C02: 

1H2C0  +  02  — +-  1C02  +  H20 

Since  hydrogen  atoms  appear  in  only  one  of  the 
products,  water,  there  must  be  one  molecule  of 


water  to  accommodate  the  two  atoms  of  hydro- 
gen. Now  we  have 

lH2CO  +  y02  — *-  1C02  +  1H20 

Notice  that  we  have  not  yet  determined  the 
coefficient  of  02;  we  have  designated  it  as  y  to 
remind  ourselves  of  this.  Since  the  oxygen  atoms 
must  also  be  conserved  and  three  are  required  for 
the  products,  three  oxygen  atoms  must  have  been 
present  in  the  reactants.  One  oxygen  atom  was 
present  in  the  molecule  of  formaldehyde,  so  two 
more  are  required.  It  follows  that  y  must  be  1. 

We  now  have  the  balanced  equation 


lH2CO  +  102 


1C02  +  1H20        (70) 


EXERCISE  3-6 

A  paraffin  candle  burns  in  air  to  form  water  and 
carbon  dioxide.  Paraffin  is  made  up  of  molecules 
of  several  sizes.  We  shall  use  the  molecular  for- 
mula C28H52  as  representative  of  the  molecules 
present.  One  mole  of  candle  contains  the  Avo- 
gadro  number  of  these  molecules. 


Formulas  of  Reactants 
Q5Hs2  -+-  02  — 


Formulas  of  Products 
H20  +  C02 


Suppose  one  mole  of  paraffin  (which  weighs  353 
grams)  is  burned.  Using  the  method  shown  in 
the  preceding  example,  we  obtain 


lQsHs;.  +  v02 


26H20  +  25C02 


We  still  have  not  determined  the  coefficient 
for  02.  Since  76  O's  are  required  for  the  products 
[26  +  (2  X  25)  =  76],  they  must  have  been  pres- 
ent in  the  reactants.  Show  that  it  follows  that  y 
must  be  38: 


C25H52  +  3802 


26H20  +  25C02      (//) 


Usually  it  is  more  useful  to  think  of  equations 
in  terms  of  moles  rather  than  molecules  since  a 
mole  is  a  weighable  amount.  In  equation  (6)  two 
moles  of  magnesium  weigh  48.6  grams;  one 
mole  of  oxygen  weighs  32.0  grams;  two  moles 
of  MgO  weigh  80.6  grams.  Mass  is  conserved: 
48.6  grams  +  32.0  grams  =  80.6  grams. 

In  equation  (77)  the  3802  is  usually  read  as 


44 


CHEMICAL  REACTIONS  I  CHAP.  3 


38  moles,  not  38  molecules;  "38  moles  of  oxygen 
gas"  has  experimental  meaning.  They  weigh 
38  X  32  grams  =1216  grams. 

3-2.3    Calculations  Based  upon  Chemical 
Equations 

Equations  give  us  all  the  information  we  need 
for  computing  the  weights  of  the  substances  con- 
sumed or  produced  in  chemical  reactions.  Sup- 
pose we  wish  to  know  how  many  moles  of  water 
are  produced  when  68  grams  of  ammonia  are 
burned.  Equation  (9)  represents  the  reaction: 

2NH3  +  f02  — >-  1N2  +  3H20  (9) 

One  mole  of  ammonia  weighs  17  grams.  Two 
moles  of  ammonia,  weighing  34  grams,  produces 
three  moles  of  water.  We  wish  to  burn  68  grams 
of  ammonia.  How  many  moles  is  this? 

68  grams 


17  grams/mole 
Hence  we  can  write 
2NH3       +  f02 

two  moles 
of  ammonia 

SO, 

four  moles 
of  ammonia 


=  4.0  moles  of  ammonia 


produce 


IN,  +       3H20 

three  moles 
of  water 


produce 


six  moles 
of  water 


We  see  that  68  grams  (four  moles)  of  ammonia 
produce  six  moles  of  water. 

Suppose  we  wish  to  know  how  many  grams 
of  water  are  produced  from  the  burning  of  one- 
half  mole  of  paraffin.  Equation  (77)  shows  the 
reaction: 


IQ5H5,    +  3802 

1  mole  of 
C25H52 

\  mole  of 
Q25H52 


produces 


26H20  +  25C02 

26  moles 
of  H20 


UD 


Pr0duces  tfSSo8 


Since  one  mole  of  water  weighs  18  grams,  the 
weight  of  13  moles  of  water  is  (13  moles)  X 
(18g/mole)  =  234  g. 


EXERCISE  3-7 

Show  that  3.80  moles  of  oxygen  are  needed  to 
burn  35.3  g  of  paraffin  by  reaction  (77). 

EXERCISE  3-8 

How  many  moles  of  oxygen,  02,  are  required  to 
produce  242  grams  of  magnesium  oxide  by  equa- 
tion (<*)? 

2Mg  +  102  — ■*-  2MgO  (5) 

EXERCISE  3-9 

Write  the  equation  for  the  reaction  which  took 
place  in  Experiment  8,  Part  II.  What  was  the 
residue  you  obtained  on  evaporation  of  the  solu- 
tion in  beaker  number  2? 

EXERCISE  3-10 

In  Experiment  8  you  determined  the  number  of 
moles  of  silver  chloride  formed  in  the  reaction 
of  some  sodium  chloride  with  a  known  amount 
of  silver  nitrate.  How  many  moles  of  sodium 
chloride  reacted  with  the  silver  nitrate?  Compare 
this  with  the  number  of  moles  of  sodium  chlo- 
ride you  added. 


QUESTIONS  AND  PROBLEMS 


One  volume  of  hydrogen  gas  combines  with  one 
volume  of  chlorine  gas  to  give  two  volumes  of 
hydrogen  chloride  gas.  On  the  basis  of  many  re- 
actions, we  have  learned  that  the  molecular 
formulas  are,  for  hydrogen,  H2,  for  chlorine,  Cl2, 
and  for  hydrogen  chloride,  HC1.  The  reaction, 
in  symbols,  is 

H2  +  Cl2  — >-  2HC1 


(a)  According  to  this  reaction,  how  many  mole- 
cules of  hydrogen  chloride,  HC1,  can  be 
formed  from  one  molecule  of  hydrogen,  H2? 

(b)  How  many  moles  of  hydrogen  chloride,  HC1, 
can  be  formed  from  one  mole  of  hydro- 
gen, H2? 

(c)  Four  molecules  of  chlorine,  Cl2,  will  produce 
how  many  molecules  of  HC1? 


QUESTIONS    AND    PROBLEMS 


45 


(d)  Eight  moles  of  hydrogen  chloride  are  formed 
from  how  many  moles  of  Cl2? 

2.  The  reaction  between  nitric  oxide,  NO,  and  oxy- 
gen, 02,  is  written 

2NO  +  02  — *■  2N02 

(a)  Two  molecules  of  nitric  oxide  give  how  many 
molecules  of  nitrogen  dioxide,  N02  ? 

(b)  Two  moles  of  NO  give  how  many  moles  of 
N02? 

(c)  How  many  moles  of  oxygen  atoms  are  there 
in  two  moles  of  NO? 

(d)  How  many  moles  of  oxygen  atoms  are  there 
in  one  mole  of  02? 

(e)  How  many  moles  of  oxygen  atoms  are  there 
in  two  moles  of  N02  ? 

(0  Use  the  answers  to  parts  (c),  (d),  and  (e)  to 
verify  that  the  reaction  is  written  so  as  to 
conserve  oxygen  atoms. 

3.  (a)  Write  the  equation  for  the  reaction  between 

nitrogen  and  hydrogen  to  give  ammonia  on 
the  basis  of  your  answer  to  Problem  3  of 
Chapter  2,  and  assuming  the  following  mo- 
lecular formulas:  nitrogen,  N2;  hydrogen, 
H2;  ammonia,  NH3. 

(b)  Verify  that  your  equation  conserves  nitrogen 
atoms. 

(c)  Verify  that  your  equation  conserves  hydro- 
gen atoms. 

4.  When  ammonia  is  decomposed  into  nitrogen  and 
hydrogen,  the  reaction  absorbs  heat.  Written  in 
terms  of  moles,  the  equation  is 

2NH3  +  22  kcal  — »-  N2  +  3H2 

(a)  Two  moles  of  ammonia  produce  how  many 
moles  of  nitrogen  ? 

(b)  The  production  of  one  mole  of  nitrogen  ab- 
sorbs how  much  heat? 

(c)  The  production  of  nine  moles  of  hydrogen, 
H2,  absorbs  how  much  heat? 

(d)  Calculate  the  weight  of  two  moles  of  am- 
monia and  compare  it  to  the  sum  of  the 
weights  of  one  mole  of  nitrogen,  N2,  plus 
three  moles  of  hydrogen,  H2. 

5.  In  the  manufacture  of  nitric  acid,  HN03,  nitro- 
gen dioxide  reacts  with  water  to  form  HN03  and 
nitric  oxide,  NO: 

3N02  +  H20  — >■  2HN03  +  NO 

(a)  Verify  that  the  equation  conserves  oxygen 
atoms. 


(b)  How  many  molecules  of  nitrogen  dioxide  are 
required  to  form  25  molecules  of  nitric 
oxide  ? 

(c)  How  many  moles  of  nitric  oxide  are  formed 
from  0.60  mole  of  nitrogen  dioxide? 

6.  If  3  grams  of  substance  A  combine  with  4  grams 
of  substance  B  to  make  5  grams  of  substance  C 
and  some  D,  how  many  grams  of  D  would  you 
expect  ? 

7.  One  step  in  the  manufacture  of  sulfuric  acid  is 
to  burn  sulfur  (formula,  S8)  in  air  to  form  a 
colorless  gas  with  a  choking  odor.  The  name  of 
the  gas  is  sulfur  dioxide  and  it  has  the  molecular 
formula  S02.  On  the  basis  of  this  information : 

(a)  Write  the  balanced  equation  for  this  reac- 
tion. 

(b)  Interpret  the  equation  in  terms  of  molecules. 

(c)  Interpret  the  equation  in  terms  of  moles. 

(d)  Two  moles  of  sulfur,  S8,  would  produce  how 
many  moles  of  sulfur  dioxide,  SO,? 

8.  When  iron  rusts,  it  combines  with  oxygen  of  the 
air  to  form  iron  oxide,  Fe203.  Which  of  the  fol- 
lowing is  FALSE? 


(a)  The  equation  is 

302  +  4Fe 


2Fe203 


(b)  There  are  five  atoms  represented  by  the  for- 
mula, Fe203. 

(c)  Oxygen  gas  is  triatomic. 

(d)  The  mass  of  the  reactants  equals  the  mass 
of  the  products. 

(e)  Atoms  are  conserved. 

Balance  the  equations  for  each  of  the  following 
reactions.  Begin  on  the  basis  of  one  mole  of  the 
substance  underscored. 


(a)  Li    +  Clo 

(b)  Na  +  Cl2 

(c)  Na  +  F2 

(d)  Na  +  Br2 

(e)  Q2  +  Cl2 

(f)  o2  +  CJ2 


LiCl 

NaCl 

NaF 

NaBr 

ci2o 

CIO 


10. 


Show  that  your  answers  to  parts  (e)  and  (0  con- 
tain the  same  information. 

Balance  the  equations  for  each  of  the  following 
reactions  involving  oxygen.  Begin  on  the  basis 
of  one  mole  of  the  substance  underscored. 


46 


CHEMICAL  REACTIONS  I  CHAP.  3 


(a)  With  metallic  nickel: 

Ni  +  Q2  — >-  NiO 

(b)  With  metallic  nickel: 

Ni  +  02  — ►■  NiO 

(c)  With  metallic  lithium: 

Li  +  Q2  — >■  Li20 

(d)  With  the  rocket  fuel  hydrazine,  N2H4: 

NFL.  +  02  — ►-  N2  +  H-O 

(e)  With  acetylene,  QH2,  in  an  acetylene  torch 
flame: 

CM*  +  02  — »-  C02  +  H.O 

Answer.  C2H2  +  f 02  — >■  2C02  +  H20 

(f)  With  the  important  copper  ore,  chalcocite, 
Cu2S  (the  process  called  "roasting"  the  ore): 

Cu2S  +  02  — >-  Cu  .0  +  SO, 

(g)  With  the  important  iron  ore,  iron  pyrites, 
FeS2  (again,  "roasting"  the  ore): 

FeS2  +  02  — >■  Fe203  +  S02 

11.  (a)  Balance  the  equations  for  the  decomposition 
(to  elements)  of  ammonia,  NH3,  nitrogen 
trifluoride,  NF3,  and  nitrogen  trichloride, 
NC13.  Base  each  equation  upon  the  produc- 
tion of  one  mole  of  N2. 


NH3 

NF3 
NCI3 


1N2  +  H2 

1N2+F, 
1N2  +  Cl2 


(b)  Rewrite  the  equations  to  include  the  infor- 
mation that  the  decomposition  of  ammonia 
is  endothermic,  absorbing  22.08  kcal/mole 
N2,  the  decomposition  of  NF3  is  endother- 
mic, absorbing  54.4  kcal/mole  N2,  and  the 
decomposition  of  NC13  is  exothermic,  re- 
leasing 109.4  kcal/mole  N2. 

(c)  One  of  the  three  compounds  NH3,  NF3,  and 
NCI3  is  dangerously  explosive.  Which  would 
you  expect  to  be  the  explosive  substance? 
Why? 

12.  Graphite,  a  form  of  carbon,  C,  burns  in  air  to 
produce  the  colorless  gas,  carbon  dioxide.  On 
the  basis  of  this  information : 

(a)  Write  the  equation  for  the  reaction. 

(b)  If  one  mole  of  graphite  is  burned,  how  many 
moles  of  carbon  dioxide  are  produced?  What 
is  the  weight  in  grams  of  this  amount  of 
carbon  dioxide? 


(c)  If  two  moles  of  graphite  were  burned,  how 
many  moles  of  carbon  dioxide  would  be 
produced?  What  is  the  weight  in  grams? 

(d)  If  five  moles  of  graphite  were  burned  in  a 
vessel  containing  10  moles  of  oxygen  gas, 
what  is  the  maximum  number  of  moles  of 
carbon  dioxide  that  could  be  produced  ? 

13.  If  a  piece  of  sodium  metal  is  lowered  into  a  bottle 
of  chlorine  gas,  a  reaction  takes  place.  Table 
salt,  NaCl,  is  formed. 

(a)  Write  the  equation  for  the  reaction. 

(b)  How  many  moles  of  NaCl  could  be  formed 
from  one  mole  of  Na  ? 

(c)  How  many  moles  of  NaCl  could  be  formed 
from  2.30  grams  of  Na? 

14.  Methane,  the  main  constituent  of  natural  gas, 
has  the  formula  CH4.  Its  combustion  products 
are  carbon  dioxide  and  water. 

(a)  Write  the  equation  for  the  combustion  of 
methane.  Compare  your  answer  with  equa- 
tion (5),  p.  41. 

(b)  One  mole  of  methane  produces  how  many 
moles  of  water  vapor? 

(c)  One-eighth  mole  of  methane  would  produce 
how  many  moles  of  carbon  dioxide? 

(d)  How  many  moles  of  water  vapor  would  be 
produced  by  4.0  grams  of  methane? 

15.  If  potassium  chlorate,  KCIO3,  is  heated  gently, 
the  crystals  will  melt.  Further  heating  will  de- 
compose it  to  give  oxygen  gas  and  potassium 
chloride,  K.C1. 

(a)  Write  the  equation  for  the  decomposition. 

(b)  How  many  moles  of  K.CIO3  are  needed  to 
give  1.5  moles  of  oxygen  gas? 

(c)  How  many  moles  of  KC1  would  be  given  by 
i  mole  of  KCIO3? 

(d)  How  many  moles  of  oxygen  gas  would  be 
produced  by  122.6  grams  of  K.CIO3? 

16.  One  gallon  of  gasoline  can  be  considered  to  be 
about  25  moles  of  octane,  CsHl8. 

(a)  How  many  moles  of  oxygen  must  be  used  to 
burn  this  gasoline,  assuming  the  only  prod- 
ucts are  carbon  dioxide  and  water? 

(b)  How  many  moles  of  carbon  dioxide  are 
formed  ? 

(c)  How  much  does  this  carbon  dioxide  weigh? 
(Express  your  answer  in  kilograms.) 


QUESTIONS    AND    PROBLEMS 


47 


(d)  What  weight  of  carbon  dioxide  is  released 
into  the  atmosphere  when  your  automobile 
consumes  10  gallons  of  gasoline?  Express 
this  answer  in  pounds  (1  kg  =  2.2  pounds). 

17.  Iron  (Fe)  burns  in  air  to  form  a  black,  solid 
oxide  (Fe304). 

(a)  Write  the  equation  for  the  reaction. 

(b)  How  many  moles  of  oxygen  gas  are  needed 
to  burn  one  mole  of  iron  ? 

(c)  How  many  grams  of  oxygen  gas  is  that? 

(d)  Can  a  piece  of  iron  weighing  5.6  grams  burn 
completely  to  Fe304  in  a  vessel  containing 
0.05  mole  of  02? 

18.  Problem  5  relates  to  the  manufacture  of  nitric 
acid. 


(a)  According  to  the  equation  given  in  that 
problem,  how  many  grams  of  nitric  acid  are 
formed  from  one  mole  of  nitrogen  dioxide  ? 

(b)  How  many  more  grams  of  nitric  acid  could 
be  formed  if  the  nitric  oxide  formed  could 
be  completely  converted  into  nitric  acid  (as- 
sume one  mole  of  nitric  oxide  gives  one  mole 
of  nitric  acid)? 

19.  Hydrazine,  N2H4,  can  be  burned  with  oxygen  to 
provide  energy  for  rocket  propulsion.  The  energy 
released  is  150  kcal  per  mole  of  hydrazine 
burned. 

(a)  How  much  energy  is  released  if  10.0  kg  of 
hydrazine  fuel  are  burned? 

(b)  Compare  the  energy  that  would  be  released 
if  the  same  weight  of  hydrogen,  10.0  kg,  were 
burned  as  a  fuel  instead  (see  Section  3-1.1). 


Gilbert  Newton  Lewis,  one  of  the  greatest  American  chem- 
ists of  the  twentieth  century,  began  his  career  teaching  high 
school  chemistry.  Born  near  Boston  and  reared  in  Ne- 
braska, young  Lewis  returned  to  the  East  to  study  and 
graduated  from  Harvard  University.  After  a  year  teaching 
high  school,  he  returned  to  Harvard  and  received  the  Ph.D. 
in  1899.  There  followed  a  year  at  universities  in  Germany, 
another  as  Superintendent  of  Weights  and  Measures  in  the 
Philippine  Islands.  Then,  in  seven  years  at  the  Massachu- 
setts Institute  of  Technology  he  rose  to  the  rank  of  Profes- 
sor. Finally,  in  1912  he  accepted  the  position  of  Chairman 
of  what  was  then  a  little  known  chemistry  department  far 
from  the  recognized  scientific  centers.  He  moved  to  the 
University  of  California  and  spent  the  remainder  of  his 
career  at  Berkeley,  building  one  of  the  most  powerful 
chemistry  departments  in  the  world. 

Lewis  devoted  most  of  his  career  to  the  understanding  of 
the  structures  of  molecules  and  of  thermodynamics,  the 
energy  relations  in  chemical  changes.  His  thinking  was  far 
ahead  of  his  lime  and  his  theories  have  had  profound  in- 
fluence on  chemistry.  His  understanding  of  chemical  bond- 
ing has  strongly  influenced  modern  thinking  on  this  subject. 
Lewis  was  one  of  the  first  to  recognize  that  energy  effects 
provide  a  basis  for  predicting  what  chemical  reactions  can 
occur.  Thus  he  awakened  chemists  to  the  crucial  importance 
of  thermodynamics.  His  book  on  this  subject,  published  in 
1923,  became  a  classic  of  the  chemical  literature.  He  pub- 
lished over  150  research  publications  on  topics  extending 
from  the  phases  of  sulfur  to  quantum  mechanics. 

G.  N.  Lewis  enjoyed  chemistry.  Throughout  his  distin- 
guished career  he  remained  active  in  the  laboratory  and 
never  tired  of  the  thrill  of  discovery.  He  favored  simple 
and  direct  experiments — many  of  his  important  discoveries 
were  performed  with  a  few  test  tubes  and  simple  chemicals. 
His  enthusiasm  and  burning  interest  in  chemistry  were 
contagious — many  of  his  students  became  great  scientists. 
Lewis  virtually  eliminated  graduate  courses,  relying  in- 
stead upon  the  open  debate  of  research  seminars.  He  en- 
couraged his  colleagues  to  think  critically,  to  challenge  his 
ideas,  and  to  welcome  challenge  of  their  own. 

G.  N.  Lewis  died  March  23,  1946,  in  the  laboratory  he 
loved,  surrounded  by  the  beakers  and  books  that  were  the 
tools  of  his  trade.  He  is  remembered  and  respected  by 
chemists  the  world  over. 


CHAPTER 


4 


The  Gas  Phase: 
Kinetic  Theory 


•  •  •  it  is  my  intention  to  make  known  some  new  properties  in  gases, 
the  effects  of  which  are  regular,  by  showing  that  these  substances  com- 
bine amongst  themselves  in  very  simple  proportions   ■    •    • 

JOSEPH    L.    GAY-LUSSAC,    1808 


We  have  already  seen  that  the  behavior  of  gases 
is  important  to  a  chemist.  The  pressure-volume 
behavior  leads  to  the  particle  model  of  a  gas. 
Differences  among  gases  (in  properties  such  as 
color,  odor,  and  solubility)  show  that  the  par- 
ticles of  one  gas  differ  from  the  particles  of  an- 
other gas.  In  chemical  reactions,  the  simple 
combining  volume  relationships  support  Avo- 
gadro's  Hypothesis  and,  hence,  give  us  a  way  to 
measure  molecular  weights. 


Thus  we  see  that  the  properties  of  gases  pro- 
vide a  substantial  basis  for  developing  the  atomic 
theory.  The  gaseous  state  is,  in  many  ways,  the 
simplest  state  of  matter  for  us  to  understand. 
The  regularities  we  discover  are  susceptible  to 
detailed  mathematical  interpretation.  We  shall 
examine  these  regularities  in  this  chapter.  We 
shall  find  that  their  interpretation,  called  the 
kinetic  theory,  provides  an  understanding  of  the 
meaning  of  temperature  on  the  molecular  level. 


4-1  THE  VOLUME  OCCUPIED  BY  ONE  MOLE  OF  GAS 


To  a  chemist,  one  of  the  most  important  regu- 
larities displayed  by  a  gas  relates  to  the  volume 
occupied  by  one  mole  of  a  gas.  We  shall  begin 
investigating  this  subject  by  comparing  the  sizes 
of  gaseous  particles  with  the^average  spacing 
between  them  under  normal  conditions  of  tem- 
perature and  pressure.  The  comparison  can  be 
based  upon  the  volumes  occupied  by  a  mole  of 


nitrogen  first  as  a  solid,  then  as  a  liquid,  and 
finally  as  a  gas. 

4-1.1    The  Volume  Occupied  by  a  Mole 
off  Nitrogen,  N, 

The  molecular  formula  of  nitrogen  is  N2;  the 
nitrogen  molecule  is  diatomic.  One  mole  of  N2 

49 


50 


THE   GAS   PHASB:    KINETIC   THEORY   |   CHAP.    4 


molecules  contains,  then,  two  moles  of  nitrogen 
atoms.  The  weight  of  one  mole  of  nitrogen  mole- 
cules is  28.0  grams. 

At  a  sufficiently  low  temperature,  below 
-210°C,  nitrogen  is  a  solid  with  a  density  of 
1  03  grams  per  milliliter.  The  volume  occupied 
by  one  mole  of  the  solid,  called  the  molar 
volume,  is 
Molar  volume,  solid.    2-ff^f  =  *"  ■"*** 

Now  if  we  warm  the  solid  to  -210°C,  the 
soUd  melts  to  form  liquid  nitrogen.  The  density 
of  this  liquid  is  0.81  grams  per  milliliter.  Now  the 
volume  of  a  mole  is 
Molarvolume,/^:    2-|ff^  =  34.6 ml/mole 

If  we  raise  the  temperature  still  further,  the 
liquid  vaporizes  to  form  nitrogen  gas,  taking 
whatever  density  is  necessary  to  fill  the  container. 
The  density  now  depends  upon  the  volume  of  the 
container  and  the  temperature.  For  the  sake  of 
comparison,  suppose  the  gaseous  nitrogen  is 
placed  in  that  volume  that  gives  a  pressure  of 
one  atmosphere  when  the  container  is  placed  m 
an  ice  bath  at  0°C.  Then  the  density  is  found  to 
be  only  0.00125  gram  per  milliliter.  This  means 
that  the  volume  required  for  one  mole  of  gas  is 
Molar  volume,  gas  at  0°C,  1  atm: 

28.0  g/mole  =  22  4  X  10s  ml/mole 
0.00125  g/ml 

=  22.4  liters/mole 

The  volume  of  this  gas  is  almost  1000  times  as 
great  as  the  volume  of  the  same  weight  of  solid. 
Experiments  with  other  gases  lead  to  similar 
results.  If  the  size  of  a  single  molecule  is  assumed 
to  be  the  same  in  the  solid  and  gas,  then  the 
molecules  must  have  separated  from  each  other 
in  the  gas.  The  free  space  between  gaseous  mole- 
cules is  on  the  order  of  1000  times  the  volume 
a  molecule  occupies  in  the  solid. 


EXERCISE  4-2 

(a)  Calculate  the  volume  (in  milliliters)  occupied 
by  one  nitrogen  molecule  in  the  solid  phase. 

(b)  Recognizing  that  one  milliliter  is  1.00  cubic 
centimeter,  estimate  the  size  (in  centimeters) 
of  a  cube  that  has  the  volume  calculated 
in  part  (a).  Use  one  significant  figure.  Now 
express  your  answer  in  Angstroms  (1  A  = 
10"8  cm). 


EXERCISE  4-1 

How  many  molecules  of  nitrogen  are  present  in 
one  liter  of  the  gas  at  0°C  and  one  atmosphere 
pressure? 


4-1.2    A  Comparison  of  Molar  Volumes  of  Gases 

The  volume  calculated  above,  22.4  liters,  we 
have  seen  before.  In  Table  l-II  the  pressure- 
volume  product  of  32.0  grams  of  oxygen,  02,  was 
found  to  be  22.4  at  0°C.  (Notice  that  32.0  grams 
of  Oa  is  the  weight  of  one  mole  of  oxygen.)  So 
we  can  use  this  relation, 

liters  X  atmospheres  (ftt  0oQ 
r  A  r  mole 

to  solve  for  the  volume  of  a  mole  of  O,  at  one 
atmosphere  pressure: 

1  atm  X  V  =  22.4  liters  X  atm/mole 
liters  X  atm/mole 
V  "  2ZA        1  atm 
=  22.4  liters/mole 

This  is  the  same  volume  as  that  just  calculated 
for  a  mole  of  nitrogen  at  0°C  and  one  atmos- 
phere pressure  (in  Section  4-1.1).  Furthermore, 
it  is  the  same  volume  occupied  by  17.0  grams  of 
ammonia  at  0°C  and  one  atmosphere  pressure 
(See  Table  2-III,  p.  19).  Since  the  molecular 
formula  of  ammonia  is  NH3,  its  molecular  weight 
is  (14.0  +  3  X  1.0)  =  17.0  grams.  Thus  one  mole 
of  ammonia,  17.0  grams,  also  occupies  22.4  liters 
at  0°C  and  one  atmosphere  pressure.  Experi- 
ments on  many  other  gases  are  in  agreement  and 
lead  to  the  generalization: 

A  mole  of  gas  occupies  22.4  liters  at  PC  and 
one  atmosphere  pressure.  v  > 

What  happens  to  a  gas  as  the  temperature  is 
changed?  An  experiment  provides  the  answer. 
Table  4-1  shows  some  pressure-volume  measure- 


SEC.  4-1  I  THE  VOLUME  OCCUPIED  BY  ONE  MOLE  OF  GAS 


51 


ments  for  ammonia  gas  at  25°C  (approximately 
room  temperature).  Although  the  data  shown 
contain  some  experimental  uncertainty,  we  find 
again  the  regularity,  PV  =  a  constant. 


Table  4-1 

PRESSURE    AND    VOLUME    OF    17.0    GRAMS 
OF    AMMONIA    GAS,    NHi    t  -  25°C 


PRESSURE 

VOLUME 

(atmospheres) 

(liters) 

PX  V 

0.200 

123 

24.6 

0.400 

60.0 

24.0 

0.600 

43.0 

25.8 

0.800 

29.3 

23.4 

1.00 

25.7 

25.7 

1.50 

15.9 

23.9 

2.00 

12.1 

24.2 
Average       24.5  ±0.7 

This  time,  however,  the  pressure-volume  prod- 
uct for  a  mole  of  ammonia  is  24.5  ±  0.7.  We 


Fig.  4-1.  A  mole  of  gas  occupies  22.4  liters  at  0'C, 
1  atmosphere. 

A  mole  of  gas  occupies  24.5  liters  at  25' C, 
1  atmosphere. 


can  compare  this  with  our  earlier  result: 

For  one  mole  of  ammonia  at  0°C, 

PXV  =  22.4  liter-atm  (2) 

For  one  mole  of  ammonia  at  25°C, 

p  x  V  =  24.5  liter-atm  (5) 

From  (3)  we  see  that  the  molar  volume  of 
ammonia  at  25°C  and  one  atmosphere  pressure 
is  24.5  liters,  whereas  it  is  22.4  liters  at  0°C.  The 
molar  volume  of  ammonia  depends  upon  the 
temperature.  This  result  is  no  surprise  —a  sample 
of  gas  expands  when  heated  at  constant  pressure. 
So  when  we  compare  the  molar  volumes  of  dif- 
ferent gases,  they  should  be  at  the  same  tem- 
perature (and,  by  the  same  sort  of  argument,  at 
the  same  pressure). 

Consider  the  following  experiment.  The  air  is 
removed  from  a  one-liter  flask  and  it  is  weighed 


Table  4-II     THE  VOLUME  °  F  A  mole  of  gas  at  25  c  and  one  atmosphere 

PRESSU  RE 


GAS 


WT.  OF  WT.  OF  WT.  OF  MOLECULAR  VOLUME 

FLASK  EMPTY         FLASK  +  GAS  1  LITER  OF  GAS  WEIGHT  (liter/mole) 

Wi(g)  W»(g)  Wx  -  W ,  (g/liter)      MW(g/mole)      MW/(Wt-Wx) 


oxygen,  Os 

157.35 

158.66 

1.31 

32.0 

24.5 

nitrogen,  Nj 

157.35 

158.50 

1.15 

28.0 

24.3 

carbon 

monoxide,  CO 

157.35 

158.50 

1.15 

28.0 

24.4 

carbon 

dioxide,  COj 

157.35 

159.16 

1.81 

44.0 

24.3 

52 


THE    GAS    PHASE:    KINETIC    THEORY    I    CHAP.    4 


empty.  Then  the  flask  is  weighed  again  filled  with 
a  gas  at  one  atmosphere  pressure  and  at  25°C. 
The  difference  in  weight  is  the  weight  of  one  liter 
of  the  gas.  From  this,  we  can  calculate  the  vol- 
ume of  a  mole  of  that  gas.  Table  4-II  shows  the 
results.  We  find  that  all  the  gases  have  about  the 
same  molar  volume  at  25°C  and  one  atmosphere. 
Whether  the  gas  is  02,  N2,  CO,  or  C02,  the  same 
volume,  24.5  ±  0.2  liters,  contains  6.02  X  1023 
molecules  (at  25°C,  one  atmosphere).  Whether 
the  gas  is  N2  or  CO,  a  volume  of  22.4  ±  0.1  liter 
contains  6.02  X  1023  molecules  at  0°C  and  one 
atmosphere  pressure. 


4-1.3    Avogadro's  Hypothesis 

When  different  gases  are  compared  at  the  same 
temperature  and  pressure,  they  have  the  same 
volume  per  mole.  This  is  true  at  0°C  and  one 
atmosphere  but,  more  important,  it  is  true  at 
other  temperatures  and  pressures  as  well. 

When  the  two  gases  ammonia,  NH3,  and  hy- 
drogen chloride,  HC1,  react,  one  liter  of  ammo- 
nia reacts  with  one  liter  of  hydrogen  chloride  if 
the  two  volumes  are  measured  at  the  same  tem- 
perature and  pressure.  This  simple  one-to-one 
volume  ratio  is  observed  at  0°C  and  one  atmos- 
phere pressure  but,  more  important,  this  simple 
ratio  is  observed  at  other  temperatures  and 
pressures  as  well. 

These  results  and  many  similar  ones  led  Avo- 
gadro  to  propose  his  famous  hypothesis,  as  dis- 
cussed in  Section  2-2.3.  The  hypothesis  states 
that  equal  volumes  of  gases  contain  equal  numbers 
of  molecules  (at  the  same  temperature  and  pres- 
sure). Therefore,  the  molecular  weight  of  a  gas 


can  be  determined  by  comparing  the  weight 
of  a  known  volume  of  the  gas  with  the  weight 
of  the  same  volume  of  another  gas  of  known 
molecular  weight.  It  does  not  matter  what  t  and 
P  are  as  long  as  they  are  the  same  for  the 
two  gases. 

"Avogadro's  Hypothesis"  is  often  called 
"Avogadro's  Law"  because  it  has  such  wide  ap- 
plicability. It  is  one  of  the  important  generaliza- 
tions of  chemistry.  It  is  important,  not  because 
it  is  exact  but  because  it  applies  to  all  gases, 
regardless  of  whether  their  molecules  are  large 
or  small.  The  molecules  of  different  gases  actu- 
ally have  different  sizes  and  slightly  different 
attractions  for  one  another.  As  a  result,  different 
gases  do  not  have  exactly  the  same  number  of 
molecules  in  a  given  volume.  Such  variations  are 
small  (usually  less  than  1  %)  and  do  not  impair 
the  usefulness  of  Avogadro's  Hypothesis  as  a 
method  for  determining  the  molecular  weight  of 
a  gas. 

It  is  an  interesting  commentary  on  the  progress 
of  science  that  this  important  regularity,  now 
often  called  a  "Law,"  was  not  generally  accepted 
for  half  a  century  after  it  was  proposed.  Though 
Avogadro  published  his  idea  in  181 1,  its  validity 
was  not  widely  recognized  until  the  proposal  was 
reintroduced  at  an  international  conference  of 
chemists  at  Karlsruhe,  Germany,  in  1858.  Today, 
we  find  it  easy  to  "discover"  or  "confirm"  Avo- 
gadro's Hypothesis  because  we  can  draw  upon 
a  wealth  of  accumulated  quantitative  weight  and 
volume  relations  to  develop  a  tightly  knit  pattern 
of  self-consistency.  In  contrast,  even  atomic 
weights  were  in  doubt  early  in  the  nineteenth 
century  and  quantitative  methods  were  rela- 
tively crude. 


4-2    THE  KINETIC  THEORY 


Avogadro's  Hypothesis  provides  a  method  for 
identifying  the  molecules  present  in  a  gas.  Also, 
it  explains  why  the  volumes  of  gases  that  react 
with  each  other  are  in  the  same  simple  ratio  as 
are  the  moles  in  the  balanced  equation.  The 
importance  of  these  results  makes  the  explana- 


tion of  the  properties  of  gases  important  to  a 
chemist. 

We  have  already  observed  that  there  are  many 
and  close  similarities  between  a  gas  and  a  col- 
lection of  particles  in  endless  motion.  It  is  essen- 
tial in  the  particle  model  that  each  particle 


SEC.    4-2    I    THE    KINETIC    THEORY 


53 


possesses  energy  of  motion,  called  kinetic 
energy.  Hence,  the  mathematical  expression 
that  describes  this  model  is  called  the  kinetic 
theory  of  gases.  According  to  this  theory,  the 
molecules  of  a  gas  are  in  rapid  motion.  They 
travel  in  straight  lines  until  they  meet  other  mole- 
cules of  the  gas  or  the  atoms  in  the  walls  of  the 
container.  They  are  then  deflected  and  scattered. 
The  net  result  is  a  helter-skelter  movement  of 
molecules  traveling  in  all  directions  and  at  dif- 
ferent speeds. 

At  room  temperature,  the  average  speed  of  a 
nitrogen  molecule  is  found  to  be  about  one- 
quarter  mile  per  second.  In  one  second,  however, 
the  nitrogen  molecule  has  collided  with  many 
other  molecules,  so  its  motion  follows  a  zig-zag 
path.  Although  the  average  distance  between 
molecules  is  small,  the  molecule  passes  by  many 
other  molecules  without  hitting  them,  so  the  dis- 
tance it  travels  between  collisions  is  about  fifteen 
times  the  average  distance  between  the  molecules 
(at  room  pressure  and  temperature). 

4-2.1    Gas  Pressure 

Pressure  is  an  important  quantity  in  a  discussion 
of  gas  behavior.  The  applicability  of  the  kinetic 
theory  to  an  understanding  of  gas  pressure  is, 
then,  an  important  success  (see  Section  2-1.1). 
We  shall  investigate  this  success  in  more  detail, 
but  first  we  should  investigate  how  pressure  is 
measured. 

MEASURING    THE    PRESSURE    OF    A    GAS 

A  gas  exerts  pressure  equally  on  all  the  walls  of 
its  container.  The  standard  method  of  measuring 
this  pressure  is  by  measuring  the  height  of  a 
mercury  column  supported  by  the  gas.  An  in- 
strument for  measuring  the  pressure  of  the  air  is 
illustrated  in  Figure  4-2A;  it  is  called  a  barome- 
ter. We  can  make  a  barometer  by  filling  a  long 
tube  (closed  at  one  end)  with  mercury  and  in- 
verting it  in  a  dish  of  mercury.  Mercury  will  flow 
from  the  tube  into  the  dish  until  the  column  of 
mercury  exerts  a  downward  pressure  which  is 
exactly  balanced  by  the  pressure  of  the  air.  In 
the  illustration,  the  pressure  of  the  air  is  ex- 
pressed as  "755  millimeters  of  mercury"  (written 


mm  Hg  or  mm).  This  is  the  height  of  the  mercury 
column.  (Notice  that  only  mercury  vapor  is  pres- 
ent to  exert  pressure  in  the  space  at  the  top  of 
the  column.  At  room  temperature,  this  vapor 
pressure  is  negligible,  about  10-3  mm.) 

The  pressure  of  a  gas  sample  can  be  measured 
in  a  device  similar  to  a  barometer,  called  a 
manometer.  Figures  4-2B  and  4-2C  show  two 
types.  Figure  4-2B  shows  a  closed-end  manome- 
ter. Here  the  downward  pressure  exerted  by  the 
column  of  mercury  is  balanced  by  the  pressure 
of  the  gas  sample  placed  in  the  flask.  The  gas 
pressure  is,  in  the  example  shown,  105  mm.  As  in 
the  barometer,  only  mercury  vapor  is  present  in 
the  right-hand  tube. 

The  apparatus  shown  in  Figure  4-2C  differs  in 
that  the  right-hand  tube  is  open.  In  this  type  of 
manometer,  atmospheric  pressure  is  exerted  on 
the  right-hand  mercury  column.  Hence  the  pres- 
sure in  the  flask  plus  the  height  of  the  mercury 
column  equals  atmospheric  pressure.  In  the  ex- 
ample shown,  the  pressure  is  755  —  650  =  105 
mm,  the  same  as  pictured  in  the  closed-end 
manometer,  Figure  4-2B. 

STP 

Two  conditions  that  are  often  important  in 
chemical  experiments  are  temperature  and  pres- 
sure. Consequently,  chemists  usually  control  and 
measure  these  conditions  during  experiments.  In 
addition,  it  is  useful  to  refer  many  experimental 
results  to  a  standard  and  generally  accepted  set 
of  temperature  and  pressure  conditions.  This 
facilitates  comparison  of  results  of  different 
types  and  from  different  laboratories. 

The  temperature  0°C  is  readily  obtained  and 
maintained  with  an  ice  water  bath.  The  tem- 
perature is  one  at  which  thermometers  are  cali- 
brated; this  aids  measurement.  A  temperature 
that  is  easy  to  maintain  and  easy  to  measure 
makes  a  good  standard  temperature. 

Air  pressure  varies  somewhat  from  day  to  day 
and  from  place  to  place.  Nevertheless,  air  pres- 
sure is  always  reasonably  near  760  mm  Hg,  so 
atmospheric  pressure  furnishes  a  convenient, 
though  approximate,  reference  pressure.  How- 
ever, it  is  not  sufficiently  constant  for  many 
purposes.    So,    by    international    agreement,    a 


54 


THE    GAS    PHASE:    KINETIC    THEORY    I    CHAP.    4 


Air  pressure, 
755  mm 


fli 


Valve 


■  Vacuum. 


Gas  inle-t 


Valve 


755 


m 


Fig.  4-2.  Pressure  measurement.  A.  Barometer:  pres- 
sure =  755  mm.  B.  Closed-end  manometer: 
pressure  =  105  mm.  C.  Open-end  manom- 
eter: pressure  =  755  —  650  =  105  mm. 


standard  pressure  for  gases  is  represented  by  a 
height  of  760  mm  of  mercury.  This  standard  pres- 
sure is  often  expressed  merely  as  one  atmos- 
phere (1  atm). 

Thus  chemists  have  accepted  0°C  and  one 
atmosphere  as  convenient  standard  conditions. 
These  conditions,  0°C  and  760  mm  pressure, 
are  called  standard  temperature  and  pres- 
sure and  are  abbreviated  STP. 

The  standard  pressure  is  defined  in  terms  of  a  pressure 
reading  on  a  standard  barometer.  A  standard  barometer 


7.55-650  =  105  mm 

takes  account  of  the  fact  that  the  gravitational  attraction 
of  the  earth  on  the  mercury  varies  slightly  from  place  to 
place,  and  the  fact  that  mercury  expands  and  becomes  less 
dense  when  it  is  heated.  Thus,  the  mercury  column  of  a 
barometer  is  several  millimeters  longer  at  20°C  than  at 
0°C.  In  the  standard  barometer,  the  mercury  is  at  0°C. 
You  can  find  in  published  tables  how  much  to  subtract 
from  (or  add  to)  your  barometer  reading  to  obtain  the 
same  pressure  value  a  standard  barometer  would  give. 
The  correction  is  seldom  more  than  1  or  2  mm,  and  is 
often  negligible  compared  to  other  possible  errors.  Unless 
your  other  experimental  measurements  are  rather  precise, 
you  need  not  correct  your  readings  in  this  way. 


THE   CAUSE   OF   GAS    PRESSURE 

In  Chapter  1  we  explained  how  gases  exert  pres- 
sure in  terms  of  collisions  of  particles  with  the 


SEC.    4-2    I    THE    KINETIC    THEORY 


55 


container  walls:  this  model  of  gas  pressure  is 
part  of  the  kinetic  theory.  Each  time  a  gas  mole- 
cule strikes  a  wall,  or  a  mercury  surface,  it  exerts 
a  small  push  or  force,  just  as  a  ball  thrown  at  a 
wall  exerts  a  force  upon  it.  The  force  per  unit 
area,  called  the  pressure,  depends  directly  upon 
the  number  of  molecules  that  strike  the  unit  area 
of  the  surface.  Twice  as  many  molecules  in  a 
given  volume  result  in  twice  as  many  collisions 
per  unit  area,  hence  twice  the  original  pressure. 
Thus  we  explain  why  the  pressure  goes  up  as  we 
pump  air  into  an  automobile  tire.  If  the  volume 
and  temperature  of  the  tire  remain  unchanged, 
the  pressure  goes  up  in  direct  proportion  to  the 
number  of  moles  of  air  pumped  in. 


EXERCISE  4-3 

A  container  of  fixed  volume  contains  two  moles 
of  gas  at  room  temperature.  The  pressure  in  the 
container  is  four  atmospheres.  Three  more  moles 
of  gas  are  added  to  the  container  at  the  same 
temperature.  Use  the  result  just  stated  to  show 
that  the  pressure  is  now  10  atmospheres. 


(93  mm).  The  second  bulb  contains  0.001 1  mole 
of  water  vapor.  The  pressure  in  this  bulb  is  20 
mm  Hg.  The  third  bulb  contains  0.0050  mole  of 
air  and  also  0.0011  mole  of  water  vapor.  The 
third  manometer  shows  that  the  pressure  in  the 
last  bulb  is  113  mm  Hg. 

This  experiment  shows  that  the  pressure  ex- 
erted by  the  mixture  of  gases  is  just  the  sum  of 
the  pressure  the  air  exerts  when  alone  in  the 
flask  and  the  pressure  the  water  vapor  exerts 
when  alone  in  the  flask: 


1 1 3  mm  =  93  mm  +  20  mm 


(4) 


The  total  pressure  can  be  regarded  as  a  sum  of 
the  parts  furnished  by  the  individual  pressures 
exerted  by  each  of  the  components  of  the  gas 
mixtures.  The  pressure  exerted  by  each  of  the 
gases  in  a  gas  mixture  is  called  the  partial 
pressure  of  that  gas.  The  partial  pressure  is  the 
pressure  that  the  gas  would  exert  if  it  were  alone 
in  the  container.  In  the  example  of  Figure  4-3, 
the  total  pressure  in  the  third  bulb  is  113  mm. 
The  partial  pressure  of  water  vapor  in  this  bulb 
is  20  mm  and  the  partial  pressure  of  air  is  93  mm. 


4-2.2    Partial  Pressure 

Figure  4-3  shows  three  one-liter  bulbs  at  25°C. 
The  first  bulb  contains  0.0050  mole  of  air.  The 
manometer  shows  that  the  pressure  is  93  mm  Hg 


Fie.  4-3.  Pressure  of  a  mixture  of  gases. 


EXERCISE  4-4 

Assume  that  0.0050  mole  of  air  contains  0.0040 
mole  of  nitrogen,  N2,  and  0.0010  mole  of  oxygen, 
02.  What  is  the  partial  pressure  of  oxygen  in  the 
first  bulb  in  Figure  4-3?  What  is  the  partial  pres- 
sure of  oxygen  in  the  third  bulb?  Use  three  sig- 
nificant figures. 


93  mm 


.0050  mole  air- 
in  one  liter 

ir=  Z5°C 


20  mm 


113  mm 


.0011  mole  yvater  vapor 
in  one    liter 

t=ZS°C 


.0050  mote  air 
-r  .0011  mole  water  vapor 
in  one  liter 

t^  25  °C 


56 


THE    GAS    PHASE:    KINETIC    THEORY   |    CHAP.    4 


The  pressure  behavior  shown  in  Figure  4-3  is 
readily  explained  in  terms  of  the  kinetic  theory 
of  gases.  There  is  so  much  space  between  the 
molecules  that  each  behaves  independently,  con- 
tributing its  share  to  the  total  pressure  through 
its  occasional  collisions  with  the  container  walls. 
The  water  molecules  in  the  third  bulb  are  seldom 
close  to  each  other  or  to  molecules  provided  by 
the  air.  Consequently,  they  contribute  to  the 
pressure  exactly  the  same  amount  they  do  in  the 
second  bulb — the  pressure  they  would  exert  if 
the  air  were  not  present.  The  0.0011  mole  of 
water  vapor  contributes  20  mm  of  pressure 
whether  the  air  is  there  or  not.  The  0.0050  mole 
of  air  contributes  93  mm  of  pressure  whether  the 
water  vapor  is  there  or  not.  Together,  the  two 
partial  pressures,  20  mm  and  93  mm,  determine 
the  measured  total  pressure. 

4-2.3    Temperature  and  Kinetic  Energy 

If  the  kinetic  theory  is  applicable  to  gases,  we 
should  expect  pressure  to  be  affected  by  other 
factors  than  the  number  of  moles  per  unit  vol- 
ume. For  example,  the  mass  of  the  molecules  and 
their  velocities  should  be  important,  as  well. 
After  all,  a  baseball  exerts  more  "push"  on  a 
catcher's  mitt  than  would  a  ping-pong  ball 
thrown  with  the  same  velocity.  Also,  a  baseball 
exerts  more  "push"  on  the  mitt  if  a  "fast  ball" 
is  thrown  rather  than  a  "slow  ball."  To  see  how 
the  mass  of  the  molecules  and  their  velocities  are 
dealt  with  in  the  kinetic  theory,  we  must  consider 
temperature. 

To  measure  the  temperature  of  a  gas  we  im- 
merse some  kind  of  thermometer  in  it.  If  the 
thermometer  is  colder  than  the  system,  heat  flows 
into  the  thermometer  until  the  gas  and  the  ther- 
mometer are  at  the  same  temperature.  Then  we 
read  the  thermometer  to  get  a  numerical  value 
for  the  temperature.  If  the  thermometer  were 
hotter  than  the  gas,  heat  would  flow  from  the 
thermometer.  When  there  is  no  net  flow  of  heat, 
the  thermometer  is  said  to  be  in  thermal 
equilibrium  with  the  gas. 

There  are  many  kinds  of  thermometers.  Any 
substance  can  be  fashioned  into  a  thermometer 
if  it  has  a  readily  measured  property  that  is  sen- 


sitive to  a  change  in  temperature.  The  familiar 
mercury  thermometer  depends  upon  the  expan- 
sion of  the  liquid  as  temperature  is  raised.  Solids 
and  gases  also  change  volume  with  temperature 
change.  Hence  either  can  be  (and  both  are)  used 
as  a  basis  for  a  thermometer.  A  gas  held  at  con- 
stant volume  also  responds  to  a  change  in  tem- 
perature, the  pressure  rising  with  rising  tem- 
perature. This  is  the  more  common  way  in  which 
a  gas  is  used  in  a  thermometer:  the  volume  is 
fixed  and  the  pressure  varies  with  temperature. 

So  let  us  measure  the  temperature  of  a  sample 
of  gas  A  by  placing  it  in  thermal  contact  with  a 
sample  of  gas  B  (our  thermometer).  There  will 
be  heat  flow  between  the  two  gas  samples  if  they 
are  initially  at  different  temperatures.  Energy  is 
transferred  from  the  hotter  gas  to  the  cooler  gas. 
When  heat  flow  ceases,  the  gases  have  reached 
thermal  equilibrium.  Then  the  gases  have  the 
same  temperature. 

We  can  visualize  what  is  going  on  with  the  aid 
of  the  kinetic  theory  of  gases.  Suppose  sample  A 
is  initially  at  a  high  temperature  relative  to  the 
thermometer  gas  B.  We  interpret  this  to  mean 
that  the  molecules  in  gas  A  have  more  energy  of 
motion  than  those  of  gas  B — the  molecules  of  gas 
A  have  higher  kinetic  energies  (on  the  average). 
When  the  samples  are  brought  into  thermal  con- 
tact, collisions  permit  the  rapidly  moving  A 
molecules  to  transfer  kinetic  energy  through  the 
thermal  connection  to  the  slowly  moving  B  mole- 
cules. This  transfer  of  kinetic  energy  from  gas  A 
to  gas  B  is  the  process  that  raises  the  temperature 
of  gas  B  and  lowers  the  temperature  of  gas  A. 
When  the  thermal  contact  between  molecules  of 
A  and  B  no  longer  results  in  a  net  transfer  of 
kinetic  energy  from  one  gas  to  the  other,  then 
gases  A  and  B  are  in  thermal  equilibrium:  they 
have  the  same  temperatures. 

Thus,  we  picture  heat  flow  between  two  gas 
samples  as  a  transfer  of  kinetic  energy.  The  proc- 
ess continues  until  the  molecules  of  both  gases 
have  the  same  average  kinetic  energy.  Then  the 
gases  are  at  the  same  temperature.  This  is  a 
basic  premise  of  the  kinetic  theory:  When  gases 
are  at  the  same  temperature,  the  molecules  of  the 
gases  have  the  same  kinetic  energy  (on  the  aver- 
age). 


SEC.    4-2    I    THE    KINETIC    THEORY 


57 


4-2.4    Absolute  Temperature 

The  quantitative  effects  of  temperature  on  gases 
were  first  studied  by  Jacques  Charles,  a  French 
scientist,  in  1787.  He  found  that  all  gases  expand 
by  the  same  fraction  of  their  original  volumes 
when  they  are  heated  over  the  same  temperature 
range.  (In  these  experiments  the  pressure  re- 
mained the  same.)  A  simple  experiment  shows 
the  relations.  Into  a  small-bore  glass  tube,  one- 
half  meter  in  length  and  closed  at  one  end,  we 
place  a  drop  of  mercury.  This  mercury  falls  and 
finally  traps  a  sample  of  air  in  the  bottom  of  the 
tube.  (See  Figure  4-4.)  Since  the  tube  has  a  uni- 
form bore,  .we  can  use  the  length  of  the  air 
sample  as  a  measure  of  its  volume.  The  mercury 
plug  moves  up  or  down  and  maintains  a  con- 
stant pressure. 


Fig.  4-4.  Apparatus  for  demonstrating  the  effect  of 
temperature  on  the  volume  of  a  gas. 


Table  4-/II. 

CHANGE  OF  VOLUME  OF  A  GAS 
WITH  CHANGE  IN  TEMPERATURE 


RELATIVE  VOLUME 

TEMPERATURE 

(as  measured  by 

(°C) 

length  of  sample) 

200 

1.73 

100 

1.37 

50 

1.18 

0 

1.00 

When  we  plot  these  results  with  relative  vol- 
umes on  the  ordinate  (vertical  axis)  and  tem- 
peratures on  the  abscissa  (horizontal  axis),  we 
obtain  the  graph  shown  in  Figure  4-5.  The 
straight  line  passes  through  the  experimental 
points.  When  extrapolated  upward,  it  shows  that 
the  volume  at  273°C  is  double  that  at  0°C. 
Extrapolated  downward,  the  line  shows  that  the 


2.0 


1.0 


-273  "C 


1 
1 

-200 

-100         0       +100 

+200 

+300  °C 

0 

73 

173      273       373 
Te  mpe  ra  tttre 

■473 

573  °K 

Fig.  4-5.  An    absolute    temperature    scale    from    the 
change  of  volume  of  a  gas  with  temperature. 


We  may  place  the  tube  in  ice  water  (0°C)  and 
measure  the  relative  volume  of  the  air  sample. 
If  the  tube  is  immersed  in  water  boiling  at  one 
atmosphere  pressure  (100°C),  the  relative  vol- 
ume has  a  higher  value.  From  these  data  and 
from  similar  measurements  at  other  tempera- 
tures, we  collect  data  such  as  those  in  Table  4-III. 


volume  would  become  zero  at  —  273°C.  The  vol- 
ume change  per  degree  centigrade  is  yfg  of  the 
volume  at  0°C.  Actually,  all  gases  liquefy  before 
their  temperature  reaches  —  273°C. 

If  gases  are  heated  or  cooled  at  constant  vol- 
ume, the  pressure  changes,  also  at  the  rate  of  ^ 
of  its  value  at  0°C.  Then  the  pressure  of  a  gas 


58 


THB   GAS   PHASE:    KINETIC   THEORY   I   CHAP.    4 


would  become  zero  at  —  273°C.  In  terms  of  the 
kinetic  theory,  the  motion  of  the  molecules 
ceases  at  this  temperature.  The  kinetic  energy 
has  become  zero. 

There  are  great  advantages  to  an  absolute  tem- 
perature scale  that  has  its  zero  point  at  —  273°C. 
Whereas  the  "zero"  of  temperature  in  the  Centi- 
grade scale  is  based  upon  an  arbitrary  tempera- 
ture, selected  because  it  is  easily  measured,  the 
zero  point  of  the  absolute  scale  has  inherent 
significance  in  the  kinetic  theory.  If  we  express 
temperatures  on  an  absolute  temperature  scale, 
we  find  that  the  volume  of  a  fixed  amount  of  gas 
(at  constant  pressure)  varies  directly  with  tem- 
perature* Also,  the  pressure  of  a  fixed  amount 
of  ?ns  (at  constant  volume)  varies  directly  with 
temperature.  And,  according  to  the  kinetic  the- 
ory, the  kinetic  energy  of  the  molecules  varies 
directly  with  the  absolute  temperature.  For  these 
reasons,  in  dealing  with  gas  relations,  we  shall 
usually  express  temperature  on  an  absolute  tem- 
perature scale. 

This  temperature  scale,  with  the  same  size 
degrees  as  the  Centigrade  scale,  is  called  the 
Kelvin  scale  and  values  on  this  scale  are  ex- 
pressed in  degrees  Kelvin  (°K).  Both  Kelvin  and 
Centigrade  temperatures  are  shown  in  Figure 
4-5.  Notice  that  all  numerical  values  on  the 
Kelvin  scale  are  273  degrees  higher  than  the 
corresponding  temperatures  on  the  Centigrade 
scale. 


EXERCISE  4-5 

(a)  Express  the  following  temperatures  in  de- 
grees Kelvin: 

Boiling  point  of  water :  1 00°C 

Freezing  point  of  mercury:  —  38.9°C 

Boiling  point  of  liquid  nitrogen:     —  196°C 

(b)  Express  the  following  temperatures  in  de- 
grees Centigrade: 


Melting  point  of  lead : 

A  normal  room  temperature: 

Boiling  point  of  liquid  helium: 


600°K 

298°K 

4°K 


*  This  direct  relation  between  volume  and  temperature 
(at  constant  pressure)  is  called  Charles'  Law. 


EXERCISE  4-6 

In  Experiment  9  a  student  obtained  the  result 
that  2.00  X  10~3  mole  of  magnesium  produced 
a  volume  of  hydrogen  that  would  occupy  49.0 
ml  at  25°C  and  one  atmosphere  pressure. 

(a)  If  one  mole  of  magnesium  produces  one  mole 
of  hydrogen,  use  these  data  to  calculate  the 
volume  of  one  mole  of  hydrogen  at  25°C 
(298°K)  and  one  atmosphere. 

(b)  Calculate  the  volume  one  mole  of  hydrogen 
would  occupy  at  0°C  (273°K)  and  one  at- 
mosphere. 


We  have  remarked  that  a  temperature  of  zero 
on  the  absolute  temperature  scale  would  cor- 
respond to  the  absence  of  all  motion.  The  kinetic 
energy  would  become  zero.  Very  interesting  phe- 
nomena occur  at  temperatures  near  0°K  (the 
superconductivity  of  many  metals  and  the  super- 
fluidity of  liquid  helium  are  two  examples). 
Hence,  scientists  are  extremely  interested  in 
methods  of  reaching  temperatures  as  close  to 
absolute  zero  as  possible.  Two  low  temperature 
coolants  commonly  used  are  liquid  hydrogen 
(which  boils  at  20°K)  and  liquid  helium  (which 
boils  at  4°K).  Helium,  under  reduced  pressure, 
boils  at  even  lower  temperatures  and  provides  a 
means  of  reaching  temperatures  near  1°K.  More 
exotic  techniques  have  been  developed  to  pro- 
duce still  lower  temperatures  (as  low  as  0.00 1°K) 
but  even  thermometry  becomes  a  severe  problem 
at  such  temperatures. 

4-2.5    Avogadro's  Hypothesis  and 
the  Kinetic  Theory 

The  kinetic  theory  is  based  upon  the  premise 
that  if  two  gases  are  at  the  same  temperature,  the 
molecules  of  the  gases  have  the  same  average 
kinetic  energy.  The  ability  of  this  kinetic  theory 
to  explain  Avogadro's  Hypothesis  is  one  of  its 
most  important  successes. 

We  may  state  Avogadro's  Hypothesis  in  this 
form:  If  two  gases  at  the  same  temperature  have 
the  same  number  of  particles  in  a  given  volume, 
they  must  exert  the  same  pressure.  Yet,  as  re- 


SEC.    4-2    I    THE    KINETIC    THEORY 


59 


marked  in  Section  4-2.3,  the  mass  of  a  molecule, 
as  well  as  its  velocity,  should  influence  the  pres- 
sure exerted.  If  the  molecules  of  our  two  gas 
samples  have  different  masses,  they  must  have 
different  speeds  in  order  to  have  the  same  kinetic 
energies.  The  lighter  molecules  must  travel  faster, 
so  they  will  strike  the  container  walls  more  times 
per  second.  The  effect  of  the  more  frequent  col- 
lisions exactly  counteracts  the  lower  "push"  per 
collision  from  these  lower-mass  molecules.  The 
result  is  in  perfect  accord  with  Avogadro's  Hy- 
pothesis: Two  gases  at  the  same  concentration 
and  at  the  same  temperature  exert  the  same 
pressure  even  though  their  molecules  have  dif- 
ferent masses. 

Avogadro's  Hypothesis  can  be  shown  quite  readily  in 
an  approximate  way.  The  kinetic  energy  of  a  moving 
particle  is  expressed  by  the  equation 

KE  =  *mv2  (5) 

where  m  is  the  mass  of  the  particle  and  v  is  the  velocity. 
Therefore,  for  gas  A  and  gas  B  at  the  same  temperature, 
we  have 

(KE)A  =  (KE)B  (6) 

IntAVA2  =  jmaVfl1 
or 

rriAVA1  =  mBVBl  (7) 

Now  suppose  we  place  n  molecules  in  a  cubical  box  of 
dimension  d.  The  pressure  is  fixed  by  the  number  of  wall 
collisions  per  second  on  each  square  centimeter  times  the 
momentum  transferred  per  collision: 

_  _  /collisionsN/    1    \/ momentum \     .» 

\  second  /\area/\   collision   / 

Momentum  depends  upon  mass  and  velocity.  The  par- 
ticle approaches  the  wall  with  momentum  mv  and  leaves 
with  this  same  momentum  in  the  opposite  direction.  The 
momentum  transferred  to  the  wall  is,  then 

Momentum  =  2mv  (9) 

The  collisions  per  second  with  the  wall,  on  the  other 
hand,  depend  upon  the  container  dimension  and  the 
velocity  (as  the  molecule  bounces  back  and  forth  between 
the  walls).  We  can  assume  that  one-third  of  the  molecules 
bounce  back  and  forth  in  a  given  direction  between  two 
opposite  walls.  Then  if  there  are  n  molecules  in  the  con- 
tainer, there  are  «/3  hitting  these  two  walls.  One  of  these 
walls  receives  a  collision  each  time  one  of  the  molecules 
travels  the  box  dimension  d  and  back,  a  distance  of  2d. 

collisions  _  /no.  particles  bouncing  back  and  forth \ 
second    ~  \  time  for  a  particle  to  travel  distance  2d) 

collisions  _  /  w/3  \  _  (n\(  v_\  _  nv  .... 

second    "  \2d/v)  ~  \3j\2d)  ~  6d  (    } 


Combining  (8),  (9),  and  (10),  we  find 

_  /collisions\/    1    N/momentumX 
—  \  second   /\area/\   collision   / 

Applying  equation  (7/)  to  each  of  the  gases  A  and  B, 

PA  =  \(^)(mAVA')  U2) 

Pb  =  \  (^)(»W)  (13) 

If  the  gases  have  the  same  pressure,  Pa  =  Pb,  we  can 
equate  (12)  and  (13)  so  that 

If  the  temperatures  of  the  gases  are  the  same,  equation 
(7)  is  applicable  and  equation  (14)  becomes 


tiA      riB 
<P  ~  d* 


(15) 


Thus  we  see  that  at  the  same  temperature  and  pressure, 
the  two  gases  have  the  same  number  of  molecules  per 
unit  volume.  This  is  Avogadro's  Hypothesis. 


4-2.6    The  Perfect  Gas 

We  have  examined  experimental  pressure-volume  data 
for  oxygen  gas  (Table  l-II),  ammonia  gas  (Table  4-1), 
and  hydrogen  chloride  gas  (Table  2-II).  In  each  case, 
within  the  experimental  uncertainty  of  the  data  shown, 
the  gases  have  the  regular  behavior,  PV  =  a  constant. 
We  find  from  many  such  experiments  that  many  gases 
follow  this  simple  behavior.  Of  course  such  a  generaliza- 
tion is  subject  to  uncertainty,  as  is  any  other  scientific 
statement.  The  generalization  was  derived  from  a  set  of 
measurements,  each  of  which  involves  some  uncertainty, 
and,  hence,  the  constancy  of  the  PV  product  is  established 
only  within  corresponding  bounds  of  uncertainty.  What's 
more,  there  are  limits  to  the  pressure  range  over  which 
the  behavior  has  been  tested. 

To  be  specific,  consider  the  data  for  17.0  grams  of 
ammonia  gas  at  25°C,  as  presented  in  Table  4-1  (p.  51). 
These  data  show  that  PV  =  24.5,  but  a  complete  state- 
ment should  include  both  the  uncertainty  and  the  range 
over  which  the  data  are  known  to  apply.  In  this  case  the 
uncertainty  is  ±0.7  and  the  range  is  0.2-2  atmospheres 
pressure.  It  would  not  be  safe,  from  these  data  alone,  to 
assume  that  the  pressure-volume  product  is  constant  to 
four  significant  figures,  PV  =  24.50.  Neither  would  it  be 
safe  to  assume  that  the  pressure-volume  product  is  con- 
stant outside  the  range  of  pressure  studied,  0.2-2.0  at- 
mospheres. Remember,  a  generalization  is  reliable  within 


60 


THE    GAS    PHASE:    KINETIC    THEORY   I    CHAP.    4 


the  bounds  defined  by  the  experiments  that  led  to  the  rule. 
If  we  need  four  significant  figures,  or  wish  to  know  the 
behavior  at  a  higher  pressure,  more  experiments  are 
needed. 

More  accurate  pressure-volume  measurements  extend- 
ing to  much  higher  pressures  have  been  performed.  Table 
4-IV  shows  the  results  of  such  experiments. 


Table  4-IV 

ACCURATE  PRES 
M  EASU  REM  ENT 
OF    AMMONIA    G 

PRESSURE 

(atmospheres) 


SURE-VO 
S  FOR  17 
AS    AT    25 

VOLUME 

(liters) 


LUME 

00    GRAMS 

C 

PXV 


0.1000 

244.5 

24.45 

0.2000 

122.2 

24.44 

0.4000 

61.02 

24.41 

0.8000 

30.44 

24.35 

2.000 

12.17 

24.34 

4.000 

5.975 

23.90 

8.000 

2.925 

23.40 

9.800 

2.360 

23.10 

condensation 
beginning 

9.800 

0.020 

0.20 

no  gas  left; 
liquid  only 

20.00 

0.020 

0.40 

liquid  only 
present 

50.00 

0.020 

1.0 

liquid  only 
present 

The  most  startling  fact  revealed  in  Table  4-IV  is  the 
drastic  deviation  from  PV  =  24.5  that  occurs  when  the 
pressure  is  raised  above  9.800  atmospheres.  Suddenly  the 
relation  PV  =  a  constant  is  no  longer  applicable.  Here 


is  dramatic  evidence  of  the  danger  lurking  in  careless 
extrapolation  beyond  the  range  of  experience. 

Even  below  the  condensation  pressure  the  pressure- 
volume  product  was  not  perfectly  constant.  With  meas- 
urements of  sufficient  accuracy  and  precision,  we  can  see 
that  the  PV  product  of  ammonia  at  25°C  is  not  really 
constant  after  all.  It  varies  systematically  from  24.45  at 
0.1000  atmospheres  to  23.10  at  9.800  atmospheres,  just 
before  condensation  begins.  Similar  measurements  on 
28.0  grams  of  carbon  monoxide  at  0°C  show  that  the  PV 
product  is  22.410  at  0.2500  atmospheres  pressure,  but  if 
the  pressure  is  raised  to  4.000  atmospheres,  the  PV  prod- 
uct becomes  22.308.  This  type  of  deviation  is  common. 
Careful  measurements  reveal  the  fact  that  no  gas  follows 
perfectly  the  generalization  PV  =  a  constant  at  all  pres- 
sures. On  the  other  hand,  every  gas  follows  this  rule 
approximately,  and  the  fit  becomes  better  and  better  as 
the  pressure  is  lowered.  So  we  find  that  every  gas  ap- 
proaches the  behavior  PV  =  a  constant  as  pressure  is 
lowered. 

There  is  a  reasonable  explanation  for  this  type  of 
deviation.  The  kinetic  theory,  which  "explains"  the 
pressure-volume  behavior,  is  based  upon  the  assumption 
that  the  particles  exert  no  force  on  each  other.  But  real 
molecules  do  exert  force  on  each  other!  The  condensation 
of  every  gas  on  cooling  shows  that  there  are  always 
attractive  forces.  These  forces  are  not  very  important 
when  the  molecules  are  far  apart  (that  is,  at  low  pressures) 
but  they  become  noticeable  at  higher  pressures.  With  this 
explanation,  we  see  that  the  kinetic  theory  is  based  on  an 
"idealized"  gas — one  for  which  the  molecules  exert  no 
force  on  each  other  whatsoever.  Every  gas  approaches 
such  ideal  behavior  if  the  pressure  is  low  enough.  Then 
the  molecules  are,  on  the  average,  so  far  apart  that  their 
attractive  forces  are  negligible.  A  gas  that  behaves  as 
though  the  molecules  exert  no  force  on  each  other  is 
called  an  ideal  gas  or  a  perfect  gas. 


Table  4- V.    molar  volumes  of  some  gases 


FORMULA 


MOLAR  WEIGHT 

(grams) 


MOLAR  VOLUME  AT 
0°C  AND  1  ATM 

(liters) 


hydrogen 

H2 

2.0160 

22.430 

helium 

He 

4.003 

22.426 

("perfect"  gas) 

— 

— 

(22.414) 

nitrogen 

N2 

28.016 

22.402 

carbon  monoxide 

CO 

28.011 

22.402 

oxygen 

o2 

32.000 

22.393 

methane 

CH4 

16.043 

22.360 

carbon  dioxide 

co2 

44.011 

22.262 

hydrogen  chloride 

HC1 

36.465 

22.248 

ammonia 

NH3 

17.032 

22.094 

chlorine 

Cl2 

70.914 

22.063 

sulfur  dioxide 

so2 

64.066 

21.888 

QUBSTIONS    AND    PROBLEMS 


61 


Avogadro's  Hypothesis  is  consistent  with  the  kinetic 
theory.  Therefore  a  perfect  gas  follows  Avogadro's  Hy- 
pothesis. At  one  atmosphere  pressure  and  0°C,  one  mole 
(6.02  X  10"  molecules)  of  a  perfect  gas  occupies  22.414 
liters.  How  closely  real  gases  approximate  a  perfect  gas 
at  one  atmosphere  pressure  and  0°C  is  shown  by  measur- 


ing the  volume  occupied  by  one  mole  of  that  gas,  the  molar 
volume.  Table  4-V  shows  the  molar  volumes  of  a  num- 
ber of  gases.  We  see  that  real  gases  do  approximate 
closely  (to  three  significant  figures)  the  perfect  gas  be- 
havior at  one  atmosphere  and  0°C.  Every  gas  becomes  a 
perfect  gas  as  the  pressure  is  reduced  toward  zero. 


4-3    REVIEW 


Regularities  observed  in  the  behavior  of  gases 
have  contributed  much  to  our  understanding  of 
the  structure  of  matter.  One  of  the  most  im- 
portant regularities  is  Avogadro's  Hypothesis: 
Equal  volumes  of  gases  contain  equal  numbers 
of  particles  (at  the  same  pressure  and  tempera- 
ture). This  relationship  is  valuable  in  the  deter- 
mination of  molecular  formulas — these  formulas 
must  be  known  before  we  can  understand  chemi- 
cal bonding. 

We  have  explored  the  meaning  of  temperature. 
According  to  the  kinetic  theory,  when  two  gases 
are  at  the  same  temperature,  the  molecules  of  the 
two  gases  have  the  same  average  kinetic  energies. 
Changing  the  temperature  of  a  sample  of  gas  at 
constant  pressure  reveals  that  the  volume  is  di- 


rectly proportional  to  the  temperature  if  the 
temperature  is  expressed  in  terms'  of  a  new, 
absolute  scale.  The  melting  point  of  ice  (0°C)  on 
this  new  scale  (called  the  Kelvin  scale)  is  273°K. 
The  boiling  point  of  water  at  one  atmosphere 
(100°C)  is  373°K.  The  zero  temperature  on  the 
Kelvin  scale  corresponds  to  the  hypothetical  loss 
of  all  molecular  motion. 

This  progress  gives  us  substantial  basis  for 
confidence  in  the  usefulness  of  the  atomic  theory 
and  it  encourages  us  to  develop  the  model  fur- 
ther. We  shall  see  that  the  concepts  we  have 
developed  in  our  consideration  of  gases  are  also 
useful  when  we  consider  the  behavior  of  con- 
densed phases— liquids  and  solids. 


QUESTIONS  AND  PROBLEMS 

1.  How  many  molecules  are  there  in  a  molar  vol- 
ume of  a  gas  at  100°C?  At  0°C? 

2.  What  is  the  molar  volume  of  water  under  each 
of  the  following  conditions? 

(a)  Solid,  0°C; 

density  of  ice  =  0.915  g/ml. 

(b)  Liquid,  0°C; 

density  of  water  (liquid,  0°C)  =  1.000  g/ml. 

(c)  Gas,  100°C; 

density  of  water  vapor  (100°C,  1  atm)  = 
5.88  X  10~4  g/ml. 

3.  What  is  the  molecular  weight  of  a  gas  if  at  0°C 

and  one  atmosphere  pressure,  1.00  liter  of  the 

gas  weighs  2.00  grams? 

Answer.  44.8  g/mole 

4.  The  gas  sulfur  dioxide  combines  with  oxygen  to 
form  the  gas  sulfur  trioxide: 


What  ratio  would  you  expect  for  the  following? 
number  of  SQ3  molecules  produced 
number  of  02  molecules  consumed 
volume  of  SQ3  gas  produced 
volume  of  02  gas  consumed 


(a) 
(b) 


5.  A  glass  bulb  weighs  108.1 1  grams  after  all  of  the 
gas  has  been  removed  from  it.  When  filled  with 
oxygen  gas  at  atmospheric  pressure  and  room 
temperature,  the  bulb  weighs  109.56  grams. 
When  filled  at  atmospheric  pressure  and  room 
temperature  with  a  gas  sample  obtained  from 
the  mouth  of  a  volcano,  the  bulb  weighs  111.01 
grams.  Which  of  the  following  molecular  for- 
mulas for  the  volcano  gas  could  account  for  the 
data? 


2SO,(gas)  +  O-Xgas) 


2SO,(gas) 


co2 
ocs 

Si2H6 

so2 

NF3 


S03 

S8 

A  gas  mixture,  half  C02, 
half  Kr 


62 


THE    GAS    PHASE!    KINETIC    THEORY    I    CHAP.    4 


6.  Compressed  oxygen  gas  is  sold  at  a  pressure  of 
130  atm  in  steel  cylinders  of  40  liters  volume. 

(a)  How  many  moles  of  oxygen  does  such  a 
filled  cylinder  contain  ? 

(b)  How  many  kilograms  of  oxygen  are  in  the 

cylinder? 

Answer.  6.7  kg. 

7.  A  carbon  dioxide  fire  extinguisher  of  3  liters 
volume  contains  about  10  pounds  (4.4  kg)  of 
C02.  What  volume  of  gas  could  this  extinguisher 
deliver  at  room  conditions? 

8.  Hydrogen  for  weather  balloons  is  often  supplied 
by  the  reaction  between  solid  calcium  hydride, 
CaH2,  and  water  to  form  solid  calcium  hydrox- 
ide, Ca(OH)2,  and  hydrogen  gas,  H2. 

(a)  Balance  the  equation  for  the  reaction  and 
decide  how  many  moles  of  CaH2  would  be 
required  to  fill  a  weather  balloon  with  250 
liters  of  hydrogen  gas  at  normal  conditions. 

(b)  What  weight  of  water  would  be  consumed  in 

forming  the  hydrogen  ? 

Answer.  0.18  kg. 

9.  Gas  is  slowly  added  to  the  empty  chamber  of  a 
closed-end  manometer  (see  Figure  4-2B).  Draw 
a  picture  of  the  manometer  mercury  levels,  show- 
ing in  millimeters  the  difference  in  heights  of  the 
two  mercury  levels: 

(a)  before  any  gas  has  been  added  to  the  empty 
gas  chamber; 

(b)  when  the  gas  pressure  in  the  chamber  is 
300  mm; 

(c)  when  the  gas  pressure  in  the  chamber  is 
760  mm; 

(d)  when  the  gas  pressure  in  the  chamber  is 
865  mm. 

10.  Repeat  Problem  9  but  with  an  open-end  ma- 
nometer (see  Figure  4-2C).  Atmospheric  pres- 
sure is  760  mm. 

1 1 .  The  balloons  that  are  used  for  weather  study  are 
quite  large.  When  they  are  released  at  the  surface 
of  the  earth  they  contain  a  relatively  small  vol- 
ume of  gas  compared  to  the  volume  they  acquire 
when  aloft.  Explain. 

12.  A  1.50  liter  sample  of  dry  air  in  a  cylinder  exerts 
a  pressure  of  3.00  atm  at  a  temperature  of  25°C. 
Without  change  in  temperature,  a  piston  is 
moved  in  the  cylinder  until  the  pressure  in  the 


cylinder  is  reduced  to  1.00  atm.  What  is  the 
volume  of  the  gas  in  the  cylinder  now  ? 

13.  Suppose  the  total  pressure  in  an  automobile  tire 
is  30  pounds/inch2  and  we  want  to  increase  the 
pressure  to  40  pounds/inch2.  What  change  in  the 
amount  of  air  in  the  tire  must  take  place  ?  As- 
sume that  the  temperature  and  volume  of  the  tire 
remain  constant. 

14.  The  density  of  liquid  carbon  dioxide  at  room 
temperature  is  0.80  grams/ml.  How  large  a  car- 
tridge of  liquid  C02  must  be  provided  to  inflate 
a  life  jacket  of  4.0  liters  capacity  at  STP? 

15.  A  student  collects  a  volume  of  hydrogen  over 
water.  He  determines  that  there  is  2.00  X  10~3 
mole  of  hydrogen  and  6.0  X  10~6  mole  of  water 
vapor  present.  If  the  total  pressure  inside  the 
collecting  tube  is  760  mm,  what  is  the  partial 
pressure  of  each  gas? 

Answer.  Partial  pressure  H2     =  738  mm. 
Partial  pressure  H20  =    22  mm. 

16.  A  sample  of  nitrogen  is  collected  over  water  at 
18.5°C.  The  vapor  pressure  of  water  at  18.5°C 
is  16  mm.  When  the  pressure  on  the  sample  has 
been  equalized  against  atmospheric  pressure,  756 
mm,  what  is  the  partial  pressure  of  nitrogen? 
What  will  be  the  partial  pressure  of  nitrogen  if 
the  volume  is  reduced  by  a  factor  740/760? 

17.  A  candle  is  burned  under  a  beaker  until  it  ex- 
tinguishes itself.  A  sample  of  the  gaseous  mix- 
ture in  the  beaker  contains  6.08  X  1020  molecules 
of  nitrogen,  0.76  X  1020  molecules  of  oxygen, 
and  0.50  X  1020  molecules  of  carbon  dioxide. 
The  total  pressure  is  764  mm.  What  is  the  partial 
pressure  of  each  gas? 

18.  A  cylinder  contains  nitrogen  gas  and  a  small 
amount  of  liquid  water  at  a  temperature  of  25°C 
(the  vapor  pressure  of  water  at  25°C  is  23.8  mm). 
The  total  pressure  is  600.0  mm  Hg.  A  piston  is 
pushed  into  the  cylinder  until  the  volume  is 
halved.  What  is  the  final  total  pressure  ? 

Answer.  1176  mm. 

19.  Consider  two  closed  glass  containers  of  the  same 
volume.  One  is  filled  with  hydrogen  gas,  the 
other  with  carbon  dioxide  gas,  both  at  room 
temperature  and  pressure. 

(a)  How  do  the  number  of  moles  of  the  two 
gases  compare? 


QUESTIONS    AND    PROBLEMS 


63 


(b)  How  do  the  number  of  molecules  of  the  two 
gases  compare? 

(c)  How  do  the  number  of  grams  of  the  two 
gases  compare? 

(d)  If  the  temperature  of  the  hydrogen  container 
is  now  raised,  how  do  the  two  gases  now 
compare  in: 

(i)  pressure, 

(ii)  volume, 
(iii)  number  of  moles, 
(iv)  average  molecular  kinetic  energy. 


20.  The  boiling  points  and  freezing  points  in  degrees 
Centigrade  of  certain  liquids  are  listed  below. 
Express  these  temperatures  on  the  absolute  tem- 
perature (degree  Kelvin)  scale. 

Liquid  helium,         boiling  point    =  —269 
Liquid  hydrogen,    freezing  point  =  —259 

Answer.  14°K. 
Liquid  hydrogen,     boiling  point    =  —253 

Answer.  20°K. 

Liquid  nitrogen,      freezing  point  =  —210 
Liquid  nitrogen,      boiling  point    =  —196 

Liquid  oxygen,        freezing  point  =  —  219 
Liquid  oxygen,        boiling  point    =  — 183 


21.  If  exactly  100  ml  of  a  gas  at  10°C  are  heated  to 
20°C  (pressure  and  number  of  molecules  remain- 
ing constant),  the  resulting  volume  of  the  gas 
will  be  which  of  the  following? 

(a)  50  ml, 

(b)  1000  ml, 

(c)  100  ml, 

(d)  375  ml, 

(e)  103  ml. 


(a)  Will  the  final  pressure  be  greater  or  lower 
than  the  original  pressure? 

(b)  By  what  factor  does  the  pressure  change  if 
one  mole  of  methane  and  one  mole  of  oxygen 
are  mixed  and  reacted  (with  the  temperature 
changing  from  25°C  to  200°Q? 

Answer.  2.38. 

24.  Automobiles  are  propelled  by  burning  gasoline, 
typical  formula  C8Hi8,  inside  a  container  (the 
cylinder)  that  can  change  volume  and  drive  the 
wheels.  Oxygen  reacts  with  the  gasoline  to  form 
carbon  dioxide  and  water,  releasing  enough  en- 
ergy to  heat  the  gas  from  about  300°K  to  about 
1500°K. 

Balance  the  equation  for  the  reaction  and 
decide  whether  the  work  done  by  the  gas  in  the 
cylinder  is  mainly  due  to  pressure  rise  caused  by 
change  in  number  of  moles  of  gas  or  due  to 
pressure  rise  resulting  from  heating. 

25.  Why  does  the  pressure  build  up  in  a  tire  on  a 
hot  day?  Answer  in  terms  of  the  kinetic  theory. 

26.  A  vessel  contains  equal  numbers  of  oxygen  and 
of  hydrogen  molecules.  The  pressure  is  760  mm 
Hg  when  the  volume  is  50  liters.  Which  of  the 
following  statements  is  FALSE? 

(a)  On  the  average,  the  hydrogen  molecules  are 
traveling  faster  than  the  oxygen  molecules. 

(b)  On  the  average,  more  hydrogen  molecules 
strike  the  walls  per  second  than  oxygen 
molecules. 

(c)  If  the  oxygen  were  removed  from  the  system, 
the  pressure  would  drop  to  190  mm  Hg. 

(d)  Equal  numbers  of  moles  of  each  gas  are 
present. 

(e)  The  average  kinetic  energies  of  oxygen  and 
hydrogen  are  the  same. 


22.  Why  is  it  desirable  to  express  all  temperatures 
in  degrees  Kelvin  when  working  with  problems 
dealing  with  gas  relationships? 


23.  A  gaseous  reaction  between  methane,  CH4,  and 
oxygen,  02,  is  carried  out  in  a  sealed  container. 
Under  the  conditions  used,  the  products  are 
hydrogen,  H2,  and  carbon  dioxide,  C02.  Energy 
is  released,  so  the  temperature  rises  during  the 
reaction. 


27.  The  vapor  pressure  of  a  molten  metal  can  be 
measured  with  a  device  called  a  Knudsen  cell. 
This  is  a  container  closed  across  the  top  by  a 
thin  foil  pierced  by  a  small,  measured  hole.  The 
cell  is  heated  in  a  vacuum,  until  the  vapor  above 
the  melt  streams  from  the  small  hole  (it  effuses). 
The  weight  of  the  material  escaping  per  second 
tells  the  rate  at  which  gaseous  atoms  leave. 

Two  identical  Knudsen  cells  are  heated  at 
1000°C,  one  containing  lead  and  the  other  con- 
taining magnesium. 


64 


THE    GAS    PHASE:     KINETIC    THEORY    I    CHAP.    4 


(a)  Contrast  the  average  kinetic  energies  of  the 
lead  and  magnesium  atoms  within  each  cell. 

(b)  Contrast  the  average  velocities  of  the  lead 
and  magnesium  atoms  leaving  each  cell. 

(c)  At  this  fixed  temperature,  the  rate  at  which 
atoms  leave  is  determined  by  two  factors, 
the  vapor  pressure  and  the  mass  of  the 
gaseous  particles.  Explain. 

28.  The  following  table  indicates  the  boiling  points 
and  the  molar  volumes  (0°C  and  1  atm)  of  some 
common  gases. 

(a)  What  regularity  is  suggested  in  the  relation- 
ship between  the  boiling  points  and  molar 
volumes? 

(b)  Account  for  this  regularity. 


GAS 


FORMULA 


BOILING 
POINT 
(°C) 


MOLAR 
VOLUME 

(liters) 


helium 

He 

-269 

22.426 

nitrogen 

N2 

-196 

22.402 

carbon 

monoxide 

CO 

-190 

22.402 

oxygen 

o2 

-183 

22.393 

methane 

CH4 

-161 

22.360 

hydrogen 

chloride 

HC1 

-84.0 

22.248 

ammonia 

NH3 

-33.3 

22.094 

chlorine 

a, 

-34.6 

22.063 

sulfur 

dioxide 

SO: 

-10.0 

21.888 

CHAPTER 


5 


Liquids  and  Solids: 
Condensed  Phases 
of  Matter 


Almost  all  the  chemical  processes  which  occur  in  nature,  whether  in 
animal  or  vegetable  organisms,  or  in  the  non-living  surface  of  the  earth, 
•    •    •   take  place  between  substances  in  solution. 

w.  ostwald,   1890 


Only  a  handful  of  substances  are  gases  under 
normal  conditions  of  temperature  and  pressure. 
Of  the  hundred  or  so  elements,  most  are  nor- 
mally solids;  two  or  three  are  liquids.  As  for 
compound  substances,  more  than  a  million  have 
been  prepared  by  chemists,  yet,  more  than  99% 
of  these  are  liquids  or  solids,  each  with  distinc- 
tive and  characteristic  properties.  It  is  no  sur- 
prise, then,  that  there  is  great  variety  among  all 
of  these  substances.  Rather,  it  is  remarkable  that 
they  can  be  classified  into  a  small  number  of 
types  and  that  the  wealth  of  information  repre- 
sented by  the  diversity  of  all  of  these  compounds 


can  be  treated  within  a  simple  framework.  We 
shall  begin  our  study  of  this  framework  by  con- 
sidering the  properties  of  pure  substances  in  their 
liquid  and  solid  phases. 

EXERCISE  5-1 

The  dozen  or  so  elements  that  are  normally 
found  as  gases  include  nitrogen,  oxygen,  fluo- 
rine, helium,  neon,  argon,  krypton,  xenon,  and 
chlorine.  Where  are  these  placed  in  the  periodic 
table  (see  inside  front  cover)? 


5-1    PURE  SUBSTANCES 

A  gas,  when  cooled,  condenses  to  a  liquid.  Fur- 
ther cooling  causes  solidification.  The  condensa- 
tion of  ammonia  under  pressure,  the  condensa- 
tion of  water  vapor  (steam)  on  cooling,  and  the 


freezing  of  liquid  water  to  ice  are  familiar  cases. 
These  changes  are  called  phase  changes.  We  shall 
consider  liquid-gas  changes  first  and  then,  solid- 
liquid  phase  changes. 

65 


66 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    |    CHAP.    5 


5-1.1    Liquid-Gas  Phase  Changes 

When  a  pan  of  water  is  warmed,  the  input  of 
heat  causes  the  water  temperature  to  rise.  At  a 
certain  point,  however,  the  water  begins  to  boil. 
Then  the  temperature  is  constant  as  long  as 
liquid  water  remains,  and  continued  heating 
causes  the  formation  of  water  vapor.  Water 
changes  from  the  liquid  phase  to  the  gas  phase, 
absorbing  energy  though  the  temperature  re- 
mains constant.  The  energy  of  the  liquid  is  less 
than  the  energy  of  the  same  weight  of  gas. 

Let  us  consider  how  much  energy  is  needed 
for  this  particular  phase  change. 

H20(liquid)  — ►-  H20(gas)  (7) 

or,  in  abbreviation, 

H2OfU         — ►-  H20(g)  (7) 

Suppose  we  wish  to  evaporate  one  mole  of  water, 
as  expressed  in  equation  (7).  One  mole  contains 
the  Avogadro  number  of  molecules  (6.02  X  1023) 
and  has  a  weight  of  18.0  grams.  Using  a  calo- 
rimeter, as  you  did  in  Experiment  5,  you  could 
measure  the  quantity  of  heat  required  to  evapo- 
rate one  mole  of  water.  It  is  10  kilocalories  per 
mole.  This  value  is  called  the  molar  heat  of 
vaporization  of  water.  This  is  the  energy  re- 
quired to  separate  6.02  X  1023  molecules  of  water 
from  one  another,  as  pictured  in  Figure  5-1. 

Fig.  5-1.  Evaporation  of  liquid  water. 


EXERCISE  5-2 


+  10 


kcal 
mole 


One  mole. 
liquid   water    +    10   kcal 

HzO(t) 

6.02  *  1023 molecules 
weight  18  g 


When  two  moles  of  water  are  evaporated,  how 
much  heat  is  required?  One-half  mole  of  water? 


When  water  vapor  condenses  to  liquid  water, 
the  molecules  release  the  energy  it  took  to  sepa- 
rate them.  A  mole  of  gaseous  water,  therefore, 
will  release  10  kilocalories  of  heat  when  con- 
densed to  liquid  water  at  the  same  temperature. 
The  amount  of  heat  released  is  numerically  equal 
to  the  molar  heat  of  vaporization. 

Other  liquid-gas  phase  changes  are  similar, 
though  boiling  points  vary  over  a  wide  range. 
Table  5-1  shows  the  boiling  points  and  heats  of 
vaporization  of  a  variety  of  liquids.  In  each  case, 
energy  is  absorbed  as  the  particles  that  make  up 
the  liquid  are  separated  into  the  molecules  of  the 
gas.  We  shall  see  in  Chapter  17  that  the  extreme 
range  of  heats  of  vaporization  shown  in  this 
table  can  be  explained  using  rather  simple  prin- 
ciples. These  principles  provide  a  basis  for  quali- 
tative predictions  of  boiling  point,  heat  of 
vaporization,  and  other  properties. 

5-1.2    Liquid-Gas  Equilibrium:  Vapor  Pressure 

Our  knowledge  of  gas  behavior  helps  us  interpret 
the  evaporation  of  liquids.  We  have  considered, 
thus  far,  vaporization  of  a  liquid  at  its  usual 


one  mole   water  vapor 

H20($) 
6.02  x  1023 molecules 
"weight  18  g 


SEC.    5-1    !    PURE    SUBSTANCES 


67 


Table  5-1.     the  normal  boiling  points  and  molar  heats  of  vaporization 

OF    SOME    PURE    SUBSTANCES 


SUBSTANCE 


PHASE  CHANGE 

(liquid)  — >-  (gas) 


BOILING  POINT 

°K  °C 


MOLAR  HEAT 
OF  VAPORIZATION 

(kcal/mole) 


boiling  point.  But  liquids  vaporize  at  all  tempera- 
tures. Let  us  consider  this  process,  beginning 
with  liquid  water  again. 

If  we  place  some  liquid  water  in  a  flask  at 
20°C  and  seal  the  flask,  some  water  molecules 
leave  the  liquid  and  enter  the  gas  phase.  The 
partial  pressure  of  water  vapor  in  the  flask  rises, 
but  when  it  reaches  17.5  mm  no  more  change 
can  be  observed.  The  amount  of  excess  liquid 
remains  constant  thereafter,  and  the  partial 
pressure  of  water  vapor  in  the  flask  remains  at 
17.5  mm,  as  long  as  the  temperature  is  main- 
tained at  20°C.  This  partial  pressure  is  called  the 
vapor  pressure  of  water  at  20°C.  At  this  vapor 
pressure,  liquid  and  gaseous  water  can  coexist 
indefinitely  at  20°C.  This  vapor  pressure,  17.5 
mm,  is  the  same  whether  air  is  present  or  not; 
it  is  a  property  of  water.  If  the  flask  were  origi- 
nally evacuated,  liquid  would  evaporate  until  the 
pressure  rose  from  0  mm  to  17.5  mm.  If  the  flask 
originally  contained  dry  air  at  a  pressure  of  750 
mm,  liquid  would  evaporate  until  the  pressure 
rose  from  750  mm  to  767.5  mm  (the  partial  pres- 
sure of  water  vapor  changing  from  0  mm  to  17.5 
mm).  When  a  liquid  is  in  contact  with  its  vapor 
at  the  vapor  pressure,  the  liquid  and  gas  are  said 
to  be  in  equilibrium.  At  equilibrium,  no  measur- 
able changes  are  taking  place. 

EFFECT   OF   TEMPERATURE 

The  vapor  pressure  of  water  at  20°C  is  17.5  mm. 
At  40°C,  the  vapor  pressure  is  55.3  mm;  at  60°C, 
it  is  149.4  mm.  The  vapor  pressure  of  water  in- 
creases with  increasing  temperature. 


Ethyl  alcohol  is  also  a  liquid  at  room  tempera- 
ture. Its  vapor  pressure  at  20°C  is  44  mm,  higher 
than  the  vapor  pressure  of  water  at  this  same 
temperature.  At  40°C,  ethyl  alcohol  has  a  vapor 
pressure  of  134  mm;  at  60°C,  the  vapor  pressure 
is  352  mm.  Again  we  find  that  the  vapor  pressure 
increases  rapidly  with  increasing  temperature. 
This  is  always  so.  The  vapor  pressure  of  every 
liquid  increases  as  the  temperature  is  raised. 

THE    BOILING    POINT 

At  any  temperature,  molecules  can  escape  from 
the  surface  of  a  liquid  (vaporizing  or  evaporat- 
ing) to  enter  the  gas  phase  as  vapor.  At  the  spe- 
cial temperature  at  which  the  vapor  pressure  just 
equals  the  atmospheric  pressure,  a  new  phe- 
nomenon occurs.  There,  bubbles  of  vapor  can 
form  anywhere  within  the  liquid.  At  this  tempera- 
ture, the  liquid  boils. 

We  see  that  the  boiling  point  is  fixed  by  the 
surrounding  pressure.  For  example,  if  the  sur- 
rounding pressure  is  760  mm,  water  boils  at 
100°C.  This  is  the  temperature  at  which  the 
vapor  pressure  of  water  is  just  760  mm.  Ethyl 
alcohol,  having  a  higher  vapor  pressure,  achieves 
a  vapor  pressure  of  760  mm  at  78.5°C.  Ethyl 
alcohol  boils  at  78.5°C  with  this  surrounding 
pressure.  Suppose,  however,  that  the  atmos- 
pheric pressure  drops  to  750  mm  (as  it  might  just 
before  a  storm).  Then  bubbles  of  vapor  could 
form  anywhere  in  liquid  water  at  a  temperature 
of  99.6°C  since  the  vapor  pressure  of  water  is 
750  mm  at  99.6°C.  Water  boils  at  99.6°C  when 
the  surrounding  pressure  is  750  mm. 


68 


LIQUIDS    AND    SOLIDS:    CONDENSED    PH-ASES    OF    MATTER    I    CHAP.    5 


The  normal  boiling  point  of  a  liquid  is  de-  exercise  5-3 

fined  as  the  temperature  at  which  the  vapor  w^  is  the  normal  boiling  point  of  ethyl  al- 

pressure  of  that  liquid  is  exactly  one  standard  cohol? 
atmosphere,  760  mm  Hg. 

EXERCISE  5-4 


Suppose  a  closed  flask  containing  liquid  water  is 
Fig.  5-2.  Vapor  pressure  increases  as  temperature  is      connected  t0  a  vacuum  pump  and  the  pressure 
raised 

over  the  liquid  is  gradually  lowered.  If  the  water 
temperature  is  kept  at  20°C,  at  what  pressure 
will  the  water  boil? 


EXERCISE  5-5 


Answer  Exercise  5-4,  substituting  ethyl  alcohol 
for  water. 


5-1.3    Solid-Liquid  Phase  Changes 

Solids  and  liquids  are  called  condensed  phases. 
The  attractive  forces  in  a  condensed  phase,  either 
a  solid  or  a  liquid,  tend  to  hold  the  molecules 
close  together.  In  liquids,  molecules  are  irregu- 
larly spaced  and  randomly  oriented.  In  a  crystal- 
line solid,  the  molecules  occupy  regular  posi- 
tions, resulting  in  additional  stability  (relative  to 
the  liquid). 

The  difference  between  the  energy  of  a  sub- 
stance in  liquid  form  and  its  energy  in  solid  form 
is  usually  much  smaller  than  the  difference  be- 
tween the  energies  of  the  liquid  and  gaseous 
forms.  For  example,  consider  the  heat  of  melting 
a  mole  of  ice, 


H20(solid) 
or,  in  abbreviation, 


H20(liquid) 


U20(l) 


(2) 


(2) 


Liquid 


in  ice.  hath.,    O  C 


The  heat  accompanying  the  phase  change  (2) 
is  1.44  kcal/mole.  This  is  much  less  than  the  mo- 
lar heat  of  vaporization  of  water,  10  kcal/mole. 
Table  5-II  contrasts  the  melting  points  and  the 
heats  of  melting  per  mole  (the  molar  heat  of 
melting,  or  the  molar  heat  of  fusion)  of  the  same 
pure  substances  listed  in  Table  5-1. 

Once  again,  we  find  an  extreme  range  among 
the  properties  of  these  substances.  The  molar 
heats  of  melting  vary  from  0.080  kcal/mole  for 


SEC.    5-2    I    SOLUTIONS 


69 


Table  5-1 1,     the  melting  points  and  heats  of  melting  of  some  pure 

SUBSTANCES 


SUBSTANCE 


neon 

chlorine 

water 

sodium 

sodium  chloride 

copper 


PHASE  CHANGE 

(solid)  — >■  (liquid) 


MELTING  POINT 
°K  °C 


MOLAR  HEAT 
OF  MELTING 

(kcal/mole) 


0.080 

1.53 

1.44 

0.63 

6.8 

3.11 


neon  to  6.8  kcal/mole  for  the  substance  sodium  very  great  differences  in  the  forces  that  bind 

chloride— a  change  by  a  factor  of  85.  There  are  these  solids.  Since  these  differences  affect  prop- 

erties  other  than  melting  point  and  heat  of  melt- 

Fig.  5-3.  Melting  of  ice.  ~  "»*'  ^  are  imPortant  to  a  chemist. 


+  i.+ 


5-2    SOLUTIONS 


Sodium  chloride,  sugar,  ethyl  alcohol,  and  water 
are  four  pure  substances.  Each  is  characterized 
by  definite  properties,  such  as  vapor  pressure, 
melting  point,  boiling  point,  density.  Suppose  we 
mix  some  of  these  pure  substances.  Sodium  chlo- 
ride dissolves  when  placed  in  contact  with  water. 
The  solid  disappears,  becoming  part  of  the  liq- 
uid. Likewise,  sugar  in  contact  with  water  dis- 
solves. When  ethyl  alcohol  is  added  to  water,  the 
two  pure  substances  mix  to  give  a  liquid  similar 
in  appearance  to  the  original  liquids.  The  salt- 
water mixture,  the  sugar-water  mixture,  and  the 


alcohol-water  mixtures  are  called  solutions.  Solu- 
tions differ  from  pure  substances  in  that  their 
properties  vary,  depending  upon  the  relative 
amounts  of  the  constituents.  The  behavior  of 
solutions  during  phase  changes  is  dramatically 
different  from  that  just  described  for  pure  sub- 
stances. These  differences  provide,  at  once,  rea- 
son for  making  a  distinction  between  pure 
substances  and  solutions  and,  as  well,  a  basis  for 
deciding  whether  a  given  material  is  a  pure  sub- 
stance or  a  solution. 


70 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


5-2.1    Differentiating  Between  Pure 
Substances  and  Solutions 

The  earth  has  many  unlike  parts — it  is  heteroge- 
neous. Some  of  the  parts  are  uniform  through- 
out; that  is,  they  are  homogeneous.  Familiar 
examples  of  heterogeneous  materials  are  granite 
(which  consists  of  various  minerals  suspended  in 
another  mineral),  oil  and  vinegar  salad  dressing 
(which  consists  of  droplets  of  oil  suspended  in 
aqueous  acetic  acid),  and  black  smoke  (which 
consists  of  particles  of  soot  suspended  in  air). 
Examples  of  homogeneous  materials  are  dia- 
mond, fresh  water,  salt  water,  and  clear  air. 
Heterogeneous  materials  are  hard  to  describe 
and  classify  but  we  can  describe  homogeneous 
materials  rather  precisely. 

Both  pure  substances  and  solutions  are  ho- 
mogenous. A  homogeneous  material  that  contains 
only  one  substance  is  called  a  pure  substance. 
A  solution  is  a  homogeneous  material  that  con- 
tains more  than  one  substance. 

We  have  used  the  terms  gas  phase,  liquid 
phase,  and  solid  phase.  A  phase  is  a  homogene- 
ous part  of  a  system — a  part  which  is  uniform 
and  alike  throughout.  A  system  in  turn,  is  any 
region  and  the  material  in  i*  that  we  wish  to 
consider.  It  may  include  oniy  one,  or  more  than 
one,  phase. 

Suppose  we  compare  two  liquid  samples,  one 
of  distilled  water,  and  one  of  salt  water.  Each 
sample  is  a  homogeneous  system  consisting  of  a 
single  phase.  However,  one  of  the  liquids  is  a 
pure  substance  whereas  the  other  is  a  solution. 
We  cannot  tell,  merely  by  visual  observation, 
which  of  these  clear  liquids  is  the  pure  substance 
and  which  is  the  solution.  True,  there  are  dif- 
ferences— for  example,  the  salt  water  has  a 
greater  density  than  the  pure  water — but  even 
this  property  does  not  indicate  which  is  the  pure 
substance. 

Let  us  compare  the  behavior  of  these  two  sys- 
tems during  a  phase  change.  Consider,  first,  how 
water  acts  when  it  is  frozen  or  vaporized.  Pure 
water  freezes  at  a  fixed  temperature,  0°C.  If  we 
freeze  half  of  a  water  sample  to  ice,  remove  the 
ice,  melt  it  in  another  container,  and  compare 
the  separate  samples,  we  find  that  the  two  frac- 
tions of  the  original  sample  are  indistinguishable. 


In  a  similar  way,  if  we  boil  a  sample  of  water 
until  half  of  it  has  changed  to  steam,  condense 
the  steam  to  water  in  a  different  vessel,  and  then 
compare  the  separate  samples,  we  find  that  the 
fractions  of  the  original  sample  are  indistin- 
guishable. Such  behavior  on  boiling  (condensing) 
or  freezing  (melting)  characterizes  pure  sub- 
stances. Solutions  behave  differently. 

Suppose  we  boil  off  part  of  a  sample  of  salt 
water.  The  temperature  of  the  liquid  rises,  as 
shown  in  curve  b  of  Figure  5-4,  until  boiling 


Boiling  begins 


x — * — x — x — x — x — x — x — x — x — x — x — x — x — x- 


Boiling  begins 


•  Salt  water 
x  Pure  water 


Time 

Fig.  5-4.  Behavior  on  boiling,  (a)  A  pure  substance, 
(b)  A  solution. 


begins.  Already,  a  difference  from  the  behavior 
of  pure  water  can  be  noted,  shown  in  curve  a 
of  Figure  5-4:  the  boiling  point  of  the  salt  solu- 
tion is  higher.  As  boiling  continues,  the  tempera- 
ture of  the  pure  water  remains  constant  whereas 
the  temperature  of  the  salt  solution  rises.  As  the 
boiling  point  goes  up,  the  remaining  liquid  be- 
comes saltier.  If  we  collect  the  steam  from  the 
salt  solution  and  condense  it  in  a  separate  vessel, 
we  find  that  the  resulting  liquid  behaves  like  pure 
water  rather  than  like  the  solution  from  which  it 
came.  If  we  boil  off  all  the  water,  solid  salt 
remains  behind.  Thus,  by  distilling — that  is, 
evaporating  and  recondensing  in  a  separate  vessel — 
we  can  separate  a  pure  liquid  from  a  solution; 
and,  by  crystallizing — that  is,  forming  a  crystal- 
line solid — we  can  obtain  a  pure  solid  from  a 
solution.  Chemists  call  the  pure  liquid  obtained 
by  distilling  and  the  pure  solid  obtained  by  crys- 
tallizing, the  components  of  the  solution.  In  our 


SEC.    5-2    I    SOLUTIONS 


71 


Condenser 


Cold  water  inlet- 


Pure 
■water 


Fig.  5-5.  A  simple  distillation  apparatus. 

experiment  shown  in  Figure  5-4,  the  components 
are  salt  and  water. 

Pure  sodium  chloride,  like  pure  water,  has  a 
definite  melting  (freezing)  temperature  (at  a 
given  pressure).  Separating  operations — such  as 
distilling  or  freezing — do  not  separate  the  salt 
into  components.  The  composition  of  the  salt, 
whether  expressed  in  relative  numbers  of  sodium 
and  chlorine  atoms  or  in  the  relative  weights  of 
these  atoms,  is  fixed  and  is  represented  by  the 
formula  NaCl.  Sodium  chloride,  like  water,  is  an 
example  of  a  pure  substance. 

On  the  other  hand,  operations  such  as  distill- 
ing or  freezing  usually  tend  to  separate  solutions 
into  the  pure  substances  that  were  the  compo- 
nents of  the  solution.  The  nearer  alike  the  com- 
ponents are,  the  harder  it  is  to  separate  them 
from  the  solution,  but  even  in  difficult  cases,  a 
variety  of  methods  in  succession  usually  brings 
about  a  separation.  In  nature,  solutions  are  much 
more  common  than  pure  substances,  and  hetero- 
geneous systems  are  more  common  than  solu- 
tions. When  we  want  pure  substances,  we  often 
must  prepare  them  from  solutions  through  suc- 
cessive phase  changes. 

We  are  all  familiar  with  liquid  solutions.  Gas 
and  solid  solutions  also  exist.  We  shall  consider 
them  briefly  and  then  return  to  liquid  solutions, 
the  most  important  from  a  chemist's  point  of 
view. 


5-2.2    Gaseous  Solutions 

All  gas  mixtures  are  homogeneous;  hence  all  gas 
mixtures  are  solutions.  Air  is  an  example.  There 
is  only  one  phase — the  gas  phase — and  all  the 
molecules,  regardless  of  the  source,  behave  as 
gas  molecules.  The  molecules  themselves  may 
have  come  from  gaseous  substances,  liquid  sub- 
stances, or  solid  substances.  Whatever  the  source 
of  the  constituents,  this  gaseous  solution,  air,  is 
a  single,  homogeneous  phase.  As  with  other 
solutions,  the  constituents  of  air  are  separated  by 
phase  changes. 

5-2.3    Solid  Solutions 

Solid  solutions  are  more  rare.  Crystals  are  stable 
because  of  the  regularity  of  the  positioning  of 
the  atoms.  A  foreign  atom  interferes  with  this 
regularity  and  hence  with  the  crystal  stability. 
Therefore,  as  a  crystal  forms,  it  tends  to  exclude 
foreign  atoms.  That  is  why  crystallization  pro- 
vides a  good  method  for  purification. 

But  in  metals  it  is  relatively  common  for  solid 
solutions  to  form.  The  atoms  of  one  element  may 
enter  the  crystal  of  another  element  if  their  atoms 
are  of  similar  size.  Gold  and  copper  form  such 
solid  solutions.  The  gold  atoms  can  replace  cop- 
per atoms  in  the  copper  crystal  and,  in  the  same 
way,  copper  atoms  can  replace  gold  atoms  in  the 
gold  crystal.  Such  solid  solutions  are  called  al- 
loys. Some  solid  metals  dissolve  hydrogen  or 
carbon  atoms — steel  is  iron  containing  a  small 
amount  of  dissolved  carbon. 

Solid  solutions,  alloys  in  particular,  will  be 
considered  again  in  Chapter  17. 

5-2.4    Liquid  Solutions 

In  your  laboratory  work  you  will  deal  mostly 
with  liquid  solutions.  Liquid  solutions  can  be 
made  by  mixing  two  liquids  (for  example,  alcohol 
and  water),  by  dissolving  a  gas  in  a  liquid  (for 
example,  carbon  dioxide  and  water),  or  by  dis- 
solving a  solid  in  a  liquid  (for  example,  sugar  and 
water).  The  result  is  a  homogeneous  system  con- 
taining more  than  one  substance — a  solution. 
In  such  a  liquid,  each  component  is  diluted  by 
the  other  component.  In  salt  water,  the  salt 


72 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


dilutes  the  water  and,  of  course,  the  water  dilutes 
the  salt.  This  solution  is  only  partly  made  up  of 
water  molecules  and  it  is  found  that  the  vapor 
pressure  of  the  solution  is  correspondingly  lower 
than  the  vapor  pressure  of  pure  water.  Whereas 
water  must  be  heated  to  100°C  to  raise  the  vapor 
pressure  to  760  mm,  it  is  necessary  to  heat  a  salt 
solution  above  100°C  to  reach  this  vapor  pres- 
sure. Therefore,  the  boiling  point  of  salt  water  is 
above  the  boiling  point  of  pure  water.  The 
amount  the  boiling  point  is  raised  depends  upon 
the  relative  amounts  of  water  and  salt.  The  more 
salt  that  is  added,  the  higher  is  the  boiling  point. 

In  a  similar  way,  a  lower  temperature  is  re- 
quired to  crystallize  ice  from  salt  water  or  from 
an  alcohol-water  solution  than  from  pure  water. 
"Antifreeze"  substances  added  to  an  automobile 
radiator  act  on  this  principle.  They  dilute  the 
water  in  the  radiator  and  lower  the  temperature 
at  which  ice  can  crystallize  from  the  solution. 
Again,  the  amount  the  freezing  temperature  is 
lowered  depends  upon  the  relative  amounts  of 
water  and  antifreeze  compound. 

In  general,  the  properties  of  a  solution  depend 
upon  the  relative  amounts  of  the  components. 
It  is  important  to  be  able  to  specify  quantita- 
tively what  is  present  in  a  solution,  that  is,  to 
specify  its  composition.  There  are  many  ways  to 
do  this,  but  one  method  will  suffice  for  our 
purposes. 


for  various  purposes.  We  shall  use  only  one  in 
this  course. 

Chemists  often  indicate  the  concentration  of 
a  substance  in  water  solution  in  terms  of  the 
number  of  moles  of  the  substance  dissolved  per 
liter  of  solution.  This  is  called  the  molar  con- 
centration. A  one-molar  solution  (1  M)  contains 
one  mole  of  the  solute  per  liter  of  total  solution, 
a  two-molar  solution  (2  M)  contains  two  moles 
of  solute  per  liter,  and  a  0.1 -molar  solution 
(0.1  M)  contains  one-tenth  mole  of  solute  per 
liter.  Notice  that  the  concentration  of  water  is 
not  specified,  though  we  must  add  definite 
amounts  of  water  to  make  the  solutions. 

We  can  make  a  1  M  solution  of  sodium  chlo- 
ride by  weighing  out  one  mole  of  the  salt.  From 
the  formula,  NaCl,  we  know  that  one  mole 
weighs  58.5  grams  (23.0  grams  +  35.5  grams). 
We  dissolve  this  salt  in  some  water  in  a  1  liter 
volumetric  flask — a  flask  holding  just  1  liter 
when  filled  exactly  to  an  etched  mark.  After  the 
salt  dissolves,  more  water  is  added  until  the  water 
level  reaches  the  etched  mark  to  make  the  vol- 
ume exactly  1  liter.  Equally  well,  we  can  prepare 
a  1  M  sodium  chloride  solution  using  a  100  ml 
volumetric  flask.  Then  the  final  volume  of  the 
solution  will  be  0.100  liter  and  we  need  only  one- 
tenth  mole  of  salt.  In  this  case,  we  weigh  out 
5.85  grams  of  salt,  place  it  in  the  flask,  dissolve 
it,  and  add  water  to  the  100  ml  mark. 


5-2.5    Expressing  the  Composition  off  Solutions         5-2.6    Solubility 


The  components  of  a  solution  are  the  pure  sub- 
stances that  are  mixed  to  form  the  solution.  If 
there  are  two  components,  one  is  sometimes 
called  the  solvent  and  the  other  the  solute.  These 
are  merely  terms  of  convenience.  Since  both  must 
intermingle  to  form  the  final  solution,  we  cannot 
make  any  important  distinction  between  them. 
When  chemists  make  a  liquid  solution  from  a 
pure  liquid  and  a  solid,  they  usually  call  the 
liquid  component  the  solvent. 

To  indicate  the  composition  of  a  particular 
solution  we  must  show  the  relative  amounts  as 
well  as  the  kind  of  componenis.  These  relative 
amounts  chemists  call  concentrations.  Chemists 
use  different  ways  of  expressing  concentration 


When  solid  is  added  to  a  liquid,  solid  begins  to 
dissolve  and  the  concentration  of  dissolved  ma- 
terial begins  to  rise.  After  all  of  the  solid  has 
dissolved,  the  concentration  remains  constant, 
fixed  by  the  amount  of  dissolved  solid  and  the 
volume  of  the  solution  If  more  solid  is  now 
added,  the  concentration  will  rise  further.  Fi- 
nally, however,  the  addition  of  more  solid  no 
longer  raises  the  concentration  of  dissolved  ma- 
terial. When  a  fixed  amount  of  liquid  has  dis- 
solved all  of  the  solid  that  it  can,  the  concentra- 
tion reached  is  called  the  solubility  of  that  solid. 
A  solution  in  contact  with  excess  solid  is  said  to 
be  saturated. 
The  solubilities  of  solids  in  liquids  vary  widely. 


SEC.    5-2    I    SOLUTIONS 


73 


For  example,  sodium  chloride  continues  to  dis- 
solve in  water  at  20°C  until  the  concentration 
is  about  six  moles  per  liter.  The  solubility  of 
NaCl  in  water  is  6  M  at  20°C.  In  contrast,  only 
a  minute  amount  of  sodium  chloride  dissolves  in 
ethyl  alcohol  at  20°C.  This  solubility  is  0.009  M. 
Even  in  a  single  liquid,  solubilities  differ  over 
wide  limits.  The  solids  calcium  chloride,  CaCl2, 
and  silver  nitrate,  AgN03,  have  solubilities  in 
water  exceeding  one  mole  per  liter.  The  solid 
called  silver  chloride,  AgCl,  has  a  solubility  in 
water  of  only  10-5  mole  per  liter. 

Because  of  this  range  of  solubilities,  the  word 
soluble  does  not  have  a  precise  meaning.  There 
is  an  upper  limit  to  the  solubility  of  even  the 
most  soluble  solid,  and  even  the  least  soluble 
solid  furnishes  a  few  dissolved  particles  per  liter 
of  solution.  If  a  compound  has  a  solubility  of 
more  than  one-tenth  mole  per  liter  (0.1  M), 
chemists  usually  say  it  is  soluble.  When  the  solu- 
bility lies  below  0.1  M  (10-1  M),  chemists  usu- 
ally say  the  compound  is  slightly  soluble.  Com- 
pounds with  solubility  below  about  10-3  M  are 
sometimes  said  to  be  very  slightly  soluble,  and 
if  the  solubility  is  so  low  as  to  be  of  no  interest, 
the  solid  is  said  to  have  negligible  solubility.  We 
use  glass  containers  for  pure  water  because  glass 
has  a  negligible  solubility  in  water. 

5-2.7    Variations  Among  the  Properties 
off  Solutions 

Though  many  solutions  are  colorless  and  closely 
resemble  pure  water  in  appearance,  the  differ- 
ences among  solutions  are  great.  This  can  be 
demonstrated  with  the  five  pure  substances,  so- 
dium chloride  (salt),  iodine,  sugar,  ethyl  alcohol, 
and  water.  Two  of  these  substances,  ethyl  alcohol 
and  water,  are  liquids  at  room  temperature. 
Let's  investigate  the  properties  of  the  solutions 
these  two  substances  form. 

First  we  can  investigate,  qualitatively,  the  ex- 
tent to  which  the  solids  dissolve  in  the  liquids. 
By  adding  a  small  piece  of  each  solid  to  a  milli- 
liter of  liquid,  we  easily  discover  that  sugar  dis- 
solves both  in  water  and  ethyl  alcohol,  sodium 
chloride  dissolves  readily  in  water  but  not  in 
ethyl  alcohol,  and  iodine  does  not  dissolve  much 


Fig.  5-6.  Salt  water  readily  conducts  electricity;  sugar 
solution  does  not. 


in  water  but  dissolves  readily  in  ethyl  alcohol. 
Thus  we  see  that  the  solvent  properties  of  the 
two  liquids  are  quite  distinctive,  at  least  as  far 
as  sugar,  salt,  and  iodine  are  concerned. 

The  experiment  just  described  gives  us  four 
solutions  containing  a  substantial  amount  of 
solute: 


74 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


Sugar  in 
water 


II 

Sugar  in 
ethyl 
alcohol 


III 

Sodium 
chloride 
in  water 


IV 

Iodine  in 
ethyl 
alcohol 


Of  these  four  solutions,  IV  is  readily  distin- 
guished. This  solution  has  a  dark  brown  color. 
The  other  three,  I,  II,  and  III,  are  colorless.  They 
can  be  easily  distinguished  by  taste  but  chemists 
have  safer  and  more  meaningful  ways  of  distin- 
guishing them.  These  solutions  differ  markedly 
in  their  ability  to  conduct  an  electric  current.  The 
two  sugar  solutions,  I  and  II,  have  virtually  the 
same  conductivity  properties  as  the  pure  liquids 
— they  do  not  conduct  electric  current  readily. 


Solution  III  conducts  electric  current  much  more 
readily  than  does  pure  water. 

Thus  we  find  great  variation  among  solutions. 
Iodine  dissolves  in  ethyl  alcohol,  coloring  the 
liquid  brown,  but  does  not  dissolve  readily  in 
water.  Sodium  chloride  does  not  dissolve  readily 
in  ethyl  alcohol  but  does  dissolve  in  water,  form- 
ing a  solution  that  conducts  electric  current. 
Sugar  dissolves  readily  both  in  ethyl  alcohol  and 
in  water,  but  neither  solution  conducts  electric 
current.  These  differences  are  very  important  to 
the  chemist,  and  variations  in  electrical  con- 
ductivity are  among  the  most  important.  We 
shall  investigate  electrical  conductivity  further 
but,  first,  we  need  to  explore  the  electrical  nature 
of  matter. 


5-3     ELECTRICAL  NATURE  OF  MATTER 


We  have  mentioned  the  electrical  conductivity  of 
solutions  as  a  means  of  distinguishing  solutions. 
The  interest  of  a  chemist  in  the  electrical  nature 
of  matter  goes  far  deeper  than  this.  We  shall  find 
that  an  understanding  of  electrical  behavior  fur- 
nishes a  key  to  the  explanation  of  chemical  prop- 
erties. We  shall  find  that  electrical  effects  aid  us 
in  predicting  molecular  formulas,  in  explaining 
chemical  reactions,  and  in  understanding  energy 
changes  that  accompany  them. 

5-3.1    Electrical  Phenomena 

Try  to  name  several  electrical  phenomena  that 
you  have  often  observed.  Before  you  read  on, 
see  if  you  can  name  five.  Does  your  list  include 
the  following? 

(1)  The  attraction  of  a  comb  for  your  hair  on  a 
dry  day. 

(2)  The  flash  of  a  bolt  of  lightning. 

(3)  The  shock  you  get  if  you  touch  a  bare  wire 
in  a  radio  set. 

(4)  The  heat  generated  by  an  electric  current 
passing  through  the  heating  element  of  an 
electric  stove. 

(5)  The  light  emitted  by  the  filament  of  a  light 
bulb  as  electric  current  is  passed  through  it. 


(6)  The  magnetic  field  generated  by  a  current 
passing  through  a  coil  of  wire. 

(7)  The  work  done  by  an  electric  motor  when 
electric  current  passes  through  its  coils. 

(8)  The  emission  of  "radio  waves"  by  the  an- 
tenna of  a  radio  or  television  station. 

We  see  electrical  devices  all  around  us — fur- 
nishing power,  light,  means  of  communication — 
influencing  every  facet  of  life.  What  does  it  mean 
to  say  that  an  electric  current  "passes  through" 
a  coil  of  wire?  What  is  an  electric  current?  To 
begin  answering  these  questions,  we  must  explore 
an  electrometer,  a  device  for  detecting  and  meas- 
uring electric  charge. 

5-3.2    Detection  of  Electric  Charge 

Figure  5-7  shows  a  simple  electrometer.  It  con- 
sists of  two  spheres  of  very  light  weight,  each 
coated  with  a  thin  film  of  metal.  The  spheres  are 
suspended  near  each  other  by  fine  metal  threads 
in  a  closed  box  to  exclude  air  drafts.  Each  sus- 
pending thread  is  connected  to  a  brass  terminal. 
Next  to  the  box  is  a  "battery"— a  collection  of 
electrochemical  cells.  There  are  two  terminal 
posts  on  the  battery.  We  shall  call  these  posts  Pi 
and  Pi.  If  post  Pi  is  connected  by  a  copper  wire 


SEC.    5-3        ELECTRICAL    NATURE    OF    MATTER 


75 


Fig.  5-7.  A  simple  electrometer. 


to  the  left  terminal  of  the  electrometer  and  post 
Pi  is  connected  to  the  right  terminal,  we  observe 
that  the  two  spheres  move  toward  each  other. 
Evidently  the  wires  have  transmitted  to  the 
spheres  the  property  of  exerting  force  on  each 
other — an  attractive  force.  The  force  is  still  pres- 
ent when  the  air  in  the  electrometer  is  removed 
with  a  vacuum  pump.  The  spheres  react  to  each 
other  "across  space."  They  feel  "force  at  a  dis- 
tance." 

If  now  the  wires  are  disconnected,  the  attrac- 
tive force  remains.  However,  if  the  two  elec- 
trometer terminals  are  connected  by  a  copper 
wire  the  spheres  return  to  their  original  positions 
and  hang  vertically  again.  The  attraction  is  lost. 

We  see  that  the  battery  transfers  to  the  elec- 
trometer spheres  the  property  of  attracting  each 
other.  It  is  natural  to  imagine  that  something  has 
been  transferred  from  the  battery  to  the  spheres. 
This  "something"  is  called  electric  charge.  The 
movement  of  this  electric  charge  from  the  battery 
through  the  metal  threads  to  the  spheres  is  called 
an  electric  current.  This  electric  charge  is  lost 
when  the  two  electrometer  terminals  are  con- 
nected by  a  copper  wire. 

We  can  learn  more  about  electric  charge  by 
another  use  of  the  electrometer.  Connect  one 
wire  from  the  battery  (say,  post  P{)  to  the  base 
of  the  electrometer  and  the  other  wire  (from  post 
P2)  to  both  terminals,  as  in  Figure  5-8. 

This  time  the  two  spheres  move  apart — they 
repel  each  other!  When  both  spheres  are  given 
electric  charge  from  the  battery  post  labeled  P2, 
they  repel  instead  of  attract.  In  Figure  5-7  we 


saw  that  when  one  sphere  received  charge  from 
post  A  and  the  second  sphere  received  charge 
from  post  Pi,  the  two  spheres  attracted.  There 
must  be  at  least  two  kinds  of  charge! 

Now  let  us  reverse  the  wires  so  that  both 
spheres  are  charged  from  battery  post  Pi.  This 
time,  Pi  is  connected  to  the  base  of  the  electrome- 
ter. Again  we  observe  that  the  spheres  move 
apart.  Whenever  both  spheres  are  connected  to 
the  same  battery  post,  the  two  spheres  repel  each 
other. 

This  represents  a  large  gain  in  our  knowledge 
of  electric  charge.  One  kind  of  charge  comes 
from  battery  post  Pi.  Until  we  have  reason  to  do 
otherwise,  we  shall  call  this  kind  of  charge  C\. 
The  other  kind  of  charge  comes  from  battery 
post  P2;  we  shall  call  this  kind  of  charge  d.  The 
spheres  attract  or  repel  each  other  when  they 
carry  charge  according  to  the  following  pattern: 


C\  attracts  d 
C\  repels  G 
C2  repels  C2 


unlike  charges  attract 
like  charges  repel 
like  charges  repel 


(i) 


Fig.  5-8.  The   electrometer    with    both   spheres  con- 
nected to  battery  post  Pt. 


76 


LIQUIDS    AND    SOLIDS!    CONDENSED    PHASES    OF    MATTER   |    CHAP.    5 


We  have  one  more  observation  to  symbolize. 
When  the  spheres  were  given  different  charges 
(as  in  Figure  5-7),  the  charges  could  be  removed 
by  connecting  the  two  terminals  by  a  copper 
wire.  Then  the  spheres  lost  all  attraction  for  each 
other.  We  interpret  this  behavior  to  mean  that 
either  G  or  C2  (or  both)  has  moved  through  the 
wire  so  as  to  join  the  other  kind  of  charge.  When 
C\  and  C2  are  united,  no  charge  remains.  Sym- 
bolically we  can  say, 

Ci  +  Ci  =  no  charge  (4) 

Our  accumulated  evidence  shows  that  there 
are  at  least  two  kinds  of  electric  charge,  which 
we  have  symbolized  G  and  C2.  These  two  kinds 
of  charge  possess  the  properties  (3)  and  (4).  We 
may  wonder  if  there  are  other  kinds  of  charge. 
This  is  answered  by  looking  into  other  ways  of 
producing  electric  charges.  Chemists  can  assem- 
ble a  variety  of  types  of  electrochemical  cells 
which  show  the  same  electrometer  behavior 
just  described.  Some  frictional  processes  leave 
charges  on  the  two  surfaces  rubbed  together.  The 
attraction  of  a  comb  for  your  hair  is  caused  by 
charges  left  on  the  comb  as  it  rubbed  against 
your  hair.  Many  decades  ago  the  properties  of 
electric  charge  were  investigated  as  they  are  pro- 
duced by  rubbing  a  hard  rubber  rod  with  cat  fur. 
The  hard  rubber  is  found  to  carry  charge  G,  and 
the  cat  fur  carries  charge  C2.  If  a  glass  rod  is 
rubbed  with  silk,  the  glass  rod  is  left  with  charge 
C2  and  the  silk  with  charge  G. 

No  matter  how  electric  charge  is  produced,  we 
always  find  these  same  two  types,  G  and  C2,  and 


only  these  two.  Any  method  of  producing  G  also 
produces  an  equivalent  amount  of  C2.  We  con- 
clude there  are  two  and  only  two  types  of  electric 
charge. 

5-3.3    The  Effect  of  Distance 

Figure  5-9  shows  two  electrometers  which  differ 
in  the  spacing  of  the  spheres.  Though  the  charges 
on  the  spheres  come  from  the  same  battery,  there 
is  more  deflection  of  the  spheres  when  they  are 
closely  spaced  (left)  than  when  they  are  widely 
spaced.  When  the  spheres  are  closer  together,  the 
deflection  is  larger.  Hence,  we  conclude  that  the 
force  of  attraction  varies  with  distance  and  is 
stronger  when  the  charges  are  close  to  each 
other.  Careful  quantitative  studies  show  that  the 
force  is  inversely  proportional  to  the  square  of 
the  distance  r  between  the  two  spheres : 


(Electric  force)  is  proportional  to  — 


(5) 


where 


r  =  distance  between  centers  of  the  two  spheres. 


5-3.4    The  Electron-Proton  Model 

These  new  facts  about  electrical  phenomena  can 
be  incorporated  into  our  particle  model  of  the 
structure  of  matter  if  we  again  allow  some 


Fig. 


5-9.  Contrast  of  deflections  in  two  electrometers 
with  different  distances  between  spheres. 


SEC.    5-3    I    ELECTRICAL    NATURE    OF    MATTER 


77 


growth  of  the  model.  The  new  idea  is  that  matter 
is  made  up  of  particles  which  carry  the  property 
called  electric  charge.  To  be  specific,  we  propose 
that  in  atoms  there  are  two  kinds  of  particles 
that  carry  unit  charge,  one  which  carries  one 
portion,  or  unit,  of  charge  G  and  one  which 
carries  one  portion,  or  unit,  of  charge  C2.  These 
particles  are  called  electrons  and  protons. 

Proposal:  Matter    includes    particles,    each    of 

which    carries    a    unit    of  electric 

charge. 
Electrons:  Each  electron  carries  one 

unit  of  charge  G- 
Protons:  Each  proton  carries  one  unit 

of  charge  G- 

These  particles  exert  force  at  a  distance  on 
each  other  in  accordance  with  the  electrical  be- 
havior we  have  observed. 

Since: 

G  repels  G,  electrons  repel  electrons; 

G  repels  G,  protons  repel  protons ; 

G  attracts  C2,  electrons  attract  protons; 

G  +  C2  =  no  charge,    one  electron  +  one  proton 

=  no  charge; 
or,  one  unit  G  +  one  unit  C, 

=  no  charge. 

The  atomic  model  now  can  cope  with  the  facts 
we  have  learned  about  electrical  behavior.  If  a 
piece  of  matter  (such  as  one  of  the  electrometer 
spheres)  has  the  same  number  of  electrons  and 
protons,  there  are  just  as  many  units  of  charge 
of  type  G  as  of  type  C2.  Since  G  +  C2  =  no 
charge,  the  sphere  will  have  no  charge.  A  body 
with  no  net  charge  (with  equal  numbers  of  pro- 
tons and  electrons)  is  said  to  be  electrically 
neutral.  If  we  remove  some  of  the  electrons  from 
the  sphere,  it  will  then  have  an  excess  of  protons, 
hence  a  net  charge  of  type  C2.  If  we  add  an  excess 
of  electrons  to  the  sphere,  it  will  have  a  net 
charge  of  type  Q.  The  amount  of  net  charge  is  the 
difference  between  the  amount  of  charge  G  and 
charge  C2. 

It  is  a  mathematical  convenience  if  we  express 
the  net  charge  in  terms  of  algebraic  symbols. 
Henceforth  we  shall  identify  the  type  of  charge 
called  G  as  "negative  charge"  and  the  type 


called  G  as  "positive  charge."  Notice  the  ad- 
vantages. 

The  combination  of  5  units  of  G  and  3  units 
of  G  leaves  a  net  of  2  units  of  G-  This  now  can 
be  expressed 

5(-l)  +  3(  +  l)  =  -5  +  3  =  -2 


EXERCISE  5-6 


Suppose  ten  protons  and  eleven  electrons  are 
brought  together.  These  charges,  grouped  to- 
gether, have  the  same  net  charge  as  how  many 
electrons?  Remember  that  one  proton  plus  one 
electron  gives  no  charge. 


EXERCISE  5-7 


Write  an  algebraic  expression  to  obtain  the  re- 
sult of  Exercise  5-6,  using  numbers  with  algebraic 
signs  to  represent  charges. 


5-3.5     Electric  Force:  A  Fundamental 
Property  of  Matter 

We  have  learned  that  a  battery  can  transfer  to 
the  spheres  of  an  electrometer  a  property  called 
electric  charge.  When  this  happens,  the  spheres 
exert  force  on  each  other.  The  discussion  brings 
up  two  "wondering  why"  questions.  The  first  is, 
"Why  do  the  electric  charges  appear?"  What 
caused  the  battery  to  transfer  to  the  electrometer 
the  property  called  electric  charge?  We  shall 
examine  this  question  carefully  later  in  the  course 
because  the  subject  of  the  operation  of  an  elec- 
trochemical cell  is  extremely  important  in  chem- 
istry. It  is  the  topic  of  an  entire  chapter  in  this 
book  (Chapter  12).  For  the  moment,  all  we  can 
say  is  that  the  electric  charge  did  come  from  the 
battery  of  electrochemical  cells,  thus  indicating 
that  the  matter  within  the  cells  contains  electric 
charges. 

The  second  question  probes  deeper:  "Why  do 
the  two  electrometer  spheres,  when  charged,  ex- 
ert force  on  each  other?"  What  is  our  explana- 
tion of  this  phenomenon?  We  say  that  the 
spheres  have  an  excess  of  electrons  (or  protons) 
and  these  electrons  (or  protons)  exert  force  on 


78 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


each  other.  This  does  not  really  explain  electric 
force  at  a  distance.  We  are  left  with  the  equiva- 
lent questions,  "Why  do  two  electrons  (or  two 
protons)  repel  each  other?  Why  does  an  electron 
attract  a  proton?"  Without  an  answer,  we  say, 
"It  is  a  fundamental  property  of  matter  that  it 
can  acquire  electric  charge  and,  when  it  does  so, 
it  exerts  force  on  other  charged  bodies."  Such  a 
statement  may  be  taken  as  a  definition  of  a 
"fundamental  property" — a  property  which  is 
generally  observed  but  for  which  diligent  search 
has  failed  to  yield  a  useful  model.  Without  an 
explanation  of  a  property,  we  call  the  property 
"fundamental."  It  is  a  curious  fact  that  after  a 
property  has  resisted  explanation  for  quite  a  time 
and  it  becomes  classified  as  a  fundamental  prop- 


erty, an  explanation  no  longer  seems  to  be  neces- 
sary. 

EXERCISE  5-8 

There  was  a  time  when  atoms  were  said  to  be 
fundamental  particles  of  which  matter  is  com- 
posed. Now  we  describe  the  structure  of  the 
atom  in  terms  of  the  fundamental  particles  we 
have  just  named,  protons  and  electrons,  plus 
another  kind  of  particle  called  a  neutron.  Why 
are  atoms  no  longer  said  to  be  fundamental  par- 
ticles? Do  you  expect  neutrons,  protons,  and 
electrons  always  to  be  called  fundamental  par- 
ticles? 


5-4     ELECTRICAL  PROPERTIES  OF  CONDENSED  PHASES 


Now  we  are  ready  to  investigate  behavior  of 
condensed  phases  that  shows  evidence  of  the 
presence  and  movement  of  electric  charge.  We 
have  already  referred  to  one  of  the  most  impor- 
tant examples — the  movement  of  electric  current 
through  water  solutions. 

5-4.1    The  Electrical  Conductivity 
off  Water  Solutions 

The  movement  of  electric  charge  is  called  an 
electric  current.  Hence  when  we  say  electric  cur- 
rent flows  through  a  salt  solution,  we  mean  there 
is  a  movement  of  electric  charge  through  the 
solution.  We  shall  be  concerned  here  with  the 
manner  in  which  this  charge  moves. 

Water  is  a  very  poor  conductor  of  electricity. 
Yet  when  sodium  chloride  dissolves  in  water,  the 
solution  conducts  readily.  The  dissolved  sodium 
chloride  must  be  responsible.  How  does  the  dis- 
solved salt  permit  charge  to  move  through  the 
liquid?  One  possibility  is  that  when  salt  dissolves 
in  water,  particles  with  electric  charge  are  pro- 
duced. The  movement  of  these  charged  particles 
through  the  solution  accounts  for  the  current. 
Salt  has  the  formula,  NaCl — for  every  sodium 
atom  there  is  one  chlorine  atom.  Chemists  have 


decided  that  when  sodium  chloride  dissolves  in 
water,  the  charged  species  present  are  chlorine 
atoms,  each  carrying  the  negative  charge  of  one 
electron,  and  sodium  atoms,  each  carrying  the 
positive  charge  of  one  proton.  We  symbolize  a 
chlorine  atom  with  a  negative  charge  as  Cl~. 
A  sodium  atom  with  a  positive  charge  is  sym- 
bolized Na+.  Atoms  or  molecules  that  carry  elec- 
tric charge  are  called  ions. 

With  these  symbols,  we  can  write  the  equation 
for  the  reaction  that  occurs  when  sodium  chlo- 
ride dissolves  in  water 

NaCl(solid)  +  water  — ►- 

Na^(in  water)  +  Cl~(in  water)    (6) 

Equation  (6)  shows  that  when  water  dissolves 
solid  sodium  chloride,  Na+  ions  (sodium  ions) 
and  Cl~  ions  (chloride  ions)  are  present  in  the 
solution.  Chemists  usually  abbreviate  this  equa- 
tion as  much  as  is  consistent  with  retaining  es- 
sential information.  On  the  left  of  equation  (<5), 
the  term  "water"  is  usually  not  written  since  its 
presence  is  implied  by  the  symbols  on  the  right. 

NaCl(solid)  — >■ 

Na+(in  water)  +  Cl_(in  water)    (6) 

We  have  already  seen  that  NaCl(solid)  is  usually 
written  NaClfsj.  There  is  a  similar  abbreviation 


SEC.    5-4    |    ELECTRICAL    PROPERTIES    OF    CONDENSED    PHASES 


79 


for  "in  water."  To  represent  this,  the  expression 
"in  water"  is  replaced  by  the  term  "aqueous," 
commonly  abbreviated  "aq".  *  Thus  equation 
(6),  showing  the  reaction  of  sodium  chloride  dis- 
solving in  water  to  form  a  conducting  solution, 
is  usually  written  in  the  form: 


NaClfsj  — ►-  Na+(aq)  +  C\~(aq) 


(6) 


Now  we  have  a  model  of  a  salt  solution  that 
aids  us  in  discussing  electrical  conductivity.  The 
solid  dissolves,  forming  the  charged  particles 
Na.+(aq)  and  C\~(aq),  which  can  move  about 
in  the  solution  independently.  An  electric  current 
can  pass  through  the  solution  by  means  of  the 
movement  of  these  ions.  The  C\~(aq)  ions  move 
in  one  direction,  causing  negative  charge  to  move 
that  way.  The  Na+(aq)  ions  move  in  the  oppo- 
site direction,  causing  positive  charge  to  move 
this  way.  These  movements  carry  charge  through 
the  solution  and  current  flows. 

Sugar  dissolves  in  water,  but  the  resulting  solu- 
tion conducts  electric  current  no  better  than  does 
pure  water.  We  conclude  that  when  sugar  dis- 
solves, no  charged  particles  result:  no  ions  are 
formed.  Sugar  must  be  quite  different  from  so- 
dium chloride. 

Calcium  chloride,  CaCl2,  is  another  crystalline 
solid  that  dissolves  readily  in  water.  The  resulting 
solution  conducts  electric  current,  as  does  the 
sodium  chloride  solution.  Calcium  chloride  is,  in 
this  regard,  like  sodium  chloride  and  unlike 
sugar.  The  equation  for  the  reaction  is 


CaCVsj 


Ca+*(aq)  +  2C\-(aq)         (7) 


Equation  (7)  shows  that  when  calcium  chlo- 
ride dissolves,  ions  are  present — Ca+2(aq)  and 
C\~(aq)  ions.  In  this  case,  each  calcium  ion  has 
the  positive  charge  of  two  protons.  Therefore  it 
has  twice  the  positive  charge  held  by  a  sodium 
ion,  Na.+(aq).  The  chloride  ion  that  forms, 
C\~(aq),  is  the  same  negative  ion  that  is  present 
in  the  sodium  chloride  solution,  though  it  comes 
from  the  calcium  chloride  solid  instead  of  the 
sodium  chloride.   Because  both  CaCh(s)  and 


NaClfsJ  dissolve  in  water  to  form  aqueous  ions, 
they  are  considered  to  be  similar. 

Silver  nitrate,  AgN03,  is  a  third  solid  sub- 
stance that  dissolves  in  water  to  give  a  conduct- 
ing solution.  The  reaction  is 

AgN03(s)  — >-  Ag+(aq)  +  NOs-  (aq)        (8) 

This  time  the  ions  formed  are  silver  ions, 
Ag+(aq),  and  nitrate  ions,  N03~(aqj.  The  aque- 
ous silver  ion  is  a  silver  atom  with  the  positive 
charge  of  a  proton ;  it  carries  the  same  charge  as 
does  an  aqueous  sodium  ion.  The  aqueous  ni- 
trate ion  carries  the  negative  charge  of  an  elec- 
tron— the  same  charge  carried  by  the  aqueous 
chloride  ion.  This  time,  however,  the  negative 
charge  is  carried  by  four  atoms,  a  nitrogen  and 
three  oxygen  atoms,  that  remain  together.  Since 
this  group,  N03~,  remains  together  and  acts  as 
a  unit,  it  has  a  distinctive  name,  nitrate  ion. 

These  three  solids,  sodium  chloride,  calcium 
chloride,  and  silver  nitrate  are  similar,  hence 
they  are  classified  together.  They  all  dissolve  in 
water  to  form  aqueous  ions  and  give  conducting 
solutions.  These  solids  are  called  ionic  solids. 

The  ease  with  which  an  aqueous  salt  solution 
conducts  electric  current  is  determined  by  how 
much  salt  is  dissolved  in  the  water,  as  well  as  by 
the  fact  that  ions  are  formed.  A  solution  con- 
taining 0.1  moles  per  liter  conducts  much  more 
readily  than  a  solution  containing  0.01  moles  per 
liter.  Thus  the  conductivity  is  determined  by  the 
concentration  of  ions,  as  well  as  by  their  pres- 
ence. 

Silver  chloride  is  a  solid  that  shows  this  effect. 
This  solid  does  not  dissolve  readily  in  water. 
When  solid  silver  chloride  is  placed  in  water, 
very  little  solid  enters  the  solution  and  there  is 
only  a  very  slight  increase  in  the  conductivity  of 
the  solution.  Yet  there  is  a  real  and  measurable 
increase — ions  are  formed.  Careful  measure- 
ments show  that  even  though  silver  chloride  is 
much  less  soluble  in  water  than  sodium  chloride, 
it  is  like  sodium  chloride  in  that  all  the  solid  that 
does  dissolve  forms  aqueous  ions.  The  reaction  is 


AgC\(s) 


Ag+(aq)  +  Cl-(aq) 


(9) 


*  The  adjective,  aqueous,  comes  from  the  Latin  name 
for  water,  aqua. 


Silver  chloride,  like  sodium  chloride,  is  an  ionic 
solid. 


80 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


5-4.2     Precipitation  Reactions 
in  Aqueous  Solutions 

Though  both  silver  nitrate  and  sodium  chloride 
have  high  solubility  in  water,  silver  chloride  is 
very  slightly  soluble.  What  will  happen  if  we  mix 
a  solution  of  silver  nitrate  and  sodium  chloride? 
Then,  we  will  have  a  solution  that  includes  the 
species  present  in  a  solution  of  silver  chloride, 
Ag+(aq)  and  C\~(aq),  but  now  they  are  present 
at  high  concentration!  The  Ag+(  aq)  came  from 
reaction  (8)  and  the  C\~(aq)  came  from  reaction 
(6)  and  their  concentrations  far  exceed  the  solu- 
bility of  silver  chloride.  The  result  is  that  solid 
will  be  formed.  The  formation  of  solid  from  a 
solution  is  called  precipitation: 

Ag+(aq)  +  C\-(aq)  — >  AgClfsJ  (70) 

Notice  that  reaction  (70)  indicates  the  change 
that  takes  place  when  silver  nitrate  solutions  and 
sodium  chloride  solutions  are  mixed.  We  could 
have  written  a  more  complete  equation: 

Ag+(aq)  +  NOz(aq)  +  Na+(aq)  +  C\~(aq) 

—^  AgC\(s)  +  NtV  (aq)  +  Na+(a<jj 

However,  the  two  ions  N03~fagj  and  Na+(aq) 
do  not  play  an  active  role  in  the  reaction,  nor  do 
they  influence  the  reaction  that  does  occur  [re- 
action (70)].  Consequently,  they  are  not  included 
in  the  equation  for  the  reaction.  The  balanced 
chemical  equation  should  show  only  species  which 
actually  participate  in  the  reaction.  These  species 
are  called  the  predominant  reacting  species. 

Equations  (6),  (7),  (8),  (9),  and  (70)  involve 
charged  species,  ions.  When  we  considered  how 
to  balance  equations  for  chemical  reactions  (Sec- 
tion 3-2.1),  we  dealt  with  reactions  involving  elec- 
trically neutral  particles.  We  were  guided  by  the 
rule  that  atoms  are  conserved.  This  principle  is 
still  applicable  to  reactions  involving  ions.  In 
addition,  we  must  consider  the  charge  balance. 
A  chemical  reaction  does  not  produce  or  con- 
sume electric  charge.  Consequently,  the  sum  of 
the  electric  charges  among  the  reactants  must  be 
the  same  as  the  sum  of  the  electric  charges  among 
the  products.  In  reaction  (7),  calcium  chloride 
dissolves  to  give  aqueous  Ca+2  and  Cl~  ions.  The 
balanced  equation  tells  us  that  the  neutral  solid 
calcium  chloride  dissolves  to  give  one  Ca^2  ion 


for  every  two  CI-  ions.  Summing  these  electric 
charges, 

["charge  on  "1  __  fcharge  on~|    ,   9  ["charge  on~| 
|_CaCl2  solidj       |_Ca~2  ion  J  ~*~  4  |_C1-  ion 


(2+)        +2        (1-) 

0  (77) 


In  a  balanced  equation  for  a  chemical  reaction, 
charge  is  conserved. 


EXERCISE  5-9 

Balance  the  equations  for  the  reactions  given 
below.  For  each  balanced  equation,  sum  up  the 
charges  of  the  reactants  and  compare  to  the  sum 
of  the  charges  of  the  products. 

(a)  PbCUfs)  — 

(b)  K2Cr207fsJ 


?b+*(aq)  +  C\-(aq) 


K+(aq)  +  Cr2Or(aq) 


(c)  Cr20;'(aq)+H20 


CTO;2(aq)  +  H+(aq) 


5-4.3    The  Electrical  Conductivity  of  Solids 

We  have,  in  this  chapter,  encountered  a  number 
of  properties  of  solids.  In  Table  5-II,  we  found 
that  melting  points  and  heats  of  melting  of  dif- 
ferent solids  vary  widely.  To  melt  a  mole  of  solid 
neon  requires  only  80  calories  of  heat,  whereas 
a  mole  of  solid  copper  requires  over  3000  calo- 
ries. Some  solids  dissolve  in  water  to  form  con- 
ducting solutions  (as  does  sodium  chloride), 
others  dissolve  in  water  but  no  conductivity 
results  (as  with  sugar).  Some  solids  dissolve  in 
ethyl  alcohol  but  not  in  water  (iodine,  for  ex- 
ample). Solids  also  range  in  appearance.  There 
is  little  resemblance  between  a  transparent  piece 
of  glass  and  a  lustrous  piece  of  aluminum  foil, 
nor  between  a  lump  of  coal  and  a  clear  crystal 
of  sodium  chloride. 

The  great  variations  among  solids  make  it 
desirable  to  find  useful  classification  schemes. 
Though  this  topic  is  taken  up  much  later  in  the 
course  (Chapter  17),  a  beginning  is  provided  by 
a  look  at  the  electrical  conductivity  of  solids. 

The  high  electrical  conductivity  of  a  substance 
like  copper  or  silver  is  familiar  to  all.  Conduc- 


SEC.    5-4    I    ELECTRICAL    PROPERTIES    OF    CONDENSED    PHASES 


tivity  measurements  on  many  other  solids  show 
that  all  of  the  substances  that  conduct  electricity 
as  readily  as  copper  and  silver  are  of  similar 
appearance.  These  good  conductors  could  al- 
most all  be  classified  visually  as  metals.  The  most 
distinctive  property  of  metallic  substances  is  high 
electrical  conductivity. 

When  we  study  a  solid  that  does  not  have  the 
characteristic  lustrous  appearance  of  a  metal,  we 
find  that  the  conductivity  is  extremely  low.  This 
includes  the  solids  we  have  called  ionic  solids: 
sodium  chloride,  sodium  nitrate,  silver  nitrate, 
and  silver  chloride.  It  includes,  as  well,  the 
molecular  crystals,  such  as  ice.  This  solid,  shown 
in  Figure  5-3,  is  made  up  of  molecules  (such  as 
exist  in  the  gas  phase)  regularly  packed  in  an 
orderly  array.  These  poor  conductors  differ 
widely  from  the  metals  in  almost  every  property. 
Thus  electrical  conductivity  furnishes  the  key  to 
one  of  the  most  fundamental  classification 
schemes  for  substances. 

5-4.4     Ionic  Solids 

Referring  to  Tables  5-1  and  5-II,  we  find  that 
both  sodium  chloride  and  copper  have  extremely 
high  melting  and  boiling  points.  These  two  solids 
have  little  else  in  common.  Sodium  chloride  has 
none  of  the  other  properties  that  identify  a  metal. 
It  has  no  luster,  rather,  it  forms  a  transparent 
crystal.  It  does  not  conduct  electricity  nor  is  it  a 
good  heat  conductor.  The  kind  of  forces  holding 
this  crystal  together  must  be  quite  different  from 
those  in  metals. 

The  sodium  chloride  crystal  contains  an  equal 
number  of  sodium  atoms  and  chlorine  atoms, 


but  they  are  not  present  as  molecules.  On  the 
basis  of  much  experimental  evidence,  chemists 
have  concluded  that  sodium  chloride  crystals  are 
built  up  of  sodium  ions,  Na+,  and  chloride  ions, 
Cl~,  rather  than  of  neutral  atoms  or  molecules. 
The  numbers  of  Na+  and  Cl~  ions  must  be  equal 
because  the  entire  crystal  is  electrically  neutral. 
Nevertheless,  there  is  electrical  attraction  be- 
tween these  oppositely  charged  particles.  This 
attraction  between  positive  and  negative  ions 
accounts  for  the  binding  in  an  ionic  solid. 

To  represent  the  composition  of  such  a  solid, 
we  use  the  formula  NaCl.  However,  this  formula 
does  not  indicate  that  molecules  of  sodium  chlo- 
ride are  present — it  is  not  a  molecular  formula. 
It  shows  the  composition  by  giving  the  relative 
number  of  each  kind  of  atom  present  and  is 
called  an  empirical  formula. 

Figure  5-10  shows  a  representation  of  the  ar- 
rangement of  the  ions  in  the  sodium  chloride 
crystal.  The  ions  are  arranged  in  layers.  A  layer 
in  the  interior  of  a  crystal  has  a  similar  layer 
lying  in  front  of  it  and  a  similar  layer  lying 
behind  it.  These  layers  are  displaced  so  that  a 
Cl~  ion  lies  in  front  of  each  Na+  ion  and  a  Cl~ 
ion  lies  behind  each  Na+  ion.  Thus,  each  ion  is 
surrounded  by  six  oppositely  charged  ions.  We 
call  this  arrangement  the  sodium  chloride  ar- 
rangement or  sodium  chloride  lattice.  Because  of 
the  proximity  of  the  oppositely  charged  ions  in 
this  arrangement,  it  is  strongly  bonded  and  the 
melting  point  of  such  a  crystal  is  high. 

Fig.  5-10.  The   packing  of  ions  in  an   ionic  crystal: 

sodium  chloride. 


82 


LIQUIDS    AND    SOLIDS:     CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


When  an  ionic  solid  like  sodium  chloride  is 
melted,  the  molten  salt  conducts  electric  current. 
The  conductivity  is  like  that  of  an  aqueous  salt 
solution:  Na+  and  C\~  ions  are  present.  The  ex- 
tremely high  melting  temperature  (808°C)  shows 
that  a  large  amount  of  energy  is  needed  to  tear 
apart  the  regular  NaCl  crystalline  arrangement 
to  free  the  ions  so  they  can  move. 

In  contrast,  solid  sodium  chloride  dissolves 
readily  in  water  at  room  temperature  and  with- 
out a  large  heat  effect.  This  can  only  mean  that 
the  water  interacts  strongly  with  the  ions — so 
strongly  that  aqueous  ions  are  about  as  stable 
as  are  ions  in  the  crystal.  In  fact,  water  interacts 


so  strongly  with  ions  that  some  molecular  crys- 
tals dissolve  in  water  to  form  conducting  solu- 
tions. For  example,  solid  hydrogen  chloride, 
HC\(s),  is  a  molecular  crystal  similar  to  the  ice 
crystal.  The  solid  is  made  up  of  HC1  molecules, 
not  of  ions  like  the  ionic  sodium  chloride.  Yet 
UC\(s)  dissolves  in  water  to  form  a  conducting 
solution  containing  hydrogen  ions,  H+(aq),  and 
chloride  ions,  Cl~(aq).  Thus  we  cannot  safely 
interpret  the  conductivity  of  an  aqueous  solution 
to  mean  that  the  solid  dissolved  was  an  ionic 
solid.  We  can,  however,  state  the  opposite:  When 
an  ionic  solid  dissolves  in  water,  a  conducting 
solution  is  obtained. 


QUESTIONS  AND  PROBLEMS 


1.  A  liquid  is  heated  at  its  boiling  point.  Although 
energy  is  used  to  heat  the  liquid,  its  temperature 
does  not  rise.  Explain. 

2.  What  is  the  maximum  amount  of  heat  that  you 
can  lose  as  one  gram  of  water  evaporates  from 
your  skin? 

3.  Note  in  Table  5-1  the  correlation  between  the 
normal  boiling  point  and  heat  of  vaporization 
of  a  number  of  liquids.  Suggest  possible  reasons 
for  this  regularity. 

4.  Which  would  likely  cause  the  more  severe  burn, 
one  gram  of  H20(g)  at  100°C  or  one  gram  of 
H,0(7j  at  100°C? 

5.  Liquids  used  in  rocket  fuels  are  passed  over  the 
outer  wall  of  the  combustion  chamber  before 
being  fed  into  the  chamber  itself.  What  advan- 
tages does  this  system  offer? 

6.  Which  of  the  following  will  require  more  energy  ? 

(a)  Changing  a  mole  of  liquid  water  into  gaseous 
water. 

(b)  Decomposing,  by  electrolysis,  one  mole  of 
water. 

Explain. 

7.  Pick  the  liquid  having  the  higher  vapor  pressure 
from  each  of  the  following  pairs.  Assume  all 
substances  are  at  room  temperature. 


10. 


(a)  Mercury,  water. 

(b)  Gasoline,  motor  oil. 

(c)  A  perfume,  honey. 

Explain  why  the  boiling  point  of  water  is  lower 
in  Denver,  Colorado  (altitude,  5,280  feet),  than 
in  Boston,  Massachusetts  (at  sea  level). 

Both  carbon  tetrachloride,  CC14  (used  in  dry 
cleaning  and  in  some  fire  extinguishers)  and  mer- 
cury, Hg,  are  liquids  whose  vapors  are  poisonous 
to  breathe.  If  CC14  is  spilled,  the  danger  can  be 
removed  merely  by  airing  the  room  overnight 
but  if  mercury  is  spilled,  it  is  necessary  to  pick 
up  the  liquid  droplets  with  a  "vacuum  cleaner" 
device.  Explain. 

Because  of  its  excellent  heat  conductivity,  liquid 
sodium  has  been  proposed  as  a  cooling  liquid 
for  use  in  nuclear  power  plants. 

(a)  Over  what  temperature  range  could  sodium 
be  used  in  a  cooling  system  built  to  operate 
at  one  atmosphere  pressure  or  lower? 

(b)  How  much  heat  would  be  absorbed  per  kilo- 
gram of  sodium  to  melt  the  solid  when  the 
cooling  system  is  put  in  operation? 

(c)  How  much  heat  would  be  absorbed  per  kilo- 
gram of  sodium  if  the  temperature  rose  too 
high  and  the  sodium  vaporized? 

Use  the  data  in  Tables  5-1  (p.  67)  and  5-II  (p.  69). 


QUESTIONS    AND    PROBLEMS 


83 


11.  Water  is  a  commonly  used  cooling  agent  in 
power  plants.  Repeat  Problem  10  considering 
one  kilogram  of  water  instead  of  sodium.  Con- 
trast the  results  for  these  two  coolants. 

12.  How  much  heat  must  be  removed  to  freeze  an 
ice  tray  full  of  water  at  0°C  if  the  ice  tray  holds 
500  grams  of  water? 

13.  List  three  heterogeneous  materials  not  given  in 
Section  5-2.1. 

14.  List  three  homogeneous  materials  not  given  in 
Section  5-2.1. 

15.  Which  of  the  following  statements  about  sea 
water  is  FALSE? 

(a)  It  boils  at  a  higher  temperature  than  pure 
water. 

(b)  It  melts  at  a  lower  temperature  than  pure 
water. 

(c)  The  boiling  point  rises  as  the  liquid  boils 
away. 

(d)  The  melting  point  falls  as  the  liquid  freezes. 

(e)  The  density  is  the  same  as  that  of  pure  water. 

16.  How  many  grams  of  methanol,  CH3OH,  must 
be  added  to  2.00  moles  of  HjO  to  make  a  solu- 
tion containing  equal  numbers  of  H.O  and 
CH3OH  molecules?  How  many  molecules  (of  all 
kinds)  does  the  resulting  solution  contain? 

17.  How    many    grams    of   ammonium    chloride, 

NH4C1,  are  present  in  0.30  liter  of  a  0.40  M 

NH4C1  solution? 

Answer.  6.4  grams. 

18.  Write  directions  for  preparing  the  following 
aqueous  solutions: 

(a)  1.0  liter  of  1.0  M  lead  nitrate,  Pb(N03)2, 
solution. 

(b)  2.0  liters  of  0.50  M  ammonium  chloride, 
NH4C1,  solution. 

(c)  0.50   liter   of  2.0  M  potassium  chromate, 
K2Cr04,  solution. 

19.  How  many  liters  of  a  0.250  M  K2Cr04  solution 
contain  38.8  grams  of  K2Cr04? 

20.  List  three  properties  of  a  solution  you  would 
expect  to  vary  as  the  concentration  of  the  solute 
varies. 

21.  Give  two  forces  other  than  electric  that  are  felt 
at  a  distance. 


22.  What  would  you  expect  to  observe  if  one  elec- 
trometer sphere  were  charged  by  your  hair  and 
the  other  by  the  comb  used  to  comb  your  hair? 

23.  Why  do  scientists  claim  there  are  only  two  kinds 
of  electric  charge? 

24.  It  is  known  that  electric  charges  attract  or  repel 
each  other  with  a  force  that  is  inversely  propor- 
tional to  the  square  of  the  distance  between 
them.  If  two  spheres  like  those  in  the  electrome- 
ter (Figure  5-7)  are  negatively  charged,  what 
would  be  the  change  in  the  force  of  repulsion  if 
the  distance  between  them  were  increased  to  four 
times  the  original  distance? 

25.  Why  do  two  electrically  neutral  objects  with 
mass  attract  each  other? 

26.  Each  of  the  following  ionic  solids  dissolves  in 
water  to  form  conducting  solutions.  Write  equa- 
tions for  each  reaction. 

(a)  potassium  chloride,  KC1 

(b)  sodium  nitrate,  NaN03 

(c)  calcium  bromide,  CaBr2 

Answer.  CaBr/sj  — >-  Ca+*~(aq)  +  lBr~(aq). 

(d)  lithium  iodide,  Lil 

27.  A  chloride  of  iron  called  ferric  chloride,  FeCl3, 
dissolves  in  water  to  form  a  conducting  solution 
containing  ferric  ions,  Fe+3,  and  chloride  ions, 
CI". 

(a)  Write  the  equation  for  this  reaction. 

(b)  If  0.10  mole  of  FeCl3  is  dissolved  in  water 
and  diluted  to  1.0  liter,  what  is  the  concen- 
tration of  ferric  ion  and  of  chloride  ion? 

Answer.  Concentration  of  Fe+3  =  0.10  M. 
Concentration  of  CI"    =  0.30  M. 

28.  The  salt  ammonium  sulfate,  (NH4)2S04,  dissolves 
in  water  to  form  a  conducting  solution  contain- 
ing ammonium  ions,  NH^,  and  sulfate  ions, 
SO4"2. 

(a)  Write  the  balanced  equation  for  the  reaction 
when  this  ionic  solid  dissolves  in  water. 

(b)  Verify  the  conservation  of  charge  by  com- 
paring the  charge  of  the  reactant  to  the  sum 
of  the  charges  of  the  products. 

(c)  Suppose  1.32  grams  of  ammonium  sulfate  is 
dissolved  in  water  and  diluted  to  0.500  liter. 
Calculate  the  concentrations  of  NH^" (aq) 
and  S04"  2(aq). 


84 


LIQUIDS    AND    SOLIDS:    CONDENSED    PHASES    OF    MATTER    I    CHAP.    5 


29.  1.00  liter  of  solution  contains  0.100  mole  of 
ferric  chloride,  FeCl3,  and  0.100  mole  of  ammo- 
nium chloride,  NH4C1.  Calculate  the  concentra- 
tions of  Fe+3,  C\~,  and  NH4+  ions. 

Answer.  Concentration  Fe+3  =  0.100  M, 
Concentration  NH4+  =0.100  M, 
Concentration  CI"      =  0.400  M. 

30.  In  Experiment  10  you  mixed  lead  nitrate  and 
sodium  iodide.  Write  an  equation  for  the  reac- 
tion that  occurred.  Show  only  the  predominant 
reacting  species. 

31.  Write  equations  for  the  reactions  between  aque- 
ous bromide  ions  and: 

(a)  aqueous  lead  ions, 

(b)  aqueous  silver  ions. 

Both  lead  bromide,  PbBr2,  and  silver  bromide, 
AgBr,  are  only  slightly  soluble. 


32.  When  solutions  of  barium  chloride,  BaCl2,  and 
potassium  chromate,  K2Cr04,  are  mixed,  the 
following  reaction  occurs: 

2K+  (aq)  +  CxOl2  (aq)  +  Ba+YaqJ  +  2C\~(aq) 
— >■  BaCr04fs;  +  2K+(aq)  +  2Cl-(aq; 

(a)  Show  how  charge  is  conserved. 

(b)  Rewrite  the  equation  showing  predominant 
reacting  species  only. 

(c)  Suppose  1.00  liter  of  0.500  M  BaCl2  is  mixed 
with  1.00  liter  of  0.200  A/K2Cr04.  Assuming 
BaCr04  has  negligible  solubility,  calculate 
the  concentrations  of  all  ions  present  when 
precipitation  stops. 

Answer.  Concentration  K+        =  0.200  M, 
Concentration  Or        =  0.500  M, 
Concentration  CrO^2  =  negligible, 
Concentration  Ba+2      =  0.150  M. 


CHAPTER 


6 


Structure  of 
the  Atom  and 
the  Periodic  Table 


The  eighth  element,  starting  from  a  given  one,  is  a  kind  of  repetition  of 
the  first,  like  the  eighth  note  of  an  octave  in  music. 

J.    A.    R.    NEWLANDS,    1864 


We  have  already  learned  that  nature  has  great 
variety.  Around  us  we  find  gases,  liquids,  and 
solids.  To  liquefy  air,  we  must  cool  it  to  about 
—  180°C,  far  colder  than  the  coldest  winter.  To 
liquefy  rock,  we  must  heat  it  to  temperatures 
above  1000°C,  the  climate  found  in  an  active 
volcano.  When  we  examined  chemical  reactivity, 
we  found  even  more  variety.  A  candle  burns 
quietly  and  slowly,  once  lit,  though  it  does  not 
react  appreciably  until  lit.  Iron  also  reacts  with 
oxygen  very  slowly  (it  rusts),  though  not  as 
slowly  as  we  might  like.  Hydrogen,  by  contrast, 
reacts  explosively  with  oxygen  when  it  is  ignited. 
In  contrast  to  the  slow  reactions  of  paraffin  wax 
and  iron  with  oxygen,  and  the  instantaneous  re- 
action of  hydrogen  with  oxygen,  helium  gas  will 
never  react  with  oxygen. 

Turning  to  the  atomic  view  of  matter,  we  find 
more  than  a  hundred  different  elements.  Each  of 
these  elements  has  a  kind  of  atom  that  is  some- 
how different  from  all  of  the  others.  With  these 
100  elements,  chemists  have  prepared  about  one 


and  one-half  million  different  compounds,  each 
having  its  own  special  properties.  Each  year 
about  100,000  new  compounds  are  reported. 
Again  we  must  deal  with  great  variety. 

We  have  already  remarked  in  Chapter  1  that 
"the  mere  cataloguing  of  observations  is  not 
science."  We  could  never  cope  with  this  great 
variety  in  nature  if  we  did  not  make  use  of  its 
regularities  in  organizing  our  knowledge.  The 
fact  that  chemists  have  been  able  to  synthesize 
more  than  a  million  compounds  shows  that  they 
have  been  successful  in  this  organization.  Their 
success  stems  in  large  part  from  the  regularities 
embodied  in  the  periodic  table. 

The  periodic  table  groups  elements  with  simi- 
lar chemistry.  It  is  of  great  value  just  as  a 
correlating  device.  It  is  even  more  powerful  when 
coupled  with  an  understanding  of  the  structure 
of  atoms.  So,  it  is  appropriate  to  consider  this 
topic  before  examining  the  relationships  that  es- 
tablish the  periodic  table. 


85 


86 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


6-1    STRUCTURE  OF  THE  ATOM 


Scientists  have  developed  a  highly  sophisticated 
view  of  the  structure  of  the  atom.  The  currently 
accepted  model  is  called  "the  nuclear  atom."  We 
shall  present  it  without  trying  to  show  immedi- 
ately all  of  the  experimental  evidence  that  led  to 
this  particular  model.  Rest  assured,  though,  that 
every  feature  of  the  nuclear  atom  picture  rests 
upon  experimental  evidence,  as  we  shall  see  in 
Chapter  14. 

6-1.1    A  Model:  The  Nuclear  Atom 

An  atom  contains  electrons  and  protons.  Since 
mass  is  associated  with  all  matter,  it  is  natural 
to  assume  that  atoms,  which  form  matter,  have 
mass.  And  since  any  sample  of  matter  occupies 
a  certain  volume,  we  can  also  assume  that  each 
atom  has  volume.  Almost  all  the  mass  of  the 
atom  is  concentrated  in  a  region  that  is  much 
smaller  than  the  total  volume  of  the  atom.  This 
region  is  called  the  nucleus  of  the  atom.  The  rest 
of  the  volume  of  the  atom  is  occupied  by  electrons. 
The  nucleus  carries  a  positive  electric  charge. 
The  element  hydrogen  has  the  lightest  atoms, 
and  the  nuclei  of  these  atoms  have  the  smallest 
positive  charge  anyone  has  observed.  Every  atom 
of  hydrogen  has  one  proton  in  its  nucleus.  The 
charge  on  the  nucleus  of  an  atom  of  hydrogen 


is  that  of  a  single  proton,  1+  unit  of  charge. 
All  other  nuclei  have  positive  charges  that  are 
exactly  an  integer  times  the  proton  charge;  a 
nucleus  may  have  2+  units,  3+  units,  4+  units, 
and  so  on.  Each  nucleus  contains  a  definite  num- 
ber of  protons  and  the  charge  on  the  nucleus  is 
fixed  by  this  number.  All  atoms  of  a  particular 
element  have  the  same  nuclear  charge.  All  hydro- 
gen atoms  have  a  nuclear  charge  of  1  + ;  all 
helium  atoms  have  a  nuclear  charge  of  2+ ;  all 
lithium  atoms  have  a  nuclear  charge  of  3+ ;  and 
so  on.  We  shall  see  that  the  nuclear  charge  deter- 
mines the  chemistry  of  an  atom. 

Since  the  nucleus  has  positive  charge,  it  at- 
tracts electrons  (each  with  negative  charge).  If  a 
nucleus  attracts  the  number  of  electrons  just 
equal  to  the  nuclear  charge,  an  electrically  neu- 
tral atom  is  formed.  Consider  a  nucleus  contain- 
ing two  protons,  a  helium  nucleus.  When  the 
helium  atom  has  two  electrons  as  well  (2— 
charge),  an  electrically  neutral  helium  atom 
results: 

2  protons  +  2  electrons  =  no  charge 

(2+)     +      (2-)       =         0 

Electrons  can  be  removed  from  or  added  to  a 
neutral  atom,  giving  it  a  net  charge.  This  is  how 
ions  are  formed.  Thus, 


'neutral 
helium  atom, 

2  protons 
2  electrons 

He 


+  energy 


+  energy 


'positive 
helium  ion, 

2  protons 
.1  electron 

He+ 


+  (electron) 


+ 


(/) 


(/) 


and,  as  another  example, 


"neutral 
fluorine  atom, 

9  protons 
_9  electrons 


+  energy 


"positive 
fluorine  ion, 

9  protons 
.8  electrons  _ 


+  (electron) 


(2) 


+  energy 


F+ 


+ 


(2) 


SEC.    6-1    I   STRUCTURE    OF   THE    ATOM 


87 


Since  a  positive  charge  attracts  a  negative  charge, 
it  is  difficult  to  take  the  electron  away  from  the 
positive  helium  or  fluorine  nucleus.  Scientists 
say  "work  must  be  done"  or  "energy  is  required" 
to  form  a  positive  ion  from  a  neutral  atom,  as  in 
(7  or  (2).  "Work"  and  "energy"  are  synony- 
mous here  and  they  indicate  that  an  external 
agent  must  exert  force  on  the  electron  to  make 
it  leave  the  neutral  atom.  In  analogy,  the  attrac- 
tion between  electron  and  nucleus  is  like  a 
stretched  rubber  band  connecting  the  two  par- 
ticles. Continued  force  applied  to  the  two  parti- 
cles can  result  in  the  rubber  band  being  stretched 
until  it  breaks,  releasing  the  two  particles,  but  at 
the  expense  of  work. 

Some  neutral  atoms  can  gain  electrons,  form- 
ing negative  ions.  Thus  a  neutral  fluorine  atom 
can  add  an  electron  to  form  a  negative  ion,  F~. 
This  change,  for  fluorine  atoms,  does  not  require 
the  input  of  energy;  it  releases  energy: 


discovered.  The  neutron  carries  no  charge;  it  is 
a  neutral  particle.  Its  mass  is  almost  identical  to 
the  mass  of  the  proton.  Thus  the  nucleus  of  the 
helium  atom  must  consist  of  two  neutrons  and 
two  protons.  Then  its  charge  will  be  2+  but  its 
mass  will  be  four  times  the  mass  of  the  hydrogen 
atom. 

Now  our  nuclear  model  suffices.  We  can  build 
up  the  atoms  for  all  elements.  Each  atom  has  a 
nucleus  consisting  of  protons  and  neutrons.  The 
protons  are  responsible  for  all  of  the  nuclear 
charge  and  part  of  the  mass.  The  neutrons  are 
responsible  for  the  rest  of  the  mass  of  the  nu- 
cleus. The  neutron  plays  a  role  in  binding  the 
nucleus  together,  apparently  adding  attractive 
forces  which  predominate  over  the  electrical  re- 
pulsions among  the  protons.* 

Around  the  nucleus  are  enough  electrons  to 
make  the  atom  as  a  whole,  electrically  neutral. 


"neutral 
fluorine  atom, 

9  protons 
9  electrons 


-+-  (electron) 


+ 


"negative 
fluorine  ion, 

9  protons 
.10  electrons. 

F" 


+  energy 


+  energy 


(J) 


(i) 


6-1.2    The  Mass  of  an  Atom  and  Its  Parts 

Protons  are  in  the  nucleus  and  electrons  sur- 
round it.  Most  of  the  mass  of  he  atom  is  in  the 
nucleus.  These  two  statements  imply  that  an 
electron  weighs  far  less  than  a  proton;  this  is  the 
case.  Experiments  have  been  performed  in  which 
individual  electrons  and  protons  have  been 
weighed  (they  are  described  in  Chapter  14). 
These  experiments  show  that  the  mass  of  the 
electron  is  smaller  than  that  of  a  proton  by  a 
factor  of  -n^. 

This  means  that  most  of  the  mass  of  the  atom 
must  be  furnished  by  the  nucleus.  However,  the 
mass  of  the  nucleus  is  not  determined  by  the 
number  of  protons  alone.  For  example,  a  helium 
nucleus  has  two  protons  and  a  hydrogen  nucleus 
has  one  proton.  Yet  a  helium  atom  is  measured 
to  be  four  times  heavier  than  a  hydrogen  atom. 
What  can  be  the  composition  of  the  helium  nu- 
cleus? A  partial  answer  to  this  problem  was 
obtained  when  a  third  particle,  the  neulroD,  was 


Table  6-1 

CHARGE  AND  MASS  OF  SOME 
FUNDAMENTAL  PARTICLES 


PARTICLE 


CHARGE 

(relative  to  the 
electron  charge) 


APPROXIMATE  MASS 

(relative  to  the 
mass  of  a  proton) 


The  charge  and  mass  of  each  of  the  three  funda- 
mental particles  we  have  discussed  are  shown  in 
Table  6-1. 


*  This  model  still  does  not  explain  what  forces  hold 
the  nucleus  together  in  spite  of  the  proton  repulsions. 
We  do  know  that  the  helium  nucleus  is  stable — it  can 
exist  indefinitely — but  the  model  does  not  explain  why  it 
is  stable.  Thus  we  use  models  because  they  help  us  to 
explain  many  important  facts,  even  though  they  do  not 
explain  aLl  the  facts. 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


6-1.3    The  Sizes  of  Atoms 

How  large  is  an  atom?  We  cannot  answer  this 
question  for  an  isolated  atom.  We  can,  however, 
devise  experiments  in  which  we  can  find  how 
closely  the  nucleus  of  one  atom  can  approach 
the  nucleus  of  another  atom.  As  atoms  approach, 
they  are  held  apart  by  the  repulsion  of  the  posi- 
tively charged  nuclei.  The  electrons  of  the  two 
atoms  also  repel  one  another  but  they  are  at- 
tracted by  the  nuclei.  The  closeness  of  approach 
of  two  nuclei  will  depend  upon  a  balance  between 
the  repulsive  and  attractive  forces.  It  also  de- 
pends upon  the  energy  of  motion  of  the  atoms 
as  they  approach  one  another.  If  we  think  of 
atoms  as  spheres,  we  find  that  their  diameters 
vary  from  0.000  000  01  to  0.000  000  05  cm  (from 
1  X  10~8  to  5  X  lO-8  cm).  Nuclei  are  much 
smaller.  A  typical  nuclear  diameter  is  10-13  cm, 
about  1/100,000  the  atom  diameter. 

Suppose  we  take  Yankee  Stadium  (seating 
capacity,  67,000)  as  a  model  for  the  atom.  To 
keep  the  proper  scale,  the  nucleus  would  be 
about  the  size  of  a  flea!  For  the  hydrogen  atom, 
the  flea  would  represent  one  proton.  He  would 
be  located  at  the  center  of  the  stadium,  some- 


Fig.  6-1.  The  nucleus  is  much  smaller  than  the  atom. 

The  nucleus  occupies  about  the  same  rela- 
tive volume  in  the  atom  as  does  a  flea  in 
Yankee  Stadium. 


where  behind  second  base.  The  one  electron 
present  in  the  neutral  atom  would  wander  hither 
and  yon,  occupying  all  the  rest  of  the  stadium. 
For  the  helium  atom,  the  nucleus  would  be  re- 
placed by  a  cluster  of  four  fleas — representing 
two  protons  and  two  neutrons.  The  two  electrons 
of  the  neutral  helium  atom  would  now  share 
the  ample  space  of  the  huge  stadium.  This  situa- 
tion is  made  even  more  difficult  to  picture  by  the 
fact  that  the  fleas,  occupying  a  minute  volume  in 
center  field,  represent  almost  all  of  the  mass  that 
is  carried  in  the  stadium. 

6-1.4    Atomic  Number 

What  do  the  atoms  of  one  element  have  in  com- 
mon to  distinguish  them  from  the  atoms  of  all 
other  elements?  Each  hydrogen  nucleus  bears  a 
charge  of  1  +  .  Each  neutral  hydrogen  atom  has 
one  electron,  charge  1  —  ,  situated  in  the  rela- 
tively large  volume  of  the  atom  outside  the  nu- 
cleus. Each  helium  atom  has  a  nucleus  whose 
charge  is  2+  and  each  neutral  atom  has  two 
electrons  around  the  nucleus.  The  element  lith- 
ium has  atoms  heavier  than  hydrogen  or  helium, 
and  each  lithium  atom  has  a  nucleus  whose 
charge  is  3+.  Three  electrons  around  the  nucleus 
are  required  to  form  a  neutral  lithium  atom. 

Thus,  each  of  the  chemical  elements  consists 
of  atoms  whose  nuclei  contain  a  particular  num- 
ber of  protons,  hence  a  particular  nuclear  charge. 
The  number  of  protons  in  the  nucleus  is  called  the 


SEC.    6-1    I    STRUCTURE    OF    THE    ATOM 


89 


atomic  number.  Of  course,  all  atomic  num- 
bers are  whole  numbers.  Thus,  oxygen  with 
atomic  number  8,  has  eight  protons  in  the  nu- 
cleus (nuclear  charge  8+).  A  neutral  oxygen 
atom  must  have  eight  electrons  (each  with  charge 
1  —  )  as  well. 

The  atomic  number  of  each  of  the  elements  is 
listed  in  the  table  on  the  inside  of  the  back  cover 
of  this  book.  You  will  find  there  that  each  ele- 
ment has  a  distinctive  name,  symbol,  and  atomic 
number.  A  given  element  can  be  identified  by 
any  of  these.  For  example,  helium  can  be  called 
by  its  name,  helium,  by  its  symbol,  He,  or  by  its 
atomic  number,  the  element  of  atomic  number  2. 

In  the  periodic  table  we  shall  see  that  the  ele- 
ments have  been  listed  in  the  order  of  increasing 
atomic  number.  An  example  of  the  periodic  table 
is  on  the  wall  of  your  classroom  and  there  is  a 
copy  on  the  inside  of  the  cover  at  the  front  of 
this  book.  In  each  box  in  the  table  the  number 


above  the  chemical  symbol  for  the  element  is  the 
atomic  number. 

6-1.5    Mass  Number  and  Isotopes 

All  of  the  atoms  of  an  element  have  the  same 
nuclear  charge.  Do  all  of  the  atoms  of  an  element 
have  the  same  mass?  Almost  all  hydrogen  atoms 
do  have  the  same  mass — the  sum  of  the  mass  of 
one  proton  and  the  mass  of  one  electron.  For 
these  atoms  the  nucleus  consists  of  a  single  pro- 
ton. However,  a  small  fraction  of  the  hydrogen 
atoms  0.016%  of  them)  have  nuclei  whose  mass 
is  approximately  twice  as  great  as  the  mass  of 
the  proton.  (Compare  with  the  helium  nucleus.) 
To  explain  the  mass  of  these  hydrogen  atoms, 
we  conclude  that  each  of  their  nuclei  consists  of  a 
neutron  (charge  zero,  mass  1)  and  a  proton 
(charge  1  +  ,  mass  1).  This  kind  of  hydrogen  atom 
is  called  hydrogen-2;  another  name  commonly 


Table  6-II.    vital  statistics  of  some  common  isotopes 


NAME 

OF 

ISOTOPE 


ABUNDANCE 

IN 

NATURE 


ATOMIC 
NUMBER 


MASS 
NUMBER 


COMPOSITION' 


NUCLEUS  NUMBER  OF 

ELECTRONS  IN 

MASS         CHARGE  NEUTRAL  ATOM 


hydrogen- 1 
hydrogen-2 

99.984% 
0.016 

1 
1 

1 
2 

lp, 

1/7 

1 
2 

1  + 
1  + 

1 
1 

helium-3 
helium-4 

1.34  X  10-4 
100 

2 

2 

3 
4 

2p, 
2p, 

1/7 
In 

3 
4 

2  + 
2  + 

2 
2 

lithium-6 
lithium-7 

7.40 
92.6 

3 
3 

6 
7 

3p, 
3/7, 

3/7 

4/7 

6 

7 

3  + 
3  + 

3 

3 

beryllium-9 

100 

4 

9 

4/>, 

5/1 

9 

4  + 

4 

boron-10 
boron- 11 

18.83 
81.17 

5 
5 

10 
11 

Sp, 
Sp, 

5/7 
6/7 

10 
11 

5  + 
5  + 

5 
5 

carbon-12 
carbon-13 

98.892 
1.108 

6 
6 

12 
13 

6p, 

6/7, 

6/7 
In 

12 
13 

6+ 
6+ 

6 
6 

nitrogen-14 
nitrogen- 15 

99.64 
0.36 

7 

7 

14 
15 

lp, 
lp, 

In 
8/: 

14 
15 

7  + 

7  + 

7 
7 

oxygen- 16 
oxygen-17 
oxygen- 18 

99.76 
0.04 
0.20 

8 
8 
8 

16 
17 
18 

Sp, 
Sp, 
Sp, 

8/7 

9/7 

10/7 

16 
17 
18 

8  + 
8  + 
8  + 

8 
8 
8 

fluorine-19 

100 

9 

19 

9p, 

10/7 

19 

9  + 

9 

chlorine-35 
chlorine-37 

75.4 
24.6 

17 

17 

35 
37 

lip, 
I7p, 

18/7 

20/7 

35 
37 

17  + 
17  + 

17 
17 

gold- 197 

100 

79 

197 

79/7, 

118/7 

197 

79  + 

79 

uranium-235 
uranium-238 

0.71 

99.28 

92 
92 

235 
238 

92/7, 
92p, 

143/7 
146/7 

235 
238 

92  + 

92  + 

92 
92 

1  p  =  proton,  n  =  neutron. 


90 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE   I    CHAP.    6 


used  is  deuterium.  The  two  kinds  of  hydrogen 
atoms  (having  the  same  atomic  number  but  dif- 
ferent masses)  are  called  isotopes.  An  isotope  is 
identified  by  specifying,  first,  which  element  it  is 
(usually  by  the  symbol  or  name  of  the  element) 
and,  second,  the  sum  of  the  number  of  protons 
and  the  number  of  neutrons.  The  number  of  pro- 
tons plus  the  number  of  neutrons  in  a  nucleus  is 
called  the  mass  number  of  the  nucleus.  The  mass 
number  is,  of  course,  always  an  integer. 
Reviewing: 

Atomic  number  =  number   of  protons   in   the 
nucleus  (fixes  nuclear  charge) 

Mass  number  =  number  of  protons  and  neu- 
trons in  the  nucleus  (fixes 
nuclear  mass) 

Most  of  the  chemical  elements  consist  of  mix- 
tures of  isotopes.  Oxygen,  atomic  number  8,  has 
three  stable  isotopes.  The  kind  having  mass  num- 
ber 16  is  most  abundant.  About  99.76%  of  the 
oxygen  atoms  consist  of  this  isotope.  Only  0.04% 


of  the  oxygen  atoms  have  mass  number  17  and 
about  0.20%  have  mass  number  18.  The  nucleus 
of  an  oxygen- 16  atom  consists  of  eight  protons 
and  eight  neutrons — charge  8+,  mass  16.  The 
nucleus  of  an  oxygen- 17  atom  consists  of  eight 
protons  and  nine  neutrons — charge  8+,  mass 
17.  The  nucleus  of  an  oxygen- 18  atom  consists 
of  eight  protons  and  ten  neutrons — charge  8+, 
mass  18.  Table  6-II  summarizes  the  data  on  the 
atomic  structure  of  a  few  common  isotopes. 

The  nuclear  charge  and  the  electrons  it  attracts 
primarily  determine  the  ways  in  which  atoms 
behave  toward  other  atoms.  Mass  differences 
cause  only  minor  chemical  effects.  Since  the  iso- 
topes of  an  element  have  the  same  nuclear  charge 
and  the  same  number  of  electrons  per  neutral 
atom,  they  react  in  the  same  ways.  Thus  we  can 
speak  of  the  chemistry  of  oxygen  without  speci- 
fying which  one  of  the  three  stable  isotopes  is 
reacting.  Only  the  most  precise  measurements 
will  indicate  the  very  slight  chemical  differences 
among  them. 


6-2    THE  SIMPLEST  CHEMICAL  FAMILY— THE  INERT  GASES 


We  are  now  ready  to  consider  why  the  elements 
are  arranged  as  they  are  in  the  periodic  table.  We 
shall  examine  the  known  elements  to  discover 
the  significance  and  usefulness  of  this  table — a 
table  so  important  it  is  printed  on  the  inside 
cover  of  this  and  almost  every  other  general 
chemistry  textbook. 

When  the  elements  are  arranged  in  order  of 
increasing  atomic  number,  each  successive  ele- 
ment has  different  chemistry  from  its  neighbors. 
On  the  other  hand,  there  are  striking  likenesses 
among  some  of  the  elements.  For  example,  of  the 
one-hundred  or  so  known  elements,  only  about 
a  dozen  are  gases  at  normal  conditions  of  tem- 
perature and  pressure.  Of  these,  six  elements  are 
so  remarkably  similar  in  chemistry  that  they  are 
conveniently  considered  together.  Further,  they 
are  identified  by  a  family  name,  "the  inert  gases." 

If  we  look  at  the  properties  of  other  elements, 
we  find  that  this  recurrence  of  properties  is  usual. 
It  is  convenient  to  group  all  of  the  elements 


according  to  their  chemistry  into  families  or 
groups.  The  elements  in  a  particular  group  have 
similar  chemistry.  Knowledge  about  one  element 
in  a  group  then  aids  in  understanding  the  chem- 
istry of  other  elements  in  that  group.  In  the 
periodic  table  each  group  appears  in  a  vertical 


Fig.  6-2.  The  inert  gas  elements — a  chemical  family. 


helium 


SEC.    6-2    I    THE    SIMPLEST    CHEMICAL    FAMILY  -THE    INERT    GASES 


91 


column.  The  key  to  this  arrangement  lies  in  the 
elements  called  the  inert  gases.  It  is  one  of 
nature's  quirks  that  these  gases  provide  the  key 
to  the  organization  of  our  chemical  knowledge 
because  these  elements  are  distinctive  in  their 
almost  total  absence  of  chemical  reactivity.  The 
first  of  the  inert  gases  is  helium. 

6-2.1    Helium 

Helium,  the  second  element  in  the  periodic  table, 
has  atomic  number  2.  This  means  its  nucleus 
contains  two  protons  and  has  a  2+  charge.  The 
neutral  atom,  then,  contains  two  electrons.  There 
are  two  stable  isotopes,  helium-4  and  helium-3, 
but  the  helium  found  in  nature  is  almost  pure 
helium-4.  Helium  is  found  in  certain  natural  gas 
fields  and  is  separated  as  a  by-product.  Sources 
of  helium  are  rare  and  most  of  the  world  supply 
is  produced  in  the  United  States,  mainly  in 
Texas  and  Kansas. 

Helium  is  a  monatomic  gas  and,  as  yet,  no 
stable  compounds  of  helium  have  been  found. 
The  attractive  forces  between  the  atoms  of  he- 
lium are  unusually  weak,  as  shown  by  the  normal 
boiling  point.  To  liquefy  helium,  it  must  be 
cooled  to  -268.9°C  or  4.2°K.  No  other  element 
or  compound  has  a  boiling  point  as  low.  Helium 
has  another  distinction  which  reflects  these  weak 
forces:  it  is  the  only  substance  known  which 
cannot  be  solidified  at  any  temperature  unless  it 
is  subjected  to  pressure.  Helium  becomes  solid 
at  1.1  °K  at  a  pressure  of  26  atmospheres. 

Chemical  reactions  tend  to  occur  among  at- 
oms to  form  more  stable  arrangements  of  the 
atoms.  Helium  takes  part  in  no  chemical  reac- 
tions. Helium  atoms  must  be  particularly  stable. 


6-2.2    Neon,  Argon,  Krypton,  Xenon,  Radon 

Among  all  of  the  hundred  or  so  known  elements, 
there  are  only  five  with  chemistry  resembling  that 
of  helium.  We  have  remarked  that  these  ele- 
ments; neon,  argon,  krypton,  xenon,  and  radon, 
together  with  helium,  are  known  as  the  inert 
gases.  It  is  only  since  1962  that  chemists  have 
been  able  to  prepare  any  compounds  at  all  in- 
volving these  elements.  The  few  compounds  that 
have  been  prepared  are  extremely  reactive,  de- 
composing readily  to  restore  the  inert  gas  to  the 
elementary  state.  Everything  that  is  known  of 
the  chemistry  of  the  inert  gas  elements  indicates 
that  the  atoms  of  the  inert  gases  must  be  par- 
ticularly stable. 

The  physical  properties  of  the  inert  gases  are 
shown  in  Table  6-III.  There  is  much  information 
contained  in  this  table,  and  we  shall  examine  it 
in  parts. 

BOILING    POINT 

The  inert  gas  elements  are  all  gases  at  room 
temperature.  Helium  has  the  lowest  known  boil- 
ing point,  4.2°K.  Neon  has  the  third  lowest 
known  boiling  point,  27.2°K.  (Hydrogen,  H2,  has 
the  second  lowest  known  boiling  point,  20.4°K.) 
The  boiling  point  of  argon  is  still  quite  low, 
87.3°K,  but  not  so  low  as  to  distinguish  it  from 
a  number  of  diatomic  molecules.  (Compare  the 
boiling  points  of  nitrogen,  N2,  b.p.  =  77.4°K; 
fluorine,  F2,  b.p.  =  85.2°K;  oxygen,  02,  b.p.  = 
90.2°K.)  Krypton,  xenon,  and  radon  have  suc- 
cessively higher  boiling  points.  Apparently,  as 
the  atomic  number  goes  up,  the  boiling  point 
goes  up.  Figure  6-3  shows  the  boiling  point  trend 
of  the  inert  gases  in  a  plot  of  boiling  point  against 
atomic  number.  Since  the  atomic  number  gives 


Table  6-III.     some  properties  of  the  inert  gas  elements 

property  He  Ne  Ar  Kr 


Xe 


Rn 


Atomic  number 

2 

10 

18 

36 

54 

86 

Atomic  weight 

4.00 

20.2 

39.9 

83.7 

131 

222 

Boiling  point  (°K) 

4.2 

27.2 

87.3 

120 

165 

211 

Melling  point  (°K) 

— 

24.6 

83.9 

116 

161 

202 

Atomic  volume,  liquid 

(ml  /mole  of  atoms) 

31.8 

16.8 

28.5 

32.2 

42.9 

50.5 

92 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


k 

200 

■ 

.S 

*C    150 

' 

♦• 

.R 

// 

0 

$100 

K 

f 

^ 

yy 

rii 

/ 

50 

~f/ 

< 

20  40  60  80  IOC 

Atomic    number, 
Number    of  electrrorts 

Fig.  6-3.  The  correlation  of  boiling  point  and  number 
of  electrons  per  atom  for  the  inert  gases. 


the  number  of  protons  in  the  nucleus,  it  also 
gives  the  number  of  electrons  held  by  each  atom. 
We  can  interpret  a  higher  boiling  point  to  mean 
that  more  energy  must  be  added  to  disrupt  the 
liquid  state.  Hence,  the  weak  attractive  forces 
which  cause  the  inert  gases  to  liquefy  increase  as 
the  number  of  electrons  per  atom  increases. 

MELTING    POINT 

At  temperatures  only  slightly  below  the  liquefac- 
tion temperatures,  the  liquids  freeze.  The  solids 
are  all  simple  crystals  in  which  the  atoms  are 
close-packed  in  a  regular  lattice  arrangement. 
The  narrow  temperature  range  over  which  any 
one  of  these  liquids  can  exist  suggests  that  the 
forces  holding  the  crystal  together  are  very  much 
like  the  forces  in  the  liquid. 

ATOMIC    VOLUME 

We  picture  the  atoms  in  a  liquid  and  in  a  solid 
as  being  packed  rather  tightly.  The  packing  is 
random  in  the  liquid  and  regular  in  the  solid. 
With  this  picture,  we  can  deduce  from  the  vol- 
ume per  mole  of  atoms  the  volume  to  be  assigned 
to  a  single  atom.  Consulting  Table  6-1II,  we  find 
that  helium  is  distinctive  in  its  atomic  volume 


(indeed,  as  it  is  in  every  other  property  listed  in 
this  table).  Otherwise,  we  see  that  volume  per 
mole  of  atoms  increases  with  atomic  number, 
that  is,  with  the  number  of  electrons  around  the 
nucleus.  Notice  that  the  volume  per  mole  of 
atoms  of  neon  (16.8  ml)  is  only  slightly  less  than 
that  of  water  (18  ml).  The  water  molecule  occu- 
pies slightly  more  space  than  does  the  neon 
atom. 

PHYSICAL    PROPERTIES:    SUMMARY 

We  have  compared  the  physical  properties  of  the 
inert  gases  among  themselves  and  with  the  cor- 
responding properties  of  a  few  simple  diatomic 
molecules.  We  find  that  the  forces  between  the 
atoms  are  indeed  weak  though  comparable  to 
the  forces  between  some  stable  molecules.  The 
physical  properties,  however,  do  not  distinguish 
this  group  of  elements.  The  uniqueness  of  the 
inert  gases  relates  to  compound  formation;  when 
compared  with  the  other  elements,  the  inert  gases 
are  seen  to  have  a  specially  small  tendency  to 
form  stable  compounds. 

6-2.3     Number  of  Electrons  and  Stability 

The  inert  character  of  stable  compounds  makes 
them  of  special  interest  to  us.  We  might  make  a 

Fig.  6-4.  Trends  in  (he  physical  properties  of  the  inert 
gases. 

200°K-        (a)  mpl'  °K 

100  "K-  _ 1'      """" 


He 


Ne 


Ar 


Kr 


Xe 


200*-        (»*-P:°K 

100'K- 


He 


Ne 


Ar 


Kr 


Xe 


40  r         (c)  Atomic  volume, 

ml /mole 
30 


SEC.    6-3    I    THE    ALKALIES 


93 


list  of  the  number  of  electrons  possessed  by  each 
inert  gas.  (Remember  the  lost  child  organizing 
his  information?)  We  shall  find  these  numbers 
to  be  of  particular  value  in  all  of  our  future  study 
of  chemistry. 

In  Table  6-IV  there  is  much  food  for  thought. 
First,  each  of  the  especially  stable  atoms  has  an 
even  number  of  electrons.  Next  we  see  that  there 
seems  to  be  some  regularity  in  the  differences 
between  the  number  of  electrons  possessed  by  a 
given  inert  gas  and  that  of  its  predecessor.  The 
first  two  differences  are  eight  and  the  second  two 
differences  are  eighteen.  If  there  were  another 
inert  gas,  would  it  have  86  +  32  =  118  elec- 
trons? What  is  the  special  significance  of  the 
numbers  8,  18,  and  32?  Is  it  true  for  any  other 
element  that  the  electron  arrangements  of  he- 
lium, neon,  argon,  and  so  on  are  especially 
stable?  We  shall  see  that  this  is  true,  not  just  for 
some  elements,  but  for  all  elements. 

Table  6-IV 

THE  ELECTRON  POPULATIONS 
OF  THE  INERT  GASES 

INERT  TOTAL  NUMBER  CHANGE  IN  NUMBER 

GAS  OF  ELECTRONS  OF  ELECTRONS 


helium 

2 

neon 

10 

10  -    2  =    8 

argon 

18 

18  -  10  =    8 

krypton 

36 

36  -  18  =  18 

xenon 

54 

54  -  36  =  18 

radon 

86 

86  -  54  =  32 

6-2.4    Sodium  Chloride — Atoms  Trying 
to  Be  Inert  Gas  Atoms 

Sodium  chloride  is  a  compound  of  an  element, 
chlorine,  which  just  precedes  an  inert  gas,  and 


an  element,  sodium,  which  just  follows  an  inert 
gas.  Sodium  has  atomic  number  (nuclear  charge) 
1 1 ,  one  larger  than  neon,  which  has  atomic  num- 
ber 10.  Hence,  the  neutral  atom  of  sodium  has 
one  more  electron  than  the  number  held  by  the 
especially  stable  neon  atom.  Chlorine  has  atomic 
number  (nuclear  charge)  17,  one  below  that  of 
argon,  18.  The  neutral  atom  of  chlorine  has  one 
less  than  the  number  of  electrons  held  by  the 
especially  stable  argon  atom.  We  find  sodium 
and  chlorine  combined  in  a  one-to-one  ratio  in 
the  very  stable  compound  called  sodium  chlo- 
ride, NaClfsJ. 

We  have  already  discussed  the  structure  of 
solid  sodium  chloride  in  Chapter  5.  We  said 
there,  "On  the  basis  of  much  experimental  evi- 
dence, chemists  have  concluded  that  sodium 
chloride  crystals  are  built  up  of  charged  particles 
rather  than  of  neutral  atoms."  The  discussion 
went  on  to  identify  the  ions  in  the  lattice  as  Cl~ 
ions  packed  tightly  around  Na+  ions  (see  Figure 
5-10  on  page  81).  But  how  many  electrons  does 
a  chlorine  ion,  Cl~,  have?  By  gaining  an  electron, 
the  chlorine  now  has  18  electrons,  exactly  the 
same  number  of  electrons  as  does  argon,  the 
adjacent  inert  gas.  In  a  similar  way,  but  by  losing 
an  electron,  the  sodium  has  contrived  to  reach 
the  10  electron  population  of  its  adjacent  inert 
gas,  neon.  The  atoms  reached  these  inert  gas-like 
electron  arrangements  through  compound  for- 
mation and  the  resulting  compound  thereby 
acquired  some  of  the  unique  stability  of  the 
inert  gases. 

With  this  in  mind,  let  us  explore  the  chemistry 
of  all  of  the  elements  immediately  adjacent  to 
the  inert  gases.  These  two  vertical  columns  of 
the  periodic  table  are  called  the  alkalies  and  the 
halogens. 


6-3    THE  ALKALIES 


The  six  elements  adjacent  to  and  following  the 
six  inert  gases  are  lithium,  sodium,  potassium, 
rubidium,  cesium,  and  francium.  These  elements 
have   similar   chemistries    and    are   called    the 


alkalies  (or,  the  alkali  metals).  Figure  6-5  shows 
that  these  elements  are  neighbors  to  the  inert 
gases.  Their  chemistries  can  be  understood  in 
terms  of  the  special  stabilities  of  the  1  +  ions 


94 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


z 
He 

Wk.r*& 

10 

Ne 

18 
At* 

36 

54 
Xc 

as 

^■*J^H 

lithium 
Sodium 
potassium 
■rubidium 
cesium, 
francium 


Fig.  6-5.  The  alkali  metals. 


which  have  the  inert  gas  electron  arrangements. 

All  the  alkalies  are  metals  in  their  elementary 
states.  When  the  metal  surfaces  are  clean,  they 
have  a  bright,  silvery  luster.  The  metals  are  ex- 
cellent conductors  of  electricity  and  heat.  They 
are  soft  and  malleable,  and  have  low  melting 
points  (compared  with  almost  all  other  elemen- 
tary metals). 

In  Chapter  5  we  identified  metals  by  their  high 
electrical  conductivity.  Now  we  can  explain  why 
they  conduct  electric  current  so  well.  It  is  because 
there  are  some  electrons  present  in  the  crystal 
lattice  that  are  extremely  mobile.  These  "con- 
duction electrons"  move  throughout  the  metallic 
crystal  without  specific  attachment  to  particular 
atoms.  The  alkali  elements  form  metals  because 
of  the  ease  of  freeing  one  electron  per  atom  to 
provide  a  reservoir  of  conduction  electrons.  The 
ease  of  freeing  these  conduction  electrons  derives 
from  the  stability  of  the  residual,  inert  gas-like 
atoms. 


Fig.  6-6.  Potassium  is  soft;  it  can  be  cut  with  a  knife. 


6-3.1    Physical  Properties  of  the  Alkali  Elements 

Table  6-V  lists  the  same  properties  for  the  alkali 
metals  that  were  listed  in  Table  6-III  for  the 
inert  gases. 

BOILING    POINT   AND    MELTING   POINT 

All  the  alkali  metals  are  solids  at  room  tempera- 
ture, though  cesium  melts  just  above  room  tem- 
perature. Notice  that  both  melting  points  and 
boiling  points  decrease  as  atomic  number  in- 
creases (the  opposite  of  the  inert  gas  behavior). 
Figure  6-7a  and  6-7b  contrast  the  trends  in  the 
alkali  metal  melting  and  boiling  points  to  the 
opposite  trends  in  the  inert  gases.  Notice  also 
the  extremely  wide  temperature  range  over  which 
the  alkali  liquids  are  stable.  Sodium,  for  ex- 
ample, melts  at  371°K  and  boils  at  1162°K, 


Table  6-V.     some  properties  of  the  alkali  elements 


PROPERTY 

LITHIUM 

SODIUM 

POTASSIUM 

RUBIDIUM 

CESIUM 

Atomic  number 

3 

11 

19 

37 

55 

Atomic  weight 

6.94 

23.0 

39.1 

85.4 

133 

Boiling  point  (°K) 

1599 

1162 

1030 

952 

963 

(°C) 

1326 

889 

757 

679 

690 

Melting  point  (°K) 

453 

371 

336.4 

311.8 

3C1.7 

(°C) 

180 

98 

63.4 

38.8 

28.7 

Atomic  volume,  solid 

(ml/mole  of  atoms) 

13.0 

23.7 

45.4 

55.8 

70.0 

Density  of  solid  at  20°C 

0.535 

0.971 

0.862 

1.53 

1.90 

SEC.    6-3    I    THE    ALKALIES 


95 


SOO'K- 


(a)  m.p,  "K 


■--in 


1.500°K- 


1.000'K- 


500"K 


— fir— 

He   It     Ne  Ml    Ar  K     Kr  Rb     Xe   Cs 


(b)  b.P..  °K 


60 
40V 


He  Li     Ne   Ml    Ar  K    Kr    Rb    Xe    Cs 


(c)  Atomic  volume, 
ml/tnole 


20  Y 


He   Li    Ne    Ma  Ar   K     Kr   Rb   Xe    Cs 

Fig.  6-7.  Trends    in    the    physical    properties    of    the 
alkalies. 


almost  800°  higher.  Contrast  neon,  which  melts 
at  24.6°K  and  boils  only  2.6°  higher,  at  27.2°K. 
How  different  are  the  alkali  metals  from  the 
inert  gases,  only  one  atomic  number  removed! 

ATOMIC    VOLUME 

The  atomic  volumes  of  the  alkali  metals  increase 
with  atomic  number,  as  do  those  of  the  inert 
gases.  Notice,  however,  that  the  volume  occupied 
by  an  alkali  atom  is  somewhat  larger  than  that 
of  the  adjacent  inert  gas  (with  the  exception  of 
the  lithium  and  helium — helium  is  the  cause  of 
this  anomaly).  The  sodium  atom  in  sodium  metal 
occupies  30%  more  volume  than  does  neon. 
Cesium  occupies  close  to  twice  the  volume  of 
xenon. 

6-3.2    Chemistry  of  the  Alkali  Metals 

The  alkali  metals  are  exact  opposites  of  the  inert 
gases  in  chemical  reactivity.  These  metals  react 


vigorously  when  in  contact  with  oxygen  and 
chlorine,  and  even  with  such  a  placid  reagent  as 
water.  Let  us  investigate  some  of  these  reactions. 

REACTIONS    OF   THE    ALKALI    METALS 
WITH    CHLORINE 

When  chlorine  gas  is  brought  into  contact  with 
sodium  metal,  sodium  chloride  is  formed: 

Nafsj  +  \C\,(g)  — ►-  NaClfsJ  +  energy     (4) 

Sodium  chloride  has  a  lattice  built  up  of  so- 
dium ions,  Na+,  and  chloride  ions,  Cl~.  There- 
fore, the  reaction  (4)  involves  the  transfer  of 
electrons  from  sodium  atoms  to  chlorine  atoms. 
The  resulting  ions  attract  each  other  because 
they  have  opposite  electric  charges.  Of  course,  it 
takes  some  energy  to  remove  an  electron  from  a 
sodium  atom  to  form  a  sodium  ion.  However, 
it  takes  only  a  moderate  amount  of  energy  be- 
cause the  Na+  ion  that  is  formed  has  the  electron 
population  of  an  inert  gas,  neon.  This  electron, 
removed  from  a  sodium  atom,  is  then  added  to 
a  chlorine  atom  to  form  a  chloride  ion,  Cl~.  This 
reaction  releases  a  small  amount  of  energy — the 
Cl~  ion  also  has  the  electron  population  of  an 
inert  gas,  argon.  At  a  rather  low  net  cost  of  en- 
ergy, an  electron  can  be  removed  from  a  sodium 
atom  and  transferred  to  a  chlorine  atom,  giving 
Na+  and  Cl~  ions.  Now  these  two  ions  can  move 
toward  each  other  with  a  large  reduction  in  total 
energy.  The  stability  of  the  sodium  chloride 
crystal  can  be  said  to  depend  upon  the  electrical 
attraction  of  the  oppositely  charged  ions.  The 
crystal  is  said  to  be  held  together  by  ionic  bonds. 

This  chemistry  is  characteristic  of  all  of  the 
alkali  metals.  Each  of  them  reacts  with  chlorine 
gas  in  a  similar  way: 


Li(s)  +hC\2(g) 
Nafsj  +  iCUfcl 
K(s)  +  \C\,(g) 
Rb(s)  +  \CUg) 
Cs(s)  +  iCUg) 


LiClfs,)   +  energy  (5) 

NaClfsJ  -f  energy  (6) 

KClfsJ    +  energy  (7) 

RbClfsJ  +  energy  (8) 

CsC\(s)  +  energy  (9) 


In  every  case,  the  alkali  metal  reacts  to  form 
a  stable,  ionic  solid  in  which  the  alkali  is  present 
as  an  inert  gas-like  ion.  The  product  is,  in  each 
case,  a  crystalline  substance  with  high  solubility 
in  water. 


96 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


REACTIONS    OF    THE    ALKALI    METALS 
WITH    WATER 

Sodium  metal  reacts  vigorously  with  water  to 
form  hydrogen  gas  and  an  aqueous  solution  of 
sodium  hydroxide,  NaOH: 

2Na(s)  +  2H20  — *- 

2Na+(aq)  +  20H-(aq)  +  H2(g)  +  energy    (10) 

Energy  is  liberated  and  the  reaction  often  takes 
place  so  rapidly  that  the  temperature  rises  and 
the  hydrogen,  mixing  with  air,  explodes.  Thus, 
sodium  metal  is  dangerous  and  must  be  handled 
with  caution.  This  chemistry  is  also  character- 
istic of  all  of  the  alkali  metals. 

EXERCISE  6-1 

Write  the  equations  for  the  reactions  between 
water  and:  lithium,  potassium,  rubidium,  ce- 
sium. 


Again,  we  see  that  the  alkali  metals  display 
likeness  in  their  reactions  with  water.  Further- 
more, the  reaction  products  always  include  an 
aqueous  ion  of  the  alkali  element  in  which  one 
electron  has  been  removed,  giving  a  1+  ion. 


SUMMARY    OF    CHEMISTRY 
OF   THE   ALKALI    ELEMENTS 

The  alkali  metals  are  extremely  reactive.  Thus, 
there  is  a  dramatic  change  in  chemistry  as  we 
pass  from  the  inert  gases  to  the  next  column  in 
the  periodic  table.  The  chemistry  of  the  alkali 
metals  is  interesting  and  often  spectacular.  Thus, 
these  metals  react  with  chlorine,  water,  and  oxy- 
gen, always  forming  a  +1  ion  that  is  stable  in 
contact  with  most  substances.  The  chemistry  of 
these  +1  ions,  on  the  other  hand,  is  drab,  re- 
flecting the  stabilities  of  the  inert  gas  electron 
arrangements  that  they  have  acquired. 


6-4    THE  HALOGENS 

Now  let's  slide  to  the  left  in  the  periodic  table 
and  consider  the  column  of  elements  fluorine, 
chlorine,  bromine,  iodine,  and  astatine.  Each  of 
these  elements  has  one  less  electron  than  does 
its  neighboring  inert  gas.  These  elements  are 
called  the  halogens.  (The  discussion  that  follows 
does  not  include  astatine  because  this  halogen  is 
very  rare.) 

Fig.  6-8.  The  halogens. 


2 

He 

3 

Li 

fluorine 

10 
Ne 

11 
Na 

chlorine      Kjfl 

16 
Ar 

19 

K 

hromtne      Wjj£ 

36 
Kr 

37 

Rb 

iodine          ■!"■ 

5+ 

Xe 

55 

Cs 

astatine    WWM 

86 
Kn 

87 

Ft- 

6-4.1    Physical  Properties  of  the  Halogens 

Table  6-VI  lists  some  properties  of  the  halogens. 
In  the  elemental  state,  the  halogens  form  stable 
diatomic  molecules.  This  stability  is  indicated  by 
the  fact  that  it  takes  extremely  high  temperatures 
to  disrupt  halogen  molecules  to  form  the  mona- 
tomic  species.  For  example,  it  is  known  that  the 
chlorine  near  the  surface  of  the  sun,  at  a  tem- 
perature near  6000°C,  is  present  as  a  gas  consist- 
ing of  single  chlorine  atoms.  At  more  normal 
temperatures,  chlorine  atoms  react  with  each 
other  to  form  molecules: 

2C\(g)  — *-  Cl2(g)  (11) 

Then,  no  further  reactions  among  chlorine  mole- 
cules occur. 

Apparently  the  diatomic  molecules  of  the  hal- 
ogens already  have  achieved  some  of  the  stability 
characteristic  of  the  inert  gas  electron  arrange- 
ment. How  is  this  possible?  How  could  one 
chlorine  atom  satisfy  its  need  for  one  more  elec- 
tron (so  it  can  reach  the  argon  stability)  by 


SEC.    6-4    I    THE    HALOGENS 


97 


Table  6- VI.     some  properties  of  the  halogens 


PROPERTY 

FLUORINE 

CHLORINE 

BROMINE 

IODINE 

ASTATINE 

Atomic  number 

9 

17 

35 

53 

85 

Atomic  weight 

19.0 

35.5 

79.9 

127 

Molecular  formula 

F2 

Ch 

Br2 

I2 

Boiling  point  (°K) 

85 

238.9 

331.8 

457 

(°C) 

-188 

-34.1 

58.8 

184 

Melting  point  (°K) 

55 

172 

265.7 

387 

(°C) 

-218 

-101 

-7.3 

114 

Atomic  volume,  solid 

(ml/mole  of  atoms) 

14.6 

18.7 

23.5 

25.7 

combining  with  another  chlorine  atom,  an  atom 
with  a  similar  need?  We  answer  this  question  by 
suggesting  that  the  two  atoms  share  two  elec- 
trons, each  atom  contributing  one  electron.  If 
the  two  atoms  huddle  close  together  and  place 
this  communal  pair  of  electrons  between  them, 
each  atom  acts  more  as  though  it  had  the  special 
stability  of  the  inert  gas.  This  results  in  the  for- 
mation of  a  stable  aggregate  of  atoms,  a  mole- 
cule. Its  formula  is  Cl2.  The  same  argument  can- 
be  made  to  explain  the  diatomic  molecular  for- 
mulas of  the  other  halogens.  Because  each  of 
these  molecules  is  bonded  by  a  shared  pair  of 
electrons,  the  bond  is  called  a  covalent  bond. 

BOILING    POINT   AND    MELTING    POINT 

We  have  already  made  a  comparison  between 
the  physical  properties  of  some  of  the  halogens 
and  those  of  inert  gases.  The  comparison  sug- 
gests that  after  the  formation  of  a  diatomic 
molecule,  the  bonding  capacity  of  the  two  halo- 
gen atoms  is  "used  up."  The  only  attractive 
interactions  remaining  between  two  such  dia- 
tomic molecules  are  of  the  extremely  weak 
variety  that  account  for  the  liquefaction  of  the 
inert  gases.  Thus,  the  melting  points  rise  as 
atomic  number  increases  (remember  that  the 
alkali  metals  did  the  reverse)  and  there  is  but  a 
narrow  temperature  range  over  which  the  liquid 
is  stable.  Fluorine  and  chlorine  are  gases  under 
normal  conditions,  bromine  is  a  liquid,  and 
iodine  is  a  solid.  These  differences  in  physical 
state  result  from  the  accident  that  normal  condi- 
tions fall  where  they  do.  On  a  planet  with  a 
"normal"  temperature  of  25°K,  all  of  the  halo- 


gens would  be  solids  whereas  neon  would  be  a 
liquid,  helium  a  gas,  and  argon,  krypton,  and 
xenon  would  be  solids. 


Fig.  6-9.  Trends    in    the    physical    properties    of   the 
halogens. 


500°K 


(a)  m.f>.,     A" 


F  a 

Ne  Ar 

Na  K 


1500°K 


1,000°K 


500°K 


(b)  b.  p.,  °K 


F 

a 

Br 

I 

Ne 

Ar 

Kr 

Xe 

Na 

K 

Kb 

Cs 

60 


40 


20  - 


He 
Li 

(c)  Atomic  volume, 
ml/mole 


F 

a 

Br 

T 

He 

Ne 

Ar 

Kr 

Xe 

Li 

Na 

K 

Rb 

t 

Cs 


98 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE   I    CHAP.    6 


ATOMIC    VOLUMES 

Here  we  find  a  continuation  of  the  trend  dis- 
played by  the  inert  gases  and  alkali  metals.  Com- 
pare the  atomic  volumes  of  the  three  adjacent 
elements  in  the  solid  state: 


fluorine 

14.6  ml 

chlorine 

18.7  ml 

bromine 
23.5  ml 


neon 
20.2  ml 

argon 
24.2  ml 

krypton 
41.9  ml 


sodium 

23.7  ml 

potassium 
45.4  ml 

rubidium 

55.8  ml 


In  each  set,  the  atomic  volumes  increase  going 
from  halogen  to  inert  gas  to  alkali  metal,  as 
shown  graphically  in  Figure  6-9c.  Figure  6-10 
shows  models  constructed  on  the  same  scale  to 
show  the  relative  sizes  of  atoms  indicated  by  the 
atomic  volumes  and  by  the  packing  of  the  ions 
in  the  ionic  solids. 


Figure  6-10.  Models    (to    scale)    of    halogen    atoms, 
inert  gas  atoms,  and  alkali  atoms. 


6-4.2    Chemistry  off  the  Halogens 

The  reactions  of  the  alkali  metals  with  chlorine 
were  used  to  display  the  similarities  of  the  alkali 
metals.  In  a  similar  way,  the  reactions  of  the 
halogens  with  one  of  the  alkali  metals,  say  so- 
dium, show  similarity  within  this  group.  The 
reactions  that  occur  are  as  follows. 


Nafsj  +  \¥,(g) 
Nafsj  +  \C\2(g) 
NafsJ  +  hBr2(g) 
NafsJ  +  \h(g) 


NaFfsJ   +  energy  (12) 

NaClfsj  +  energy  (13) 

NaBrfsJ  +  energy  (14) 

Nal(s)    +  energy  (15) 


These  reactions  all  proceed  readily  and  they 
produce  ionic  solids  with  the  general  empirical 
formula,  NaX  Each  of  these  solids  has  a  crystal 
structure  made  up  of  positively  charged  sodium 
atoms  and  negatively  charged  halogen  atoms. 
These  negative  ions,  F-,  CI-,  Br-,  and  I-,  are 
called  halide  ions.  The  stabilities  of  these  ions 
can  be  related  to  the  stabilities  of  the  correspond- 
ing inert  gas  electron  arrangement. 

The  halogens  also  react  with  hydrogen  gas  to 
form  the  hydrogen  halides: 


Halogens 
Molecules  —1  Ions 


Inert  &ases 
Atoms 


Alkali  Elements 
+1  Ions  Metallic  Atoms 


A 


Ne 


Ma 


Ka  (in  metal) 


*m 


ci. 


Cl 


Ar 


K+  K  (in  metal) 


Br, 


(m 


Kr 


Rb^  Rb  (in  metal) 


SEC.    6-5    I    HYDROGEN  —  A    FAMILY    BY    ITSELF 


99 


Ht(g)  +  F2(g) 
HJg)  +  C\,(g) 
H2(g)  +  Brjg) 
Wg)  +  h(g) 


2UF(g) 

hydrogen  fluoride     (76) 
2HC\(g) 

hydrogen  chloride     (17) 

2HBr(g) 

hydrogen  bromide    (18) 
2Hl(g) 

hydrogen  iodide        (19) 


None  of  these  reactions,  (76),  (17),  (18),  or 
(19),  proceeds  readily  at  room  temperatures. 
This  is  because  the  bonds  holding  the  atoms  to- 
gether in  the  hydrogen  and  in  the  halogen  mole- 
cules must  be  broken  if  new  bonds  are  to  form 
between  hydrogen  atoms  and  halogen  atoms. 
The  breaking  of  the  bonds  is  favored  by  high 
temperatures,  however,  and,  once  started,  these 
reactions  tend  to  proceed  rapidly  or  even  ex- 
plosively. 

6-4.3    Chemistry  off  the  Halide  Ions 

Because  of  the  stabilities  of  halides,  most  ele- 
ments form  stable  halide  compounds.  Thus  cal- 
cium forms  the  compounds  CaF2,  CaCl2,  CaBr2, 
and  Cal2,  all  ionic  solids.  In  each  crystal,  the 
calcium  ion  carries  a  +2  charge,  and  each  of  the 
halide  ions  carries  a  —  1  charge.  The  empirical 
formulas  are  all  of  the  type  Ca.X2. 

The  alkali  halides  are  relatively  unreactive 
substances.  They  all  display  high  solubility  in 
water  and  quite  low  solubility  in  ethyl  alcohol. 

We  have  seen  in  Experiment  8  that  silver  chlo- 
ride has  low  solubility  in  water.  This  is  also  true 
for  silver  bromide  and  silver  iodide.  In  fact,  these 
low  solubilities  provide  a  sensitive  test  for  the 
presence  of  chloride  ions,  bromide  ions,  and 
iodide  ions  in  aqueous  solutions.  If  silver  nitrate 


solution,  which  contains  silver  ions,  Ag+faqJ, 
and  nitrate  ions,  N03"  (aq),  is  added  to  a  solution 
containing  l~(aq)  ions,  a  yellow  precipitate  of 
Agl(s)  forms: 


Ag+(aq)  +  l-(aq) 


AglfsJ 
yellow 


If  Br-  ions  are  present,  the  reaction  is 

Ag+faqJ  +  Br-(aq)  — >-   AgBrfsJ 

light  yellow 

and  if  Cl~  ions  are  present,  the  reaction  is 


Ag+(aq)  +  C\-(aq) 


AgClfsJ 

white 


(20) 


(21) 


(22) 


Silver  fluoride  is  soluble.  Therefore,  no  precipi- 
tate forms  as  Ag+  ions  are  added  to  a  solution 
of  F~  ions. 

All  the  hydrogen  halides  are  gaseous  at  room 
temperature  but  hydrogen  fluoride  liquefies  at 
19.9°C  and  1  atmosphere  pressure.  The  most 
important  chemistry  of  the  hydrogen  halides 
relates  to  their  aqueous  solutions.  All  of  the 
hydrogen  halides  dissolve  in  water  to  give  solu- 
tions that  conduct  electric  current,  suggesting 
that  ions  are  present.  The  reactions  may  be 
written: 


HF(g)  +  water 
HCl(g)  +  water 
HBr(g)  +  water 
Hl(g)    +  water 


H+(aq)  +  F-(aq)  (23) 

H+(aq)  +  C\-(aq)  (24) 

H+(aq)  +  Bx-(aq)  (25) 

H+(aq)  +  l-(aq)  (26) 


These  solutions  have  similar  properties  and  are 
called  acid  solutions.  The  common  species  in  the 
solutions  is  the  aqueous  hydrogen  ion,  H+(aq), 
and  the  properties  of  aqueous  acid  solutions  are 
attributed  to  this  ion.  We  shall  investigate  these 
solutions  in  Chapter  11. 


6-5     HYDROGEN— A  FAMILY  BY  ITSELF 

Perhaps  you  are  wondering  why  the  element 
hydrogen  was  not  included  among  the  halogens. 
It  is,  after  all,  an  element  with  one  less  electron 
than  its  neighboring  inert  gas,  helium.  On  the 
other  hand,  the  hydrogen  atom  has  but  one  elec- 


tron and,  in  a  sense,  it  is  like  an  alkali  metal. 
The  removal  of  one  electron  from  an  alkali  metal 
atom  leaves  a  specially  stable  electron  popula- 
tion— that  of  an  inert  gas.  The  removal  of  one 
electron  from  a  hydrogen  atom  leaves  it  with  no 


100 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


electrons,  which  also  turns  out  to  be  specially 
stable.  We  shall  see  both  of  these  influences  in 
the  chemistry  of  hydrogen.  This  element  forms 
a  family  by  itself,  one  having  some  similarities 
to  the  halogens  and  some  similarities  to  the 
alkalies. 


6-5.1    Physical  Properties 

Hydrogen  is  a  diatomic  gas  at  normal  conditions. 
Its  melting  point  is  15.9°K  and  its  normal  boiling 
point  is  20.4°K.  This  is  the  second  lowest  boiling 
point  of  any  element.  Table  6-VII  lists  these 
properties. 


Table  6-VII 

SOME    PROPERTIES    OF    HYDROGEN 


Atomic  number 

1 

Atomic  weight 

1.008 

Molecular  formula 

H2 

Boiling  point  (°K) 

20.4 

(°Q 

-252.8 

Melting  point  (°K) 

14.0 

(°C) 

-259.2 

Atomic  volume,  solid 

(ml/mole  of  atoms) 

13.1 

We  see  that  the  physical  properties  of  hydro- 
gen are  like  those  of  the  halogens.  Hydrogen  is 
a  diatomic  gas,  like  the  halogens,  rather  than  a 
metal,  like  the  alkalies.  Its  melting  point  is  very 
low  and  it  has  a  narrow  temperature  range  over 
which  the  liquid  is  stable.  However,  the  family 
relationships  among  the  elements  are  based  upon 
their  chemistry,  so  we  must  investigate  the  reac- 
tions of  hydrogen  before  classifying  this  unique 
element. 


6-5.2    Some  Chemistry  of  Hydrogen 

One  of  the  most  distinctive  reactions  charac- 
terizing both  the  alkalies  and  the  halogens  is  their 
reaction  with  each  other.  The  example  we  have 
discussed  most  is  the  reaction  between  sodium 
and  chlorine  to  sive  sodium  chloride.  Sodium 


chloride  is  an  ionic  solid  that  dissolves  in  water 
to  give  positively  charged  sodium  ions,  Na+(aq), 
and  negatively  charged  chloride  ions,  C\~(aq): 


NafsJ  +  \Ck(g) 
NaC\(s)  +  water 


NaClfs;  (27) 

Na+(aq)  +  Cl-(aq)     (28) 


Does  hydrogen  react  like  sodium  or  chlorine 
in  reaction  (27)1  Experiments  show  that  hydro- 
gen can  take  either  position  in  reaction  (27): 


Nafsj  +  \Wg) 

\Wg)  +  hch(g) 

HC\(g)  +  water 


NaHfsJ  (29) 

HCl(g)  (30) 

H+(aq)  +  C\-(aq)     (31) 


The  compound  sodium  hydride,  formed  in 
reaction  (29),  is  a  crystalline  compound  with 
physical  properties  similar  to  those  of  sodium 
chloride.  The  chemical  properties  are  very  dif- 
ferent, however.  Whereas  sodium  burns  readily 
in  chlorine,  it  reacts  with  hydrogen  only  on  heat- 
ing to  about  300°C.  While  sodium  chloride  is  a 
stable  substance  that  dissolves  in  water  to  form 
Na+(aq)  and  C\~(aq),  the  alkali  hydrides  bum 
in  air  and  some  of  them  ignite  spontaneously. 
In  contact  with  water,  a  vigorous  reaction  oc- 
curs, releasing  hydrogen: 

NaH(s)  +  H20  — *- 

H2(g)  +  Ua+(aq)  +  OH-(aq)    (32) 

Thus,  in  reaction  (29),  hydrogen  reacts  with 
sodium  like  a  halogen  [as  in  reaction  (27)],  but 
the  product,  sodium  hydride,  is  very  different  in 
its  chemistry  from  sodium  chloride. 

Reaction  (30)  shows  hydrogen  acting  like  an 
alkali  metal.  Though  the  product,  hydrogen 
chloride,  is  not  an  ionic  solid  like  sodium  chlo- 
ride, it  does  dissolve  in  water  to  give  aqueous 
ions.  The  formation  of  H+(aq)  and  C\~( aq)  is 
strikingly  like  the  analogous  formation  of 
Na+( aq)  and  C\~(aq)  by  sodium  chloride.  In 
fact,  the  tendency  of  hydrogen  to  form  a  posi- 
tively charged  ion  in  water,  H+(aq),  and  the 
absence  of  any  evidence  for  a  negatively  charged 
ion  in  water,  H~(aq),  is  one  of  the  most  signifi- 
cant differences  between  hydrogen  and  the  halo- 
gens. 

An  overall  view  of  the  chemistry  of  hydrogen 
requires  that  it  be  classified  alone— as  a  separate 


SEC.    6-6    |    THE    THIRD-ROW    ELEMENTS 


101 


1 
H 

2 

He 

3 

Li 

9 

F 

10 
Ne 

11       12 

N*M9 

13      14 

At   Si 

15      16       17      18   1 

P      S     CI    Ar\ 

19 

K 

35 
Br 

36 

Kr 

37 
Kb 

53 
I 

54 

Xe 

55 

Cs 

85 

At 

86 

87 
Fr 

Fig.  6-11.  The  placement  of  the  third-row  elements 

in  the  periodic  table. 


chemical  family.  There  are  some  important  simi- 
larities to  halogens — as  we  have  seen,  it  is  a 


stable  diatomic  gas — but  its  chemistry  is  more 
like  that  of  the  alkalies.  Therefore,  hydrogen  is 
usually  shown  on  the  left  side  of  the  periodic 
table  with  the  alkalies  but  separated  from  them 
to  indicate  its  distinctive  character. 


6-6    THE  THIRD-ROW  ELEMENTS 


The  chemistry  of  the  elements  we  have  examined 
thus  far  in  this  chapter  is  dominated  by  the 
special  stabilities  of  the  inert  gas  electron  popu- 
lations. We  can  expect  to  see  this  same  factor  at 
work  in  the  chemistry  of  the  elements  in  other 
parts  of  the  periodic  table.  We  shall  now  take  an 


excursion  across  a  horizontal  row  of  the  periodic 
table  to  see  the  trend  in  chemistry  as  we  pass 
from  element  to  element.  We  shall  consider  the 
third  row,  which  contains  the  elements  sodium, 
magnesium,  aluminum,  silicon,  phosphorus,  sulfur, 
chlorine,  and  argon. 


Table  6- VIII.  some  properties  of  the  elements  of  the  third  row 

OF    THE    PERIODIC    TABLE 


PROPERTY 

Na 

Mg 

Al 

Si 

P 

S 

CI 

Ar 

Atomic  number 

11 

12 

13 

14 

15 

16 

17 

18 

Atomic  weight 

23.0 

24.3 

27.0 

28.1 

31.0 

32.1 

35.5 

39.9 

Molecular  formula 

metal 

metal 

metal 

network 
solid 

P4 

S8 

Cl2 

Ar 

Boiling  point  (°K) 

1162 

1393 

2600 

2628 

553 

718 

238.9 

87 

(°C) 

889 

1120 

2327 

2355 

280 

445 

-34.1 

-186 

Melting  point  (°K) 

371 

923 

933 

1683 

317.2 

392 

172 

84 

(°C) 

98 

650 

660 

1410 

44.2 

119 

-101 

-189 

Atomic  volume,  solid 

(ml/mole  of  atoms) 

23.7 

14.0 

9.99 

12.1 

16.9 

15.6 

18.7 

24.2 

102 


STRUCTURE    OF    THE    ATOM    AND    THE,  PERIODIC    TABLE    I    CHAP.    6 


6-6.1    Physical  Properties  of  the 
Third-row  Elements 

Table  6-VII1  presents  some  properties  of  the 
elements  we  are  considering.  The  first  three, 
sodium,  magnesium,  and  aluminum,  are  metal- 
lic. The  melting  points  and  boiling  points  are 
high  and  increase  as  we  go  from  element  to 
element.  This  trend  reflects  stronger  and  stronger 
bonding  and  it  is  paralleled  by  a  decrease  in  the 
atomic  volume. 

The  fourth  element,  silicon,  forms  a  solid  in 
which  each  silicon  atom  is  bonded  to  four  neigh- 
boring silicon  atoms  placed  equidistant  from 
each  other.  (This  arrangement  places  the  four 
neighbors  at  the  four  corners  of  a  regular  tetra- 
hedron.) This  arrangement  generates  a  three- 
dimensional  network,  hence  is  called  a  network 
solid.  Typically  such  a  network  solid  has  a  high 
melting  point  and  a  high  boiling  point. 

The  remaining  four  elements  form  molecular 
solids.  The  atoms  of  white  phosphorus,  sulfur, 
and  chlorine  are  strongly  bonded  into  small 
molecules  (formulas,  P4,  S8,  and  Cl2,  respectively) 
but  only  weak  attractions  exist  between  the 
molecules.  The  properties  are  all  appropriate  to 
this  description.  Of  course  there  is  no  simple 
trend  in  the  properties  since  the  molecular  units 
are  so  different. 


6-6.2    Compounds  of  the  Third-Row  Elements 

To  see  the  trend  in  chemistry  as  we  move  across 
the  periodic  table,  we  will  consider  three  types 


of  compounds:  the  hydrides,  the  chlorides,  and 
the  oxides. 

HYDRIDES 

The  hydrides  are  the  compounds  formed  with 
hydrogen.  Hot,  molten  sodium  metal  reacts 
with  hydrogen  gas  to  form  a  solid,  salt-like  hy- 
dride having  the  empirical  formula,  NaH.  The 
ions  in  the  salt  are  thought  to  be  the  Na+  ion 
and  the  H~  ion.  The  H~  ion  may  exist  in  the  solid 
or  its  melt  but  it  does  not  exist  in  a  water  solu- 
tion. Magnesium  forms  a  similar  salt-like  hydride 
having  the  empirical  formula  MgH2.  Apparently 
two  electrons  per  neutral  magnesium  atom  are 
removed  to  form  the  Mg+2  ion  (with  the  stable, 
neon-like  electron  arrangement). 

Aluminum  forms  a  hydride  which  seems  to  be 
more  molecular  in  character  than  it  is  salt-like. 
The  empirical  formula  is,  however,  A1H3.  The 
remaining  elements  form  hydrides  which  are 
definitely  molecular,  gaseous  compounds.  The 
formulas  are,  respectively,  SiH4,  PH3,  H2S,  and 
HC1.  These  formulas  are  shown  in  Table  6-IX, 
together  with  the  ratio  of  number  of  atoms  of 
hydrogen  per  atom  of  the  element  (H/M).  We 
see  the  regularity  of  the  trend  in  combining  ca- 
pacity of  the  elements  (as  reflected  in  the  ratio, 
H/M).  On  the  left  side,  the  elements  sodium  and 
magnesium  (and,  to  some  extent,  aluminum) 
achieve  the  neon-like  electron  arrangement  by 
surrendering  electrons  to  hydrogen  atoms,  one 
electron  to  each  hydrogen  atom.  Note  that  each 
hydrogen  atom  thereby  achieves  the  helium-like 
electron  structure.   From  silicon  onward,  the 


Table  6-IX.  the  formulas  of  some  compounds  of  the  third-row  elements 

Na  Mg  Al  Si  P  S  CI  Ar 


Hydrides 

Formula 

NaH 

MgH2 

A1H3 

SiH4 

PH, 

H2S 

HC1 

— 

H/M 

1 

2 

3 

4 

3 

2 

1 

0 

Chlorides 

Formula 

NaCl 

MgCl2 

A12CU 

SiCU 

PCU,  PCI, 

SjCl, 

Cl2 

— 

Cl/M 

1 

2 

3 

4 

5,  3 

1 

1 

0 

Oxides 

Formula 

Na20 

MgO 

A120, 

Si02 

P4OU 

SO, 

Cl207,  ci2o 

— 

2(0/M) 

1 

2 

3 

4 

5 

6 

7.  1 

0 

SEC.    6-7    I    THE    PERIODIC    TABLE 


103 


molecular  compounds  indicate  that  the  elements 
are  achieving  the  next  higher  inert  gas  electron 
arrangement  (argon)  by  sharing  electrons  with 
hydrogen  atoms.  For  these  compounds,  the 
bonding  to  the  hydrogen  atoms  resembles  more 
the  covalent  bonding  we  find  in  Cl2  than  the 
ionic  bonding  we  find  in  sodium  chloride.  What- 
ever the  nature  of  the  bond,  however,  the  com- 
bining capacity  of  each  element  is  influenced  by 
the  tendency  to  form  the  next  lower  or  next 
higher  inert  gas  electron  population. 

CHLORIDES    AND    OXIDES 

We  can  find  the  same  sort  of  trends  in  the  com- 
bining capacity  by  examining  the  chlorides  and 
the  oxides.*  Again,  we  find  the  bonding  can  be 
understood  in  terms  of  electron  transfer  or  elec- 
tron sharing,  in  either  case,  to  reach  an  inert  gas 
electron  arrangement.  There  is  only  one  apparent 
anomaly — the  combining  capacity  of  sulfur  in 
the  compound  SjCl2.  This  is  not  a  real  anomaly, 
however,  for  the  structure  of  this  compound 
reveals  the  expected  combining  power  of  2.  The 


Fig.  6-12.  The  structure  of  S2C1>. 


structure  is  shown  in  Figure  6-12.  The  atoms  in 
this  molecule  are  arranged  so  that  each  sulfur 
atom  reaches  the  argon-like  electron  arrange- 
ment by  sharing  one  pair  of  electrons  with  a 
chlorine  atom  and  one  pair  of  electrons  with 
another  sulfur  atom. 

SUMMARY 

The  simple  trend  in  the  formulas  shown  by 
the  third-row  elements  demonstrates  the  impor- 
tance of  the  inert  gas  electron  populations.  The 
usefulness  of  the  regularities  is  evident.  Merely 
from  the  positions  of  two  atoms  in  the  periodic 
table,  it  is  possible  to  predict  the  most  likely  em- 
pirical and  molecular  formulas.  In  Chapters  16 
and  17  we  shall  see  that  the  properties  of  a 
substance  can  often  be  predicted  from  its  molec- 
ular formula.  Thus,  we  shall  use  the  periodic 
table  continuously  throughout  the  course  as  an 
aid  in  correlating  and  in  predicting  the  properties 
of  substances. 


6-7    THE  PERIODIC  TABLE 


The  power  of  the  periodic  table  is  evident  in  the 
chemistry  we  have  viewed.  By  arranging  the 
elements  in  the  array  shown  on  the  inside  of  the 
front  cover,  we  simplify  the  problem  of  under- 
standing the  variety  of  chemistry  found  in  na- 

*  It  must  be  remarked  that  some  stable  compounds 
have  been  omitted  (for  example,  Na202,  S02).  The  com- 
pounds listed  are  stable  and  display  the  trends  in  bonding 
capacity. 


ture.  The  elements  grouped  in  a  vertical  column 
have  pronounced  similarities.  General  state- 
ments can  be  made  about  their  chemistries  and 
the  compounds  they  form.  Furthermore,  the 
formulas  of  these  compounds  and  the  nature  of 
the  bonds  that  hold  them  together  can  be  under- 
stood in  terms  of  the  special  stabilities  of  the 
inert  gases. 
The  periodicity  of  chemical  properties  was  dis- 


104 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


covered  about  one-hundred  years  ago.  J.  W. 
Dobereiner  (a  German  chemist)  in  1828  recog- 
nized similarities  among  certain  elements  (chlo- 
rine, bromine,  and  iodine;  lithium,  sodium,  and 
potassium;  etc.)  and  he  grouped  them  as  "tri- 
ads." (Remember,  "Cylindrical  objects  burn?") 
J.  A.  R.  Newlands  (an  English  chemist)  in  1864 
was  ridiculed  for  proposing  a  "law  of  octaves" 
which  foresaw  the  differences  of  eight  that  we 
noted  in  Table  6-IV.  Simultaneously,  Lothar 
Meyer  (a  German  chemist  and  physicist)  pro- 
posed a  periodic  table  similar  to  that  of  New- 


lands.  Independently  and  in  this  same  year  (the 
time  was  ripe  for  the  next  step,  "Wooden  objects 
burn")  D.  I.  Mendeleev  (a  Russian)  framed  the 
periodic  table  in  more  complete  form.  He  even 
predicted  both  the  existence  and  properties  of 
elements  not  then  known.  The  subsequent  dis- 
covery of  these  elements  and  corroboration  of 
their  properties  solidified  the  acceptance  of  the 
periodic  table.  It  remains,  one-hundred  years 
later,  the  most  important  single  correlation  of 
chemistry.  It  permits  us  to  deal  with  the  great 
variety  we  find  in  nature. 


QUESTIONS  AND  PROBLEMS 


1.  For  which  of  the  following  processes  will  energy 
be  absorbed? 

(a)  Separating  an  electron  from  an  electron. 

(b)  Separating  an  electron  from  a  proton. 

(c)  Separating  a  proton  from  a  proton. 

(d)  Removing  an  electron  from  a  neutral  atom. 

2.  Which  of  the  following  statements  is  FALSE? 
The  atoms  of  oxygen  differ  from  the  atoms  of 
every  other  element  in  the  following  ways: 

(a)  the  nuclei  of  oxygen  atoms  have  a  different 
number  of  protons  than  the  nuclei  of  any 
other  element; 

(b)  atoms  of  oxygen  have  a  higher  ratio  of  neu- 
trons to  protons  than  the  atoms  of  any  other 
element; 

(c)  neutral  atoms  of  oxygen  have  a  different 
number  of  electrons  than  neutral  atoms  of 
any  other  element; 

(d)  atoms  of  oxygen  have  different  chemical  be- 
havior than  do  atoms  of  any  other  element. 

3.  For  every  atom,  less  energy  is  needed  to  remove 
one  electron  from  the  neutral  atom  than  is 
needed  to  remove  another  electron  from  the 
resulting  ion.  Explain. 

4.  List  the  number  and  kind  of  fundamental  par- 
ticles found  in  a  neutral  lithium  atom  that  has  a 
nucleus  with  a  nuclear  charge  three  times  that 
of  a  hydrogen  nucleus  and  with  seven  times  the 
mass. 


5.  The  nucleus  of  an  aluminum  atom  has  a  diame- 
ter of  about  2  X  10-13  cm.  The  atom  has  an 
average  diameter  of  about  3  X  10~8  cm.  Calcu- 
late the  ratio  of  the  diameters. 

6.  Suppose  a  copper  atom  is  thought  of  as  occupy- 
ing a  sphere  2.6  X  10~8  cm  in  diameter.  If  a 
spherical  model  of  the  copper  atom  is  made  with 
a  5.2  cm  diameter,  how  much  of  an  enlargement 
is  this? 

7.  Suppose  an  atom  is  likened  to  bees  flying  around 
their  beehive.  The  beehive  would  be  compared 
to  the  nucleus  and  the  bees  roving  about  the 
countryside  would  be  compared  to  the  electrons 
of  the  atom. 

(a)  If  the  radius  of  the  beehive  is  25  cm,  what 
would  be  the  average  radius  of  the  flight  of 
the  bees  to  maintain  proper  scale  with  the 
atom?  Express  your  answer  in  kilometers. 

(b)  At  any  instant,  where  is  the  concentration  of 
bees  apt  to  be  highest? 

(c)  Describe  qualitatively  the  distribution  of 
bees  around  the  hive  as  a  function  of  direc- 
tion and  of  distance. 

8.  Helium,  as  found  in  nature,  consists  of  two  iso- 
topes. Most  of  the  atoms  have  a  mass  number  4 
but  a  few  have  a  mass  number  3.  For  each 
isotope,  indicate  the: 

(a)  atomic  number; 

(b)  number  of  protons; 

(c)  number  of  neutrons; 

(d)  mass  number; 

(e)  nuclear  charge. 


QUESTIONS    AND    PROBLEMS 


105 


9.  Fill  in  the  blanks  of  the  following  table. 


aluminum  (Al) 
beryllium  (Be) 
bismuth  (Bi) 
calcium  (Ca) 
carbon  (C) 
fluorine  (F) 
phosphorus  (P) 
iodine  (I) 


ATOMIC 
NO. 


PARTICLES  PER  ATOM 


13 
83 

15 

HUH 

III  II 

II  III 

MASS 


PROTONS         ELECTRONS        NEUTRONS  NUMBER 


10.  How  do  isotopes  of  one  element  differ  from  each 
other?  How  are  they  the  same? 

11.  How  much  would  0.754  mole  of  chlorine-35 
atoms  weigh?  How  much  would  0.246  mole  of 
chlorine-37  atoms  weigh?  What  is  the  weight  of 
a  mole  of  "average"  atoms  in  a  mixture  of  the 
above  samples?  What  is  the  atomic  weight  of 
the  naturally  occurring  mixture  of  these  two 
isotopes  of  chlorine? 

12.  What  is  the  significance  of  the  trends  in  the  boil- 
ing points  and  melting  points  of  the  inert  gases 
in  terms  of  attractions  among  the  atoms? 

13.  Why  is  argon  used  in  many  electric  light  lamps? 

14.  Calculate  the  ratio  of  the  number  of  electrons 
in  a  neutral  xenon  atom  to  the  number  in  a 
neutral  neon  atom.  Compare  this  number  to  the 
ratio  of  the  atomic  volumes  of  these  two  ele- 
ments. On  the  basis  of  these  two  ratios,  discuss 
the  effects  of  electron-electron  repulsions  and 
electron-nuclear  attractions  on  atomic  size. 

15.  The  molar  heats  of  vaporization  of  the  inert 
gases  (in  kcal/mole)  are:  He,  0.020;  Ne,  0.405; 
Ar,  1.59;  Kr,  2.16;  Xe,  3.02;  Rn,  3.92.  Using  the 
data  in  Table  6-III,  (p.  91),  plot  the  boiling 
points  (vertical  axis)  against  the  heats  of  vapori- 
zation (horizontal  axis).  Suggest  a  generalization 
based  upon  a  simple  curve  passing  near  the 
plotted  points.  Write  an  equation  for  the 
straight  line  passing  through  the  origin  (that  is, 
through  zero)  and  through  the  point  for  radon. 


16.  Lithium  forms  the  following  compounds:  lith- 
ium oxide,  Li20;  lithium  hydroxide,  LiOH; 
lithium  sulfide,  Li2S.  Name  and  write  the  for- 
mulas of  the  corresponding  sodium  and  potas- 
sium compounds. 

17.  An  alkali  element  produces  ions  having  the  same 
electron  population  as  atoms  of  the  preceding 
inert  gas.  In  what  ways  do  these  ions  differ  from 
the  inert  gases?  In  what  ways  are  they  alike? 

18.  There  is  a  large  difference  between  the  energy 
needed  to  remove  an  electron  from  a  neutral, 
gaseous  sodium  atom  and  a  neutral,  gaseous 
neon  atom: 


NafgJ+  118.4  kcal 
Ne(g)  +  497.0  kcal 


Na+(gJ  +  e~ 
Ne+(g)  +  e~ 


19. 


Explain  how  these  energies  are  consistent  with 
the  proposal  that  the  electron  arrangements  of 
the  inert  gases  are  specially  stable. 

Refer  to  the  halogen  column  in  the  periodic 
table.  How  many  electrons  must  each  halogen 
atom  gain  to  have  an  electron  population  equal 
that  of  an  atom  of  the  adjacent  inert  gas?  What 
property  does  this  population  impart  to  each 
ion? 


20.  How  do  the  trends  in  physical  properties  for  the 
halogens  compare  with  those  for  the  inert  gases  ? 
Compare  boiling  points,  melting  points,  and 
atomic  volumes. 


106 


STRUCTURE    OF    THE    ATOM    AND    THE    PERIODIC    TABLE    I    CHAP.    6 


21.  Use  your  knowledge  of  the  usefulness  of  the 
periodic  table  to  fill  in  the  blank  spaces  in  Table 
6-VI,  p.  97,  under  Astatine.  List  some  chemical 
reactions  expected  for  astatine. 

22.  Chlorine  is  commonly  used  as  a  germicide  in 
swimming  pools.  When  chlorine  dissolves  in 
water,  it  reacts  to  form  hypochlorous  acid, 
HOC1,  as  follows: 

Cl2  +  H20  +=»= 

HOCl(aq)  +  H+(aq)  +  C\-(aq) 

Predict  what  happens  when  bromine,  Br2,  dis- 
solves in  water.  Write  the  equation  for  the  re- 
action. 

23.  Zinc  metal  dissolves  in  a  solution  of  gaseous 
chlorine  in  water  as  follows: 

Zn(s)  +  G2(aq)  — >-  Zn+2(aq)  +  2Q-(aq) 

Zinc  does  not  dissolve  in  a  solution  of  gaseous 
hydrogen  in  water  but  it  does  dissolve  in  an 
aqueous  solution  of  hydrogen  chloride : 

Zn(s)  +  2H+(aq)  +  2C\-(aq)  — >■ 

Zn+2(aq)  +  H2(g)  +  2Cl-(aq) 

Recognizing  that  zinc  metal  must  release  elec- 
trons to  form  Zn+2(aq),  explain  how  these  re- 
actions demonstrate  that  gaseous  hydrogen  does 
not  behave  like  a  halogen. 

24.  Write  the  molecular  formulas  of  the  hydrogen 
compounds  of  the  second-row  elements,  Li,  Be, 
B,  C,  N,  O,  F,  Ne.  Indicate,  for  each  compound, 
the  H/M  ratio. 

25.  Indicate  the  electron  rearrangement  (gain  or 
loss)  in  each  kind  of  atom  assuming  it  attains 
inert  gas-like  electron  structure  in  the  following 
reactions. 

2RbBr 
2CsI 
MgS 
2BaO 


(a)  2Rb  +  Br2 

(b)  2Cs  +  I2 

(c)  Mg  +  S 

(d)  2Ba  +  02 


26.  Which  of  the  following  is  NOT  a  correct  formula 
for  a  substance  at  normal  laboratory  conditions  ? 

(a)  H2S(g)  (d)  NaNefsj 

(b)  CaCl2(s)  (e)  Al203(sj 

(c)  He(g) 


27.  Magnesium  metal  burns  in  air,  emitting  enough 
light  to  be  useful  as  a  flare,  and  forming  clouds 
of  white  smoke.  Write  the  equation  for  the  re- 
action. What  is  the  composition  of  the  smoke? 

28.  Use  the  formulas  for  magnesium  oxide,  MgO, 
and  magnesium  chloride,  MgCl2,  together  with 
the  periodic  table  to  decide  that  magnesium  ions 
have  the  same  number  of  electrons  as  each  of 
the  following,  EXCEPT 

(a)  neon  atoms,  Ne; 

(b)  sodium  ions,  Na+; 

(c)  fluoride  ions,  F~; 

(d)  oxide  ions,  O-2; 

(e)  calcium  ions,  Ca+2. 

29.  Sodium  metal  reacts  with  water  to  form  sodium 
ions,  Na+,  hydroxide  ions,  OH~,  and  hydrogen 
gas,  H2,  as  follows: 

2Na(sJ  +  2H20  — ►■ 

2Na+(aq)  +  20H-(aqj  +  H2(g) 

Assuming  calcium  metal  reacts  in  a  similar  way, 
write  the  equation  for  the  analogous  reaction 
between  calcium  and  water.  Remember  that 
calcium  is  in  the  second  column  of  the  periodic 
table  and  sodium  is  in  the  first. 

30.  Use  Table  6-IX,  p.  102,  and  the  periodic  table 
to  write  possible  formulas  for  the  following  com- 
pounds: 

(a)  a  hydride  of  barium,  element  56; 

(b)  a  chloride  of  germanium,  element  32; 

(c)  an  oxide  of  indium,  element  49; 

(d)  an  oxide  of  cesium,  element  55; 

(e)  a  fluoride  of  tin,  element  50. 

31.  All  of  the  isotopes  of  the  element  with  atomic 
number  87  are  radioactive.  Hence,  it  is  not  found 
in  nature.  Yet,  prior  to  its  preparation  by  nuclear 
bombardment,  chemists  were  confident  they 
knew  the  chemical  reactions  this  element  would 
show.  Explain.  What  predictions  about  this  ele- 
ment would  you  make? 


DMITRI    MENDELEEV,    1834-1907 


Element  101  is  named  Mendelevium  in  honor  of  the  great 
Russian  chemist,  Dmitri  Mendeleev.  The  youngest  of 
seventeen  children,  he  was  born  in  Tobolska  where  his 
grandfather  published  the  first  newspaper  in  Siberia  and 
his  father  was  the  high  school  principal.  Dmitri  received  his 
early  education  from  a  political  exile,  but  when  his  father 
died,  his  mother  traveled  west  in  search  of  better  educa- 
tional opportunities  for  Dmitri. 

At  the  University  of  St.  Petersburg  {now  Leningrad),  he 
distinguished  himself  in  science  and  mathematics  and  earned 
the  doctorate  with  a  thesis  on  a  subject  that  remains  of 
current  interest,  "The  Union  of  Alcohol  and  Water.'1''  Sub- 
sequent studies  in  France  and  Germany  permitted  him  to 
attend  the  1858  Karlsruhe  (Germany)  conference  at  which 
Avogadro's  Hypothesis  was  heatedly  debated.  Later,  he 
visited  the  oilfields  of  Pennsylvania  to  see  the  first  oil  well. 
Upon  his  return  to  Russia,  he  developed  a  new  commercial 
distillation  process. 

He  became  a  professor  of  chemistry  at  St.  Petersburg 


when  only  32.  Searching  for  regularities,  he  arranged  the 
elements  by  their  properties.  This  organization  led  him  to 
propose  the  periodic  table  and  use  it  to  predict  the  existence 
and  properties  of  a  number  of  additional  elements.  When 
some  of  those  that  were  foretold  in  1869  were  actually 
discovered  a  few  years  later,  Mendeleev  was  hailed  as  a 
prophet. 

This  inspiring  teacher  and  tireless  experimenter  was  so 
deeply  concerned  over  social  issues  that  he  resigned  his 
professorship  rather  than  obey  an  order  to  cease  interfering 
with  affairs  of  government.  He  made  enemies  by  supporting 
liberal  movements  and  even  defied  the  Czar's  wishes  by 
refusing  to  cut  his  hair.  Nevertheless,  he  won  the  appoint- 
ment as  Director  of  the  Bureau  of  Weights  and  Measures. 

When  Mendeleev  first  published  his  chart,  there  were 
63  elements  known.  One  year  after  his  death,  there  were  86. 
The  rapidity  of  this  increase  was  made  possible  by  the 
most  important  generalization  of  chemistry,  the  periodic 
table. 


CHAPTER 


7 


Energy  Effects  in 
Chemical  Reactions 


Although  a  typical  chemical  reaction  •  •  •  may  appear  far  removed 
from  the  working  of  an  engine,  the  same  fundamental  principles  of  heat 
and  work  apply  to  both. 

f.  T.  wall,   1958 


Chemical  reactions  form  the  heart  of  chemistry. 
And  there  is  no  more  important  aspect  of  chemi- 
cal reactions  than  the  energy  effects  that  are 
caused.  You  will  realize  this  if  you  let  your 
thoughts  wander  between  the  warmth  the  little 
child  in  the  fable  derived  from  the  combustion 


of  wood  and  the  celestial  joy  ride  an  astronaut 
receives  from  the  reactions  of  his  rocket  fuels. 
How  much  energy  is  involved  in  a  chemical  re- 
action? How  do  we  find  this  out?  Where  does 
this  energy  come  from?  We  shall  investigate 
these  questions  in  this  chapter. 


7-1    HEAT  AND  CHEMICAL  REACTIONS 

At  a  temperature  of  600°C,  steam  passed  over 
hot  coal  (coal  is  mostly  carbon)  reacts  to  give 
carbon  monoxide  and  hydrogen: 

H20(g)  +  C(s)  — >-  CO(g)  +  H2(g)        (7) 

This  reaction  is  quite  useful  because  the  mixture 
of  product  gases,  called  "water  gas,"  is  an  ex- 
cellent industrial  fuel.  In  the  commercial  prepa- 
ration of  water  gas,  the  chemical  engineer  must 
allow  for  the  absorption  of  heat  during  the  reac- 
tion. In  fact,  he  must  periodically  turn  off  the 
steam  and  reheat  the  coal  to  keep  the  reaction 
going.  To  aid  the  engineer,  we  might  measure 
108 


the  amount  of  heat  absorbed  by  the  system  and 
write  it  into  the  chemical  reaction.  Such  a  meas- 
urement shows  that  3 1 .4  kcal  of  heat  are  absorbed 
per  mole  of  carbon  reacted.  Since  the  heat  is 
used  up  (as  are  the  reactants),  we  can  rewrite 
reaction  (7)  and  show  the  heat  on  the  left  side 
of  the  equation: 

H20(g)  +  C(s)  +  31.4  kcal  — >■ 

CO(g)  +  H2(g)    (la) 

Now  we  might  think  of  the  mechanical  engi- 
neer who  is  designing  a  boiler  to  be  heated  with 
water  gas  fuel.  He  is  interested  in  burning  the 


SEC.    7-1    I    HEAT    AND    CHEMICAL    REACTIONS 


109 


water  gas.  This  involves  two  chemical  reactions 
of  combustion: 


and 


co(g)  +  \o,(g)  — >■  co2(g)  (2) 


Wg)  +  \02(g)  —>  H2Q(g)  (3) 


These  reactions  release  heat,  and  our  mechanical 
engineer  wishes  to  know  how  much.  Again,  we 
might  help  by  measuring  these  amounts  of  heat 
and  adding  the  information  to  reactions  (2)  and 
(3).  Since  heat  is  produced  by  the  reaction  (as  is  a 
chemical  product),  we  should  place  it  on  the 
right  side  of  the  equation.  Experiments  show: 


CO(g)  +  \02(g) 
H2(g)  +  \Q2(g) 


C02(g)  +  67.6  kcal    (2a) 
H2Q(g)  +  57.8  kcal    (3a) 


Now  let  us  talk  with  the  business  manager.  He 
thinks  in  terms  of  gains  and  losses.  He  is  likely 
to  observe  that  the  consumption  of  coal  and 
water  to  generate  water  gas  is  followed  by  the 
combustion  of  the  water  gas  to  form  carbon 
dioxide  and  water.  Without  knowing  much 
chemistry,  he  can  see  that  what  is  finally  accom- 
plished is  the  combustion  of  coal  to  form  carbon 
dioxide.  The  overall  reaction  is  obtained  by  add- 
ing reactions  (7),  (2),  and  (3): 


H20(g)  +  C(s) 

CO(g)  +  \o2(g) 

Wg)  +  \02(g) 


CO(g)  +  H2(g)    (1) 
C02(g)  (2) 

H2OteJ  (3) 


Overall 
reaction : 


C(s)  +  02(g)-+C02(g) 


(4) 


The  business  manager  is  frugal  so  he  asks,  "Why 
not  burn  the  coal  directly  and  save  the  cost  ol 
manufacturing  the  water  gas?"  The  mechanical 
engineer  is  practical  so  he  asks,  "How  much  heat 
will  the  boiler  receive  if  I  use  coal  instead  of 
water  gas?"  The  chemical  engineer  goes  to  the 
laboratory  to  find  the  answers  by  measuring  the 
heat  released  per  mole  of  carbon  burned  in  reac- 
tion (4).  The  laboratory  result  shows  that  re- 
action (4)  releases  94.0  kcal/mole: 

C(s)  +  02(g)  -+  C02(g)  +  94.0  kcal    (4a) 

The  chemical  engineer  now  can  answer  all  of  the 
questions.  If  one  mole  of  carbon  is  burned  di- 
rectly, 94.0  kcal  of  heat  are  released  for  the 


mechanical  engineer.  The  same  amount  of  coal 
converted  into  water  gas  releases  the  sum  of  the 
heats  evolved  in  reactions  (2a)  and  (5a): 

67.6  +  57.8  =  125.4  kcal 

The  mechanical  engineer  has  a  better  fuel  in 
water  gas  than  in  coal. 

With  new  respect,  the  business  manager  might 
now  ask  the  chemical  engineer,  "Where  did  this 
extra  heat  come  from?"  "Did  we  get  something 
for  nothing?"  The  answer  is,  "No."  The  water 
gas  releases  more  heat  per  mole  of  carbon  be- 
cause the  chemical  engineer  put  that  amount  of 
heat  in  during  reaction  (7a).  The  business  man- 
ager's ledger  is  shown  in  Table  7-1. 


Table  7-1 

HEAT    EFFECTS    IN    THE    MANUFACTURE 
AND    USE    OF    WATER    GAS 

DEBIT  CREDIT 


Reaction  (7a):  heat  absorbed 

31.4  kcal 

— 

Reaction  (2a):  heat  released 

— 

67.6  kcal 

Reaction  (3a) :  heat  released 

— 

57.8  kcal 

Overall  reaction: 
(la)  +  (2a)  +  (3a)  =  (4a) 

31.4  kcal 
absorbed 

125.4  kcal 
released 

125.4 

-  31.4 

94.0  net 

Experiment  reaction  (4a): 
Heat  released 

94.0  kcal 

7-1.1    Heat  Content  of  a  Substance 

The  example  just  given  shows  that  the  31.4  kcal 
absorbed  during  reaction  (7a)  was  "stored"  in 
the  water  gas.  Furthermore,  the  amount  of  en- 
ergy "stored"  is  definite,  not  alterable  at  the 
demand  of  the  business  manager  or  the  whim 
of  the  chemical  engineer.  How  much  energy  is 
stored  depends  upon  the  reactants  and  products 
in  reaction.  We  must  add  a  fixed  amount  of 
energy  (as  heat)  to  coal  and  steam  to  make  a 
specified  amount  of  carbon  monoxide  and  hy- 
drogen. This  heat  is  retained  by  the  CO  and  H2 
molecules,  as  shown  in  Table  7-1.  We  might  say 
that  reaction  (7a)  increases  the  "heat  content"  of 


110 


ENERGY    EFFECTS    IN    CHEMICAL    REACTIONS    I    CHAP.    7 


the  atoms  of  the  reactants  by  rearranging  them 
to  form  the  products.  Apparently  a  mole  of  each 
molecular  substance  has  a  characteristic  "heat 
content"  just  as  it  has  a  characteristic  mass.  This 
heat  content  measures  the  energy  stored  in  a 
substance  during  its  formation.  The  heat  effect  in 
a  chemical  reaction  measures  the  difference  be- 
tween the  heat  contents  of  the  products  and  the 
heat  contents  of  the  reactants.  If  more  energy  is 
stored  in  the  reactants  than  in  the  products,  then 
heat  will  be  released  during  reaction.  Conversely, 
heat  will  be  absorbed  if  more  energy  is  stored 
in  the  products  than  in  the  reactants. 

This  idea — that  each  molecular  substance  has 
a  characteristic  heat  content — provides  a  good 
explanation  of  the  heat  effects  found  in  chemical 
reactions.  Chemists  symbolize  heat  content  by  H. 
Since  the  heat  effect  in  a  reaction  is  the  difference 
between  the  7/'s  of  the  products  and  the  7/'s  of 
the  reactants,  the  heat  of  reaction  is  called  AH, 
the  Greek  letter  A  (delta)  signifying  "difference." 

We  can  see  what  AH  means  in  terms  of  an 
example.  Consider  reaction  (7): 

H20(g)  +  C(s)  — >-  CO(g)  +  H2(g)        (1) 

A  heat  content,  H,  is  assigned  to  each  substance. 
Then  AH  for  reaction  (7)  is  the  difference : 

Afr  _    I  heat  content  \  /  heat  content  \ 

\  of  products  /  \  of  reactants  / 

=  \Hco  +  7/h«]  —  [7/hjO  +  He] 

=  Hco  4"  7/hi  —  Hhio  —  He 


Products 


Heat 
content 


AH 
positive 


Reactants 


Since  the  reaction  consumes  heat,  the  heat  con- 
tent of  the  products  is  higher  and  AH  will  be 
positive.  We  can  express  this  by  writing 

H20(g)  +  C(s)  — *  CO(g)  +  H2(g) 

AH  =  +31.4  kcal    (lb) 

Reaction  (lb)  is  exactly  equivalent  to  reaction 
Ua): 

UiO(g)  +  C(s)  +  31.4  kcal  — >- 

CO(g)  +  H2(g)    (la) 

Let's  try  this  on  reaction  (2): 

.„  _  /heat  content \  _      /heat  content \ 
\of  products  /  \of  reactants  / 

=       Hco,  —  [Hco  +  T/iOi] 
=  Hcch  —  Hco  —  Hio, 

In  this  reaction,  heat  is  evolved ;  hence,  the  heat 
content  of  the  products  is  below  that  of  the  re- 
actants. Therefore,  AH  must  be  negative: 

CO(g)  +  ±02(g)  -+  C02(g) 

AH  =  -67.6  kcal    (2b) 

This  has  exactly  the  same  meaning  as 

CO(g)  +  \02(g)  — >-  C02(g)  +  67.6  kcal    (2a) 

We  see  that  the  sign  of  AH  is  sensible.  It  is 
positive  when  heat  content  is  rising  (by  heat 
absorption)  and  it  is  negative  when  heat  content 
is  dropping  (by  heat  evolution).  This  is  shown 
diagrammatically  in  Figure  7-1. 

Fig.  7-1.  Heat  content  change  during  a  reaction. 


'Rea.c1ra.nJrs 

Heatr 
Con.iren.tr 

A 

1 

H 
ative 

f 

Products 

Reaction,    proceeding 

Heair  is  ahsorhed 
A  H  is  positive 


Reaction,   proceeding 

Ueatr  is  evotved 
AH  is  negative 


SEC.    7-1    I    HEAT    AND    CHEMICAL    REACTIONS 


111 


7-1.2    Additivity  of  Reaction  Heats 

Let  us  return  to  the  debit  and  credit  balance  we 
found  in  our  water  gas  fuel  problem.  In  terms 
of  AH,  the  heat  effects  are  as  follows: 


Source  of 
direct  current 


H20(g)  +  C(s) 

CO(g)  +  hojg) 
H,(g)  +  \Q2(g) 


CO(g)  +  H2(g) 

AH,  =  +31.4  kcal    (lb) 

C02(g) 

AH,  =  -67.6  kcal    (2b) 

H20(g) 

AHS  =  -57.8  kcal    (3b) 


Overall  reaction  (/)  +  (2)  +  (3)  =  (4): 

Qs)  +  cygj  — ►-  co,(g) 

A//,  =  -94.0  kcal    (4b) 

We  discover  that  not  only  is  reaction  (4)  equal 
to  the  sum  of  reactions  (lb)  +  (2b)  +  (3b)  in 
terms  of  atoms,  but  also  that 

AH4  =  AH,  +  AHt  +  AH, 

=  31.4 +  (-67.6) +  (-57.8) 

=  31.4  -  67.6  -  57.8 

=  -94.0  kcal 

We  see  that  when  a  reaction  can  be  expressed  as 
the  algebraic  sum  of  a  sequence  of  two  or  more 
other  reactions,  then  the  heat  of  the  reaction  is  the 
algebraic  sum  of  the  heats  of  these  reactions.  This 
generalization  has  been  found  to  be  applicable 
to  every  reaction  that  has  been  tested.  Because 
the  generalization  has  been  so  widely  tested,  it 
is  called  a  law — the  Law  of  Additivity  of 
Reaction  Heats.* 

7-1.3    The  Measurement  off  Reaction  Heat 

The  measurement  of  reaction  heats  is  called 
calorimetry — a  name  obviously  related  to  the 
unit  of  heat,  the  calorie.  You  already  have  some 
experience  in  calorimetry.  In  Experiment  5  you 
measured  the  heat  of  combustion  of  a  candle  and 
the  heat  of  solidification  of  paraffin.  Then  in 
Experiment  13  you  measured  the  heat  evolved 
when  NaOH  reacted  with  HC1.  The  device  you 
used  was  a  simple  calorimeter. 

*  This  generalization  was  first  proposed  in  the  year 
1840  by  G.  H.  Hess  on  the  basis  of  his  experimental 
measurements  of  reaction  heats.  It  is  sometimes  called 
Hess's  Law  of  Constant  Heat  Summation. 


Thermometer 


Insulation. 


Water 
Reaction  chamber 


Resistance  wire 
for 
y<  pS   ianitiru/    charge 


Fig.  7-2.  General  plan  of  a  calorimeter. 


Calorimeters  vary  in  details  and  are  adapted 
to  the  particular  reaction  being  studied.  Figure 
7-2  shows  the  general  plan  of  a  calorimeter  that 
might  be  used  in  measuring  the  heat  evolved 
during  a  combustion  reaction.  It  might  be  ap- 
plied to  the  combustion  of  a  candle  to  yield  a 
much  more  reliable  answer  than  can  be  obtained 
by  the  crude  technique  of  Experiment  5. 

Essentially,  the  device  consists  of  an  insulated 
vessel  containing  a  known  weight  of  water.  A 
weighed  amount  of  the  substance  to  be  burned 
and  an  excess  of  oxygen  are  introduced  under 
pressure  into  a  reaction  chamber  placed  in  the 
vessel.  The  reaction  mixture  is  ignited  by  means 
of  an  electrical  resistance  wire  sealed  into  the 
reaction  chamber.  The  heat  produced  by  the 
reaction  changes  the  temperature  of  the  water, 
which  is  stirred  to  keep  its  temperature  uniform. 
From  this  temperature  change  and  from  the 
amount  of  heat  required  to  raise  the  temperature 
of  the  calorimeter  and  its  contents  by  one  degree, 
the  heat  of  combustion  per  mole  of  substance 
burned  can  be  calculated.  By  using  a  large 
amount  of  water,  the  actual  change  in  tempera- 
ture is  kept  small.  This  is  desirable  in  that  it 
keeps  the  final  temperature  of  the  products  of 
reaction  fairly  close  to  the  initial  temperature  of 
the  reactants. 


112 


ENERGY    EFFECTS    IN    CHEMICAL   RBACTIONS   I   CHAP.    7 


EXERCISE  7-1 

Suppose  reactants  are  mixed  in  a  calorimeter  at 
25°C  and  the  reaction  heat  causes  the  tempera- 
ture of  the  products  and  calorimeter  to  rise  to 
35°C.  The  resultant  determination  of  AH  applies 
to  what  temperature?  Explain  why  it  is  desirable 
to  keep  the  final  temperature  close  to  the  initial 
temperature  in  a  calorimetric  measurement. 


In  the  combustion  reaction  as  carried  out  in  the 
calorimeter  of  Figure  7-2,  the  volume  of  the  system  is  kept 
constant  and  pressure  may  change  because  the  reaction 
chamber  is  sealed.  In  the  laboratory  experiments  you 
have  conducted,  you  kept  the  pressure  constant  by  leaving 
the  system  open  to  the  surroundings.  In  such  an  experi- 
ment, the  volume  may  change.  There  is  a  small  difference 
between  these  two  types  of  measurements.  The  difference 
arises  from  the  energy  used  when  a  system  expands 
against  the  pressure  of  the  atmosphere.  In  a  constant 
volume  calorimeter,  there  is  no  such  expansion;  hence, 
this  contribution  to  the  reaction  heat  is  not  present.  Ex- 
periments show  that  this  difference  is  usually  small.  How- 
ever, the  symbol  AH  represents  the  heat  effect  that 
accompanies  a  chemical  reaction  carried  out  at  constant 
pressure — the  condition  we  usually  have  when  the  reac- 
tion occurs  in  an  open  beaker. 


7-1.4    Predicting  the  Heat  of  a  Reaction 

Chemists  have  measured  the  heats  of  many  re- 
actions. With  the  measured  values,  many  un- 


measured reaction  heats  can  be  predicted  by 
applying  the  Law  of  Additivity  of  Reaction 
Heats.  Consequently,  a  compilation  of  known  re- 
action heats  is  extremely  useful.  Table  7-II  is 
such  a  compilation. 

Suppose  we  are  interested  in  the  heat  of  com- 
bustion of  nitric  oxide,  NO: 

NOfej  +  \02(g)  — ►-  N02fg;        AH5    (5) 

Since  reaction  (5)  can  be  obtained  by  combin- 
ing two  reactions  in  Table  7-II,  we  can  predict 
AH5.  In  Table  7-II  we  find 

\K2(g)  +  \02(g)  — >  NO(g) 

AH6  =  +21.6  kcal/mole  NO     (6) 

hK2(g)  +  02(g)    -^N02(g) 

AH7  =    +8.1  kcal/mole  N02    (7) 

Now  we  wish  to  obtain  reaction  (5)  by  combin- 
ing reactions  (6)  and  (7).  Since  NO  is  a  reactant 
in  reaction  (5),  we  need  the  reverse  of  reaction 
(<5).  We  obtain  the  heat  of  the  reverse  reaction 
merely  by  changing  the  algebraic  sign  of  AH6.  If 
21.6  kcal  of  heat  are  absorbed  when  one  mole 
of  NO  is  formed,  then  21.6  kcal  of  heat  will  be 
released  when  one  mole  of  NO  is  decomposed 
in  the  reverse  reaction: 

NO(g)  — >■  ±N2(g)  +  \02(g) 

AH8  =  -21.6  kcal/mole  NO    (8) 

Now  we  can  add  reactions  (7)  and  (8)  to  obtain 
reaction  (5): 


Table  7 -II.     heats  of  reaction  between  elements,  t  =  25°C,  p  =  1  atm 


ELEMENTS 


COMPOUND 


FORMULA 


NAME 


HEAT  OF  REACTION 

(kcal/mole  of  product) 


Wg)  +  hOt(g)  — ►■  H20(gj  water  vapor 

H2(gJ  +  \02(g)  — >-  H20(7J  water 

S(s)  +  02(g)  — ►■  S02(g)  sulfur  dioxide 

H2(g)  +  S(s)  +  202(g)  — >-  H2SOi(l)  sulfuric  acid 

hWg)  +  k02(g)  — ►■  NOteJ  nitric  oxide 

^N2fgJ  +  02(g)  — >-  N02(gj  nitrogen  dioxide 

h^2(g)  +  fHrf*;  — ►  NH3fej  ammonia 

C(s)  +  \02(g)  — ►■  CO(g)  carbon  monoxide 

C(s)  +  02(g)  — >-  C02(g)  carbon  dioxide 

2C(s)  +  3H2(g)  — )-  dH,(g)  ethane 

3C(s)  +  4U2(g)  — >■  CsHs(g)  propane 

*H2(gJ  +  \\t(g)  — >■  Hl(g)  hydrogen  iodide 


-57.8 
-68.3 
-71.0 
-194 
+21.6 

+8.1 
-11.0 
-26.4 
-94.0 
-20.2 
-24.8 

+6.2 


SEC.    7-2    I    THE    LAW    OF    CONSERVATION    OF    ENERGY 


113 


or 


NO(g)  — >  $Wg)  +  \Oi(g) 

hWg)  +  o2(g)  — ►■  no^j 

Overall  reaction: 

NO(g)  +  02(g)  +  h*i2(g)  -+  N02(g)  +  \02(g)  +  iN./gj 

NO(g)  +  \Q2(g)  —*■  NO/gJ 


AH,  =  -21.6  kcal/mole  NO  (5) 

AH7  =    +8.1  kcal/mole  N02  (7) 

AH6  =  -21.6  +  8.1 

AHS  =  -13.5  kcal/mole  NO  (5) 


EXERCISE  7-2 

Predict  the  heat  of  the  reaction 

CO(g)  +  ho2(g)  — +  C02(g) 
from  two  reactions  listed  in  Table  7-II.  Compare 
your  result  with  AHeb  given  in  Section  7-1.2. 

The  usefulness  of  Table  7-II  is  obvious  from 
these  examples.  Many,  many  reaction  heats  can 
be  predicted— in  fact,  the  heat  of  any  reaction 
that  can  be  obtained  by  adding  two  or  more  of 
the  reactions  in  the  table.  Furthermore,  there  is 
an  easy  way  to  decide  whether  the  table  contains 
the  necessary  information  for  a  particular  ex- 
ample. A  given  reaction  can  be  obtained  by  add- 
ing reactions  in  Table  7-II,  provided  every  com- 
pound in  the  reaction  is  included  in  the  table. 
The  elements  participating  in  the  reactions  auto- 
matically will  appear  in  proper  amount. 

Consider  a  more  complicated  example — the 
oxidation  of  ammonia,  NH3: 

KH3(g)  +  i02(g)  — >- 

K02(g)  +  %H20(g)       AH9    (9) 

In  reaction  (9)  we  find  three  compounds,  NH3(gJ, 
N02(g,),  and  H20(g).  These  are  all  found  in 
Table  7-11.  Consequently,  we  are  able  to  calcu- 
late AH9. 


EXERCISE  7-3 


Convince  yourself  that  reaction  9)  and  also 
AH9  =  —67.6  kcal  can  be  obtained  by  carrying 
out  the  indicated  summation: 


From 


Subtract 


\N2(g)  +  02(g)  —+  N02(gj 
AH  =    +8.1  kcal 

hn2(g)  +  § H2(g)  — >  Ntug) 

Subtract        AH  =  -11.0  kcal 


And  now  add  §  times    H2(g)  +  \02(g)  — >■  H20(g) 
Add  |  times  AH  =  -57.8  kcal 


Thus,  when  we  wish  to  predict  the  heat  of 
some  reaction,  it  takes  but  a  moment  to  decide 
whether  the  compilation  of  Table  7-11  includes 
the  necessary  reactions.  If  every  compound  in  the 
reaction  of  interest  is  in  the  "Compound"  col- 
umn of  Table  7-II,  then  the  prediction  can  be 
made.  This  is  a  convenience  provided  by  a  com- 
pilation of  heats  of  reaction  between  elements 
and  explains  why  these  reaction  heats  are  the 
ones  chemists  tabulate.  Of  course,  the  list  in 
Table  7-II  includes  only  a  small  fraction  of  the 
known  values.  Many  more  reaction  heats  are 
tabulated  in  handbooks;  they  are  listed  in  in- 
dexes under  Heat  of  Formation. 


7-2  THE  LAW  OF  CONSERVATION  OF  ENERGY 


We  have  seen  how  chemists  measure  the  heat  of 
a  reaction.  Using  a  compilation  of  measured 
values,  we  can  predict  the  energy  changes  of 
many  reactions  that  have  not  been  measured. 
Thus,  the  rule  of  Additivity  of  Reaction  Heats 


is  a  very  useful  and  reliable  generalization.  It 
makes  us  wonder  "Why  should  it  be  so?"  The 
explanation,  as  usual,  is  found  by  connecting  the 
behavior  of  a  chemical  system  to  the  behavior 
of  other  systems  that  are  better  understood. 


114 


ENERGY    EFFECTS    IN    CHEMICAL    REACTIONS    I    CHAP.    7 


Fig.  7-3.  Conservation  of  energy  in  a  collision  of 
billiard  balls. 


7-2.1    Conservation  off  Energy  in  a  Billiard 
Ball  Collision 

Figure  7-3  shows  an  experimental  study  of  the 
collision  of  hard  spheres.  The  experimenter  im- 
parts energy  of  motion  to  the  white  ball  (see 
Figure  7-3A,  7-3B).  He  does  so  by  doing  work 
by  striking  the  ball  with  the  end  of  a  cylindrical 
stick  (a  cue).  The  amount  of  energy  of  motion 
(kinetic  energy)  received  by  the  ball  is  fixed  by 
the  amount  of  work  done.  If  the  ball  is  struck 
softly  (little  work  being  done),  it  moves  slowly. 
If  the  ball  is  struck  hard  (much  work  being  done), 
it  moves  rapidly.  The  kinetic  energy  of  the  white 
ball  appears  because  work  was  done — the 
amount  of  work,  Wu  determines  and  equals  the 
amount  of  kinetic  energy,  (KE)X.  In  symbols, 


Wx  =  (KE)V 


UO) 


Suppose  the  direction  of  motion  of  the  white 
ball  causes  it  to  contact  the  motionless  red  ball. 
A  collision  occurs.  Figure  7-3C  shows  the  result. 
The  white  ball  has  a  lower  kinetic  energy,  (KE}i, 
but  now  the  red  ball  is  moving!  The  red  ball  now 
has  energy  of  motion — let's  call  it  (KE)3.  Meas- 
urements show  that  velocities  are  such  that  the 


kinetic  energy  gained  by  the  red  ball  is  equal  to 
the  kinetic  energy  lost  by  the  white  ball.  In  sym- 
bols, 

(KE)X  =  (KE)2  +  (KE)3  (11) 

Let  us  review  this  experiment.  First,  an 
amount  of  work  was  performed,  Wx,  with  the 
cylindrical  stick.  The  amount  of  kinetic  energy 
received  by  the  white  ball,  (KE)X,  was  exactly 
fixed  by  Wx.  After  the  white  ball  collided  with 
the  red  ball,  the  sum  of  the  energies  of  the  two 
balls,  (KE}t  +  (KE)3,  is  exactly  equal  to  (#£),. 
We  can  write 

Wx  =  (KE),  =  (KEh  +  (KE)3  (12) 

If  we  recognize  work  as  a  form  of  energy, 
we  may  say  that  energy  was  conserved  during  the 
experiment.  Every  quantity  in  (12)  is  known — 
we  can  measure  the  work  done  as  we  do  it  and 
we  can  learn  the  magnitudes  oi(KE)u  (KE)?,  and 
(KE)3  through  velocity  and  mass  measurements. 
Many  such  experiments  are  performed  every  day 
and  the  results  are  always  in  agreement.  In  a 
billiard  parlor,  energy  is  conserved. 

7-2.2    Conservation  off  Energy  in  a 
Stretched  Rubber  Band 

Figure  7-4  shows  an  experimental  study  of  the 
stretching  of  a  rubber  band.  The  rubber  band  is 


Fig.  7-4.  Conservation  of  energy  in  a  stretched  rubber 
band. 


SBC.    7-2   I   THB    LAW    OF    CONSERVATION    OF    ENERGY 


115 


stretched  and  hooked  over  the  end  of  the  testing 
device,  as  shown  in  Figure  7-4A.  Work  must  be 
done  to  stretch  it— let's  call  it  Wx.  In  Figure  7-4C 
the  rubber  band  has  been  released  and  has  re- 
turned to  its  initial  length,  but  it  now  has  energy 
of  motion,  (KE)3.  How  much  energy?  The  kinetic 
energy  the  rubber  band  has  depei.  jts  upon  how 
much  work  wa'  done  in  stretching  it.  Wx  fixes 
(KE)z.  We  may  write 


Wx  =  (KE)» 


U3) 


Again  we  recognize  work  as  a  form  of  energy 
and,  since  Wx  =  (KE)3,  the  overall  result  is  that 
energy  was  conserved.  But  was  energy  conserved 
in  Figure  7-4B?  If  so,  where  is  the  energy?  The 
work  Wx  has  already  been  done.  Though  the 
rubber  band  is  stretched,  there  is  no  outward 
manifestation  of  energy.  The  rubber  band  is  mo- 
tionless, so  it  has  received  no  kinetic  energy. 
Yet,  we  know  (from  previous  experiments)  that 
the  rubber  band  will  receive  the  energy  (KE)3  at 
any  instant  that  we  choose  to  release  it.  The 
initial  and  final  result  are  reminiscent  of  the 
billiard  balls:  energy  is  conserved.  It  would  be 
convenient  to  say  that  energy  is  conserved  in 
Figure  7-4B,  as  well.  So,  we  invent  a  new  form 
of  energy — energy  of  position,  or  potential  en- 
ergy. We  say  that  as  the  work  Wx  was  performed 
it  was  stored  in  the  rubber  band  as  potential 
energy. 

Now  we  can  review  this  experiment.  In  sym- 
bols we  have 


Wx  =  (PE)2  =  (,KE)t 

work        potential       kinetic 
energy  energy 


U4) 


The  amount  of  work  performed  fixed  Wx.  Meas- 
urements of  mass  and  velocity  of  the  rubber  band 
tell  us,  experimentally,  the  magnitude  of  (KE)Z. 
How  do  we  know  (PE^l  How  are  we  sure  that 
(P£>2  is  equal  to  Wx  and  to  (KE^l  The  evidence 
we  have  is  that  we  put  an  amount  of  energy  into 
the  system  and  can  recover  all  of  it  later  at  will. 
It  is  natural  to  say  the  energy  is  stored  in  the 
meantime.  Then  we  can  say  that  the  rubber  band 
is  just  like  the  billiard  ball  collision:  energy  is 
conserved  at  all  times. 


Electrodes 


Source 

electrical 

energy, 

direct 

current- 


Fig.  7-5.  Electrolytic  decomposition  of  water. 


7-2.3    Conservation  of  Energy  in  a 
Chemical  Reaction 

Figure  7-5  shows  an  apparatus  in  which  an  elec- 
tric current  can  be  passed  through  water.  As 
remarked  in  Section  3-1.2,  the  electric  current 
causes  a  decomposition  of  water.  As  work  is 
done  (electrical  work),  hydrogen  gas  and  oxygen 
gas  are  produced.  Measurements  of  the  electric 
current  and  voltage  show  that  68.3  kcal  of  elec- 
trical work,  Wx,  must  be  done  to  decompose  one 
mole  of  water.  The  equation  for  the  reaction  is 

68.3  kcal  +  HMD  —+  Wg)  +  Wg)    (75) 

electrical 
work 

Now  suppose  we  measure  the  heat  of  reaction 
of  hydrogen  and  oxygen  in  a  calorimeter  like 
that  shown  in  Figure  7-2.  This  experiment  has 
been  performed  many  times;  68.3  kcal  of  heat, 
Qz,  are  produced  for  every  mole  of  water  formed. 
The  equation  for  this  reaction  is 

Wg)  +  \Oi(g)  — >-  H20(l)  +  68.3  kcal  heat    (76) 

We  have  a  situation  just  like  that  of  the  rubber 
band.  We  put  a  readily  measured  amount  of 
energy,  Wx,  into  the  system  and,  at  any  time 


116 


ENERGY    EFFECTS    IN    CHEMICAL    REACTIONS   I   CHAP.    7 


later  that  we  choose,  this  energy  can  be  recovered 
as  heat,  Q%: 

W,  =  Q,  (17) 

The  overall  result  is  that  energy  is  conserved. 

Figure  7-7  shows  this  schematically  in  a  dia- 
gram like  that  of  Figure  7-1.  If  the  heat  content 
of  two  moles  of  water  is  represented  by  a  line 
on  this  diagram,  then  the  energy  of  two  moles 
of  hydrogen  plus  one  mole  of  oxygen  should  be 
represented  by  a  line  136.6  kcal  higher.  Now  the 
diagram  indicates  that  when  water  is  decom- 
posed, energy  must  be  put  in  to  raise  the  heat 
content  enough  to  form  the  products.  When  hy- 


Fig.  7-6.  Conservation  of  energy  in  a  chemical 
reaction. 


drogen  burns  in  oxygen,  the  heat  content  drops 
and  energy  is  released. 

Again  we  may  ask :  Where  was  the  energy  put 
in  reaction  (15)  before  we  carry  out  reaction 
(16)1  The  rubber  band  experiment  guides  us. 

It  is  easier  to  explain  why  Wx  =  Q3  if  we  say 
that  the  energy  W\  was  stored  in  the  chemical 
substances  U2(g)  and  O-rfg).  We  assign  to  these 
(and  all  other)  substances  the  capacity  to  store 
energy  and  we  call  it  heat  content.  This  permits 
us  to  say  that  energy  is  conserved  at  all  times 
during  a  chemical  reaction  as  it  is  in  billiard  ball 
collisions  and  in  stretched  rubber  bands. 

7-2.4    The  Basis  for  the  Law 

off  Conservation  of  Energy 

We  see  that  there  are  many  forms  of  energy.  We 
have  talked  about  work  as  a  form  of  energy  and 


Two    moles 
liqti  id    wcvte r> 

2H20  (I) 


-h      2(68.3  kcal) 


,  136.  6  kcal 

electrical    work 


+■ 


wt 


+ 


tifcl 


two   moles 
hydroge-n   gas 


+ 


2Kz(g) 


one    mole 
oxygen,    qcls 

+  02   (f) 


+ 


+  Z  [68.3  kcal) 


Two   mx>les 

+■ 

one  m-ole 
oxygen.  ga\s 

two   moles 

+ 

136.6  kcal 

hydrogen   gas 

liquid   wafer 

heat 

2H2  (9) 

+■ 

n    /    l 

+ 

&3 

°z  (<fJ 

*■               2J120  (  IJ 

Wt 

=    0s 

SEC.    7-3    |    THE    ENERGY    STORED    IN    A    MOLECULE 


117 


Heat 
content 
(potential 
energy) 


2H2(9)+  Q2(g) 


AH  =136.6   kcal 


Decomposition    of  water- 
Formation     of   water- 


Fig.  7-7 .  The  energy  change  in  a  chemical  reaction. 

referred  to  two  kinds  of  work,  muscular  work 
and  electrical  work.  Kinetic  energy  and  heat  are 
other  readily  measured  forms  of  energy.  We  have 
added  two  additional  forms  that  are  only  in- 
directly detectable — the  potential  energy  of  the 
stretched  rubber  band  and  the  heat  content  of 
chemical  substances.  With  these  two  added 
forms,  we  can  write: 

Energy  is  always  conserved. 

This  is  called  the  Law  of  Conservation  of 
Energy.  It  says  that  in  every  experiment  so  far 
performed,  energy  was  conserved  provided  all  of 
the  different  forms  of  energy  are  taken  into  ac- 
count. Because  the  number  of  such  experiments 
is  extremely  large  and  varied  in  type,  the  law 
gives  a  reliable  basis  for  predicting.  Don't  be 
upset  that  the  law  is  true  only  because  we  added 
forms  of  energy  to  account  for  energy  not  di- 
rectly measurable.  This  law  is  like  any  other  law 
— its  usefulness  justifies  it.  As  long  as  the  several 
forms  of  energy  give  us  a  model  that  is  always 
consistent  with  experiments,  the  law  remains 
useful. 


In  specific  reference  to  the  heat  effects  in  chem- 
ical reactions,  hundreds  of  different  reactions 
have  been  studied  calorimetrically.  The  results 
are  always  in  accord  with  the  Law  of  Additivity 
of  Reaction  Heats.  If  we  assign  a  characteristic 
heat  content  to  each  chemical  substance,  then 
all  of  these  experiments  support  the  Law  of  Con- 
servation of  Energy.  Since  the  Law  of  Conserva- 
tion of  Energy  is  consistent  with  so  many  dif- 
ferent reactions,  it  can  be  safely  assumed  to  apply 
to  a  reaction  which  hasn't  been  studied  before. 

The  term  "law"  is  usually  applied  to  the  older  scientific 
generalizations.  Modern  scientists  do  not  apply  the  term 
to  new  generalizations  because  they  realize  that  all  "laws 
of  nature"  are  human  statements — human  generalizations 
— and  are  subject  to  revision.  For  example,  later  in  this 
chapter  you  will  find  that  matter  and  energy  are  one  and 
the  same.  At  that  time,  you  will  see  that  the  two  conserva- 
tion laws,  the  Conservation  of  Matter  and  the  Conser- 
vation of  Energy,  are  really  but  one  law,  the  Conservation 
of  Matter  (which  is  Energy).  Yet  in  any  chemical  process 
the  mass  equivalent  of  the  reaction  heat  is  negligible. 
Under  these  conditions,  the  Law  of  Conservation  of 
Matter  and  the  Law  of  Conservation  of  Energy  can  be 
considered  as  independent  statements.  In  this  form  the 
conservation  laws  are  very  useful,  even  though  the  state- 
ments we  have  made  about  them  do  not  apply  under  all 
conditions. 


7-3    THE  ENERGY  STORED  IN  A  MOLECULE 


In  the  discussion  of  Sections  7-2.1  to  7-2.4,  we 
found  it  useful  to  talk  about  different  "forms" 
of  energy.  Two  of  these  are  heat  and  heat  con- 


tent. Heat  content  is  sometimes  called  "chemical 
energy"  because  its  magnitude  is  intimately  tied 
up  with  chemical  composition.  These  are  macro- 


118 


ENERGY    EFFECTS    IN    CHEMICAL    REACTIONS   I    CHAP.    7 


scopic*  manifestations  of  energy.  Two  other 
macroscopic  manifestations  of  energy  are  kinetic 
energy  (of  a  thrown  baseball,  for  example)  and 
potential  energy  (of  a  baseball  at  the  top  of  the 
flight  of  a  high  foul  ball,  for  example).  Thus,  we 
need  to  identify  several  "forms"  of  energy  when 
discussing  the  macroscopic  properties  of  sub- 
stances: heat,  heat  content  (chemical  energy), 
kinetic  energy,  potential  energy,  electrical  en- 
ergy, and  mechanical  work.  The  presence  and 
amount  of  each  energy  form  is  determined  by 
methods  uniquely  applicable  to  that  form.  We 
determine  the  quantity  of  heat  released  in  a 
calorimeter  by  measuring  a  temperature  rise  with 
a  thermometer.  We  measure  the  kinetic  energy 
of  a  baseball  with  a  watch  and  a  meter  stick. 
You  would  learn  little  about  the  kinetic  energy 
of  a  baseball  by  throwing  it  at  a  thermometer 
and  nothing  about  water  temperature  by  wearing 
your  wrist  watch  in  the  shower. 

When  we  turn  to  the  molecular  scale,  however, 
we  discover  that  all  of  these  macroscopic  forms 
of  energy  can  be  discussed  in  terms  of  the  two 
kinds  of  energy  we  assigned  to  the  baseball, 
kinetic  and  potential  energies.  We  can  "explain" 
all  macroscopic  forms  of  energy  with  a  micro- 
scopic model  involving  only  the  energy  of  motion 
and  energy  of  position  of  the  atomic  and  molecu- 
lar particles.  The  explanation  has  the  special 
advantage   given   in   Section    1-1.3    (pp.    5-8). 

*  Macroscopic  means  on  a  large  scale — the  opposite 
of  microscopic.  In  general,  it  is  used  to  indicate  weighable 
and  visible  amounts. 


f* 

./ 

/'■■'.'         \ 

1\ 

'  J> 

^•0--£J 

?>-i 

i~iu&  v^ 

. 

»w 


))»■ 


n 


All  of  our  experience  and  knowledge  about  the 
properties  of  moving  baseballs  (and  billiard  balls 
and  rubber  bands  and  automobiles  and  pendu- 
lums and  gyroscopes)  can  be  used  in  clarifying 
the  nature  of  heat,  heat  content,  electrical  energy, 
etc.  To  see  this,  we  must  consider  how  chemists 
discuss  the  energy  held  by  a  molecule. 

7-3.1    The  Energy  off  a  Molecule 

Let  us  picture  a  molecule  in  terms  of  a  model 
consisting  of  balls  of  proper  relative  masses 
hooked  together  by  springs.  The  springs  repre- 
sent the  bonds  between  the  atoms.  We  can  start 
the  springs  vibrating  and  then  toss  the  entire 
assembly  through  space  in  an  end-over-end  mo- 
tion. There  are  now  three  kinds  of  kinetic  energy 
associated  with  our  model,  as  pictured  in  Fig- 
ure 7-8. 

This  model  applies  quite  well  to  a  molecule 
in  the  gaseous  state,  but  in  the  liquid  state,  and 
(even  more  so)  in  the  solid  state,  all  these  mo- 
tions are  restricted.  In  these  phases  the  chief 
kinetic  energy  manifestation  is  a  back-and-forth 
motion  of  the  molecule  about  a  fixed  point. 


Fig.  7-8.  Types  of  motion  of  a  molecule  of  carbon 
dioxide,  CO2.  A.  Translational  motion;  the 
molecule  moves  from  place  to  place.  B.  Ro- 
tational motion;  the  molecule  rotates  about 
its  center  of  mass.  C.  Vibrational  motion; 
the  atoms  move  alternately  toward  and  away 
from  the  center  of  mass. 


\       I 

w 


))))  >  . 


5  \m> 


mv 


SEC.    7-3    I    THE    ENERGY    STORED    IN    A    MOLECULE 


119 


In  addition  to  the  various  kinds  of  kinetic 
energy  listed  in  Figure  7-8,  there  is  potential  en- 
ergy related  to  the  forces  which  act  between 
molecules.  These  forces  are  attractive,  having  a 
very  small  average  value  in  the  gaseous  state  in 
which  the  molecules  are  far  apart.  The  forces  are, 
on  the  average,  larger  in  the  liquid  state  and  are 
still  larger  in  the  solid  state. 

Next,  there  is  present,  within  the  molecule, 
chemical  energy  which  is  related  to  the  forces 
which  hold  the  atoms  together  in  the  molecule. 
This  is  referred  to  as  chemical  bond  energy. 

In  addition,  each  atom  has  energy,  some  as- 
sociated with  the  electrons  and  some  with  the 
nucleus.  The  electrons  in  the  atom  possess  kinetic 
energy  and,  because  of  their  attraction  to  the 
nucleus  and  repulsion  from  each  other,  they  also 
possess  potential  energy.  The  algebraic  sum  of 
these  kinetic  and  potential  energies  represents 
the  energy  necessary  to  pull  an  electron  away 
from  an  atom. 

Finally,  there  is  present  within  the  nucleus  of 
each  atom  a  store  of  energy.  This  energy  is  re- 
lated to  the  forces  holding  the  nuclear  particles 
together.  Since  each  nucleus  remains  intact  and 
apparently  uninfluenced  through  chemical  reac- 
tions, this  nuclear  energy  does  not  change. 
Hence,  the  nuclear  contribution  to  the  molecular 
heat  content  does  not  usually  concern  a  chemist. 

The  sum  of  all  of  these  forms  of  molecular 
energy  makes  up  the  molecular  heat  content.  If 
we  add  together  the  molecular  heat  content  of 
6.02  X  1023  molecules  of  a  given  kind,  we  obtain 
the  molar  heat  content  of  that  substance. 

7-3.2    Energy  Changes  on  Warming 

Having  this  view  of  the  make-up  of  the  heat  con- 
tent of  a  substance,  we  can  now  visualize  the 
effects  brought  on  by  warming  the  substance.  If 
the  temperature  is  low  at  first,  the  substance  will 
be  a  solid.  Warming  the  solid  increases  the  ki- 
netic energy  of  the  back-and-forth  motions  of 
the  molecules  about  their  regular  crystal  posi- 
tions. As  the  temperature  rises,  these  motions 
disturb  the  regularity  of  the  crystal  more  and 
more.  Too  much  of  this  random  movement  de- 
stroys the  lattice  completely.  At  the  temperature 


above  which  the  kinetic  energy  of  the  particles 
causes  so  much  random  movement  that  the  lat- 
tice is  no  longer  stable,  a  phase  change  occurs: 
the  solid  melts. 

In  the  liquid  each  molecule  has  considerably 
more  freedom  of  movement,  particularly  for 
translation  and  rotation.  Warming  the  liquid 
enhances  the  amount  of  molecular  movement. 
As  always,  kinetic  energy  provides  a  randomiz- 
ing effect,  tending  to  carry  the  molecules  every- 
where in  the  container.  As  the  energy  of  motion 
rises  (with  the  rising  temperature),  more  of  the 
molecules  are  able  to  move  away  from  the  liquid 
region  where  the  potential  energy  is  a  minimum. 
Another  phase  change  occurs:  the  liquid  va- 
porizes. 

If,  now,  we  continue  warming  the  substance 
sufficiently,  we  will  reach  a  point  at  which  the 
kinetic  energies  in  vibration,  rotation,  and  trans- 
lation become  comparable  to  chemical  bond  en- 
ergies. Then  molecules  begin  to  disintegrate. 
This  is  the  reason  that  only  the  very  simplest 
molecules — diatomic  molecules — are  found  in 
the  Sun.  There  the  temperature  is  so  high 
(6000°K  at  the  surface)  that  more  complex  mole- 
cules cannot  survive. 

Finally,  if  we  continue  the  heating  still  further, 
we  will  ultimately  reach  a  temperature  at  which 
the  kinetic  energies  are  large  enough  to  disrupt 
the  nuclei.  Then,  "nuclear  reactions"  begin.  The 
conditions  in  some  stars  are  considered  to  be 
suitable  for  rapid  nuclear  reactions. 

To  conclude  this  study,  let's  consider  the  mag- 
nitudes of  the  energy  effects.  Phase  changes  usu- 
ally involve  energies  of  several  kilocalories  per 
mole.  Chemical  reactions  usually  involve  ener- 
gies of  50  to  several  hundred  kilocalories  per 
mole.  Thus,  we  see  that  the  energies  involved  in 
chemical  reactions  are  usually  10  to  100  times 
larger  than  those  involved  in  phase  chages. 


EXERCISE  7-4 

Show  that  the  ratio  of  the  molar  heat  of  for- 
mation of  gaseous  water  from  the  elements  (a 
chemical  reaction)  to  the  molar  heat  of  the  fusion 
of  water  (a  phase  change)  is  of  the  order  of  50. 


120 


ENERGY    EFFECTS    IN    CHEMICAL    REACTIONS    I    CHAP.    7 


7-4    THE  ENERGY  STORED  IN  A  NUCLEUS 


Now  let  us  consider  nuclear  changes.  The  fact 
that  nuclei  do  remain  intact  during  chemical  re- 
actions suggests  that  much  larger  energies  are 
required  for  nuclear  changes.  Experimentally, 
this  proves  to  be  true.  Nuclear  reactions  usually 
involve  energy  changes  over  a  million  times 
greater  than  those  we  find  in  chemical  reactions. 
This  enormous  factor  accounts  for  the  current 
interest  in  nuclear  reactions  as  a  source  of  en- 
ergy. 

One  such  nuclear  reaction  is  represented  by 
the  equation 

2giU  +  In  — >-  'tlBa  +  llKr  +  Ion  +  energy    (18) 

Before  we  examine  the  details  of  this  rather 
strange  looking  equation,  let  us  focus  our  atten- 
tion on  the  "+  energy"  term.  The  numerical 
value  is  of  the  order  of  4.5  X  109  kcal/mole  of 
uranium.  Look  at  that  figure  again  and  compare 
it  to  the  molar  heat  of  combustion  of  carbon. 
Roughly,  what  is  the  ratio  of  these  two  energies? 
It  is  107,  or  10  million! 

Now  let  us  examine  the  reaction  in  more  de- 
tail. Forget  momentarily  the  subscripts  and  su- 
perscripts. Recall  from  Chapter  6  that  the 
neutron  («)  is  one  of  the  fundamental  particles 
visualized  as  present  in  nuclei.  What  has  hap- 
pened? 

U  +  n  — >-  Ba  +  Kr  +  3/7  (19) 

Instead  of  producing  new  kinds  of  substances  by 
combination  of  atoms,  the  element  uranium  has 
combined  with  a  neutron  and  as  a  result  has 
split  into  two  other  elements — barium  and  kryp- 
ton— plus  three  more  neutrons.  Atoms  of  a  given 
element  are  characterized  by  their  atomic  num- 
ber, the  number  of  units  of  positive  charge  on 
the  nucleus.  For  one  element  to  change  into 
another  element  the  nucleus  must  be  altered.  In 
our  example  the  uranium  nucleus,  as  a  result  of 
reacting  with  a  neutron,  splits  or  fissions  into  two 
other  nuclei  and  releases,  in  addition,  neutrons.* 


How  we  get  neutrons  and  how  we  get  them  to 
react  with  uranium  nuclei  is  not  essential  to  our 
present  discussion. 

A  glance  at  the  periodic  table  will  show  that 
the  subscripts  we  have  attached  to  our  symbols 
are  the  atomic  numbers  of  the  elements  desig- 
nated by  the  symbols — 92  for  U,  56  for  Ba,  36 
for  Kr.  The  zero  subscript  attached  to  the  neu- 
tron denotes  the  lack  of  charge  on  this  particle. 
If  we  look  at  the  subscripts, 

92U  +  0n  — >■  MBa  +  seKr  +  3o/i  (20) 

we  notice  that  their  sum  on  each  side  of  the 
equation  is  identical: 

92  +  0  =  56  +  36  +  (3  X  0) 

This  identity  is  but  another  way  of  expressing 
the  law  of  conservation  of  charge. 

In  the  model  of  nuclear  structure  you  were 
given  in  Chapter  6,  the  nucleus  was  pictured  as 
being  built  up  of  protons  and  neutrons.  These 
two  kinds  of  particles  are  given  the  general  name 
nucleon.  The  mass  number  of  a  nucleus  is 
equal  to  the  number  of  nucleons  present.  The 
superscripts  in  our  equation  are  mass  numbers: 


235IJ    +    1„ 


'Ba  +  wKr  +  3'/i 


(21) 


Apparently  the  mass  numbers  are  also  con- 
served : 

235  +  1  =  141  +  92  +  (3  X  1) 

We  may  rephrase  this  in  the  form  of  a  rule:  The 
total  number  of  nucleons  is  unchanged  during  nu- 
clear reactions. 

EXERCISE  7-5 

According  to  the  model  of  Chapter  6,  how  many 
nucleons  would  be  present  in  a  uranium  nucleus 
of  mass  number  235?  How  many  protons  are 
pictured  as  being  present?  How  many  neutrons? 


*  Perhaps  you  have  already  recognized  our  nuclear 
reaction  as  a  fission  reaction.  It  is  of  the  type  of  reaction 
used  in  an  atomic  pile,  the  energy  source  of  a  nuclear 


power  plant.  The  example  we  have  selected  is  only  one 
of  the  ways  the  uranium  nucleus  can  divide.  Lanthanum 
and  bromine  nuclei  are  also  produced,  cerium  and  sele- 
nium, and  so  on,  each  pair  of  fission  products  being  such 
that  the  sum  of  their  atomic  numbers  is  always  92. 


SEC.    7-4    I    THE    ENERGY    STORED    IN    A    NUCLEUS 


121 


Actually,  then,  by  our  symbol  "iJU  we  are 
representing  not  an  atom,  but  a  nucleus.  Our 
equation  is  written  in  terms  of  nuclei  and  par- 
ticles associated  with  them.  This  nuclear  equa- 
tion tells  us  nothing  about  what  compound  ot 
uranium  was  bombarded  with  neutrons  or  what 
compound  of  barium  is  formed.  We  are  sum- 
marizing only  the  nuclear  changes.  During  the 
nuclear  change  there  is  much  disruption  of  other 
atoms  because  of  the  tremendous  amounts  of 
energy  liberated.  We  do  not  know  in  detail  what 
happens  but  eventually  we  return  to  electrically 
neutral  substances  (chemical  compounds)  and 
the  neutrons  are  consumed  by  other  nuclei. 

It  is  easy  to  determine  that  there  is  an  electron 
balance  between  the  reactants  and  products: 

Nucleus:  »2U      on      wBa      36Kr      on 


Electrons  associated 
with  nucleus: 


92        0 


56 


36 


0 


92  =  56  +  36 


7-4.1    Exact  Mass  Relationships 

Although  the  mass  numbers  of  the  proton  and  neutron 
are  both  one,  the  masses  of  these  fundamental  particles 
are  not  identical.  The  mass  of  one  mole  of  protons  is 
1.00762  grams  and  that  of  one  mole  of  neutrons  is 
1.00893  grams.  Further  investigation  would  show  that 
the  experimentally  measured  mass  of  the  nucleus  of  any 
given  isotope  is  not  the  exact  sum  of  the  masses  of 
protons  and  neutrons  confined  in  the  nucleus  according 
to  our  model.  For  example,  the  mass  of  the  nucleus  of 
the  uranium  isotope  of  mass  number  235  is  less  than  the 
exact  sum  of  the  masses  of  92  protons  and  143  neutrons. 

One  of  the  consequences  of  the  special  theory  of  rela- 
tivity formulated  in  1905  by  the  great  German  theoretical 
physicist,  Albert  Einstein,  was  that  we  came  to  realize 
that  mass  and  energy  are  one  and  the  same.  Although 
this  was  a  very  radical  notion  at  the  time  Einstein  first 
presented  his  theory,  the  equation  relating  mass  to  energy 
is  probably  already  familiar  to  you.  The  formula  E  =  mc* 
has  become  almost  a  part  of  common  idiom  since  the 
successful  application  of  nuclear  energy  became  a  part 
of  modern  technology  in  the  mid  1940's.  In  this  equation 
c  is  the  speed  of  light,  3.00  X  10'"  cm/second.  Apparently 
a  small  value  of  mass  (m)  is  equivalent  to  a  tremendous 
amount  of  energy  since  the  proportionality  constant  (c2) 
relating  mass  to  energy  is  numerically  9.00  X  10'°. 

We  can  use  this  idea  of  the  relation  of  mass  to  energy 
in  several  ways.  The  mass  of  a  "  U  nucleus  is  less  than 
the  sum  of  the  masses  of  the  92  protons  and  143  neutrons 
postulated  to  be  in  it.  The  difference  in  mass  represents 
the  binding  energy  which  holds  the  nucleons  together  in 


the  nucleus.  Here  we  have  used  the  concept  of  expressing 
the  nuclear  binding  energy  in  terms  of  the  implied  de- 
crease in  mass.  We  can  do  the  same,  if  we  wish,  for  a 
chemical  reaction.  Again  let  us  return  to  the  molar  heat 
of  combustion  of  carbon,  roughly  102  kcal: 


Qs)  +  Ot(g) 


CO>(g)    AH  =  -94  kcal        (22) 


The  mass  change  associated  with  an  energy  change  of 
102  kcal*  is  of  the  order  of  5  X  10~9  grams.  This  is  a 
quantity  far  too  small  to  be  detected  on  any  balance 
capable  of  weighing  the  12  grams  of  carbon  and  32  grams 
of  oxygen  consumed  in  the  reaction.  Since  the  chemical 
"mass  defects"  are  too  small  to  measure,  we  do  not  use 
this  terminology  in  chemistry. 

If  we  wish  to  gain  some  idea  of  the  alteration  of  mass 
in  a  nuclear  change,  we  cannot  use  the  fission  reaction 
because  the  exact  masses  of  the  nuclei  involved  are  not 
known.  Let  us  look  at  another  type  of  reaction  of  pos- 
sible importance  in  the  production  of  nuclear  energy: 


iH  + 


H 


•'He  +  lQn 


(23) 


This  reaction  is  called  fusion  since  nuclei  are  combining 
to  form  a  heavier  nucleus.  The  energy  associated  with 
this  change  is  4.05  X  108  kcal/mole  of  ?H  nuclei. 

Let  us  do  a  little  bookkeeping  with  the  exact  masses 
of  these  nuclei.  Actually  we  will  simplify  a  bit  and  use 
the  exact  masses  of  the  atoms.  This  will  make  no  dif- 
ference. The  masses  of  the  atoms  differ  from  the  nuclear 
masses  by  the  masses  of  the  number  of  electrons  in  each 
atom.  We  have  shown  that  electrons  are  conserved  in 
nuclear  changes.  Exact  masses  of  atoms  (that  is,  exact 
masses  of  each  isotopic  species  and  not  the  chemical 
atomic  weights  shown  on  the  inside  back  cover)  are 
readily  available.  For  our  hydrogen-helium  reaction  we 
have 

Reactants:       2H        2.01471  g/mole 


Products: 


3H 
4He 


3.01707 
5.03178 
4.00390 
1.00893 
5.01283 


Reactants: 
Products: 
Mass  Difference: 


5.03178 
5.01283 
0.01895  g/mole 


Compare  this  mass  difference  of  about  0.02  g/mole  with 
one  of  about  5  X  10-9  g/mole  for  the  combustion  of 
carbon. 

In  closing,  let  us  remind  ourselves  of  the  difference 
between  nuclear  and  chemical  reactions.  In  nuclear  re- 
actions, changes  in  the  nuclei  take  place.  In  chemical 
reactions,  the  nuclei  remain  intact  and  the  changes  are 
explainable  in  terms  of  the  electrons  outside  the  nucleus. 


*  Before  using  the  E  =  mc-  relation  to  calculate  the 
amount  of  mass  associated  with  this  energy  change,  you 
must  pay  attention  to  the  relation  of  various  energy  units. 


122 


ENERGY    EFFECTS    IN    CHEMICAL    REACTIONS    I    CHAP.    7 


QUESTIONS  AND  PROBLEMS 

1.  Given 

3C(s)  +  2Fe203fsj  +110.8  kcal  — *- 

4Fe(s)  +  ZCO,(g) 

Rewrite  the  equation  using  one  mole  of  carbon 
and  use  the  AH  notation. 


Given 

\Wg)  +  \*r2(l) 


— >■  HBr(g) 

AH  =  -8.60  kcal/mole  HBr 


Rewrite  the  equation  for  one  mole  of  hydrogen 
gas  and  include  the  heat  effect  as  a  term  in  the 
equation. 

Which  of  the  following  reactions  are  endother- 
mic? 


(a)  H2(g)  +  \02(g)  - 

(b)  h^g)  +  \Oi(g) 


U*0(g) 

AH  =  -57.8  kcal 


NOfgJ 

AH  =  +21.6  kcal 

(c)  \K2(g)  +  02(g)  +  8.1  kcal  — >-  N02(g) 

(d)  hK2(g)  +  W*(g)  — >■  NH3(gj  +11.0  kcal 

(e)  NH3(gj  — ►-  ±N2(g)  +  |H2(gj 

AH  =  +11.0  kcal 

4.  What  is  the  minimum  energy  required  to  syn- 
thesize one  mole  of  nitric  oxide,  NO,  from  the 
elements  ? 

5.  How  much  energy  is  liberated  when  0.100  mole 
of  H2  (at  25°C  and  1  atmosphere)  is  combined 
with  enough  02(g)  to  make  liquid  water  at  25°C 
and  1  atmosphere? 

6.  How  much  energy  is  consumed  in  the  decom- 
position of  5.0  grams  of  H20(l)  at  25°C  and 
1  atmosphere  into  its  gaseous  elements  at  25°C 
and  1  atmosphere? 

7.  Using  Table  7-II,  calculate  the  heat  of  burning 
ethane  in  oxygen  to  give  C02  and  water  vapor. 

Answer.  AH  =  —341  kcal/mole  C2H6. 

8.  Given 


C(diamond)  +  02(g) 
C(graphite)  +  02(g) 


C02(g) 
AH  =  -94.50  kcal 

C02(g) 
AH  =  -94.05  kcal 


Find  AH  for  the  manufacture  of  diamond  from 
graphite. 


C(graphite) 


C(diamond) 


Is  heat  absorbed  or  evolved  as  graphite  is  con- 
verted to  diamond? 

9.  To  change  the  temperature  of  a  particular  calo- 
rimeter and  the  water  it  contains  by  one  degree 
requires  1 550  calories.  The  complete  combustion 
of  1.40  grams  of  ethylene  gas,  C2H4(g),  in  the 
calorimeter  causes  a  temperature  rise  of  10.7 
degrees.  Find  the  heat  of  combustion  per  mole 
of  ethylene. 

10.  The  "thermite  reaction"  is  spectacular  and 
highly  exothermic.  It  involves  the  reaction  be- 
tween Fe203,  ferric  oxide,  and  metallic  alumi- 
num. The  reaction  produces  white-hot,  molten 
iron  in  a  few  seconds.  Given : 

2A1  +  f  02  — >-  A1203    AHx  =  -400  kcal/mole 

2Fe  +  f02  — ►-  Fe2Os    AH2  =  -200  kcal/mole 

Determine  the  amount  of  heat  liberated  in  the 
reaction  of  1  mole  of  Fe203  with  Al. 

Answer.  AH  =  —200  kcal/mole  Fe203. 

11.  How  much  energy  is  released  in  the  manufacture 
of  1.00  kg  of  iron  by  the  "thermite  reaction" 
mentioned  in  Problem  10? 

12.  How  many  grams  of  water  could  be  heated  from 
0°C  to  100°C  by  the  heat  liberated  per  mole  of 
aluminum  oxide  formed  by  the  "thermite  reac- 
tion," as  described  in  Problem  10? 

13.  Which  would  be  the  better  fuel  on  the  basis  of 
the  heat  released  per  mole  burned,  nitric  oxide, 
NO,  or  ammonia,  NH3?  Assume  the  products 
are  N02(gj  and  H20(g). 

14.  What  is  the  minimum  energy  required  to  syn- 
thesize sulfur  dioxide  from  sulfuric  acid? 

H2S04f/J  — >  SO/gJ  +  H20(g)  +  \02(g) 

Answer.  AH  =  +65  kcal/mole  S02(g). 

15.  Why  is  the  Law  of  Conservation  of  Energy  con- 
sidered to  be  valid? 

16.  What  do  you  think  would  happen  in  scientific 
circles  if  a  clearcut,  well-verified  exception  was 


QUESTIONS    AND    PROBLEMS 


123 


found  to  the  Law  of  Conservation  of  Energy  as 
stated  in  the  text? 

17.  Is  energy  conserved  when  a  ball  of  mud  is 
dropped  from  your  hand  to  the  ground?  Explain 
your  answer. 

18.  What  becomes  of  the  energy  supplied  to  water 
molecules  as  they  are  heated  in  a  closed  con- 
tainer from  25°C  to  35°C? 

19.  Outline  the  events  and  associated  energy  changes 
that  occur  on  the  molecular  level  when  steam  at 
150°C  and  1  atmosphere  pressure  loses  energy 
continually  until  it  finally  becomes  ice  at  -  10°C. 

20.  The  heat  of  combustion  of  methane,  CH4,  is 
—210  kcal/mole: 


CHA(g)  +  202 


C02  +  2H20 

AH  =  -210  kcal 


Discuss  why  this  fuel  is  better  than  water  gas  if 
the  comparison  is  based  on  one  mole  of  carbon 
atoms. 

21.  In  a  nuclear  reaction  of  the  type  called  "nuclear 
fusion,"  two  nuclei  come  together  to  form  a 
larger  nucleus.  For  example,  deuterium  nuclei, 


23. 


iH,  and  tritium  nuclei,  jH,  can  "fuse"  to  form 
helium  nuclei,  2He,  and  a  neutron: 

aHe  +  In 

AH  =  -4.05  X  108  kcal 


?H  +  ?H 


How  many  grams  of  hydrogen  would  have  to  be 
burned  (to  gaseous  water)  to  liberate  the  same 
amount  of  heat  as  liberated  by  fusion  of  one 
mole  of  iH  nuclei?  Express  the  answer  in  tons 
(1  ton  =  9.07  X  10s  g). 

22.  Which  of  the  following  reactions  is  most  likely 
to  have  a  heat  effect  of  -  505  kcal?  Which  would 
be  -1.7  X  10«  kcal?  Which  would  be  +7.2 
kcal? 


(a)  UF,fi;  — >■  UFt(g) 

(b)  U(s)  +  3F2(g)  -^  UF6 

(c)  2mU  +  In 


239,  T 
92  *-> 


AH  =  ? 
AH  =  ? 
AH  =  ? 


Fission  of  uranium  gives  a  variety  of  fission 
products,  including  praseodymium,  Pr.  If  the 
process  by  which  praseodymium  is  formed  gives 
^gPr  and  three  neutrons,  what  is  the  other  nu- 
clear product? 


235i  t  _i_  x„ 


^Pr  +  ?  +  3j/i 


CHAPTER 


8 


The  Rates  of 
Chemical  Reactions 


•  •  •  a  molecular  system  •  •  •  [passes]  •  •  •  from  one  state  of 
equilibrium  to  another  •  •  -  by  means  of  all  possible  intermediate 
paths,  but  the  path  most  economical  of  energy  will  be  more  often  travelled. 

HENRY    EYRING,    1945 


A  candle  remains  in  contact  with  air  indefinitely 
without  observable  reaction  but  it  reacts  when 
given  a  start  with  a  lighted  match.  A  mixture  of 
household  gas  and  air  in  a  closed  room  remains 
indefinitely  without  reacting  but  it  may  explode 
violently  if  so  much  as  a  glowing  cigarette  is 
brought  into  the  room.  A  piece  of  iron  reacts 
quite  slowly  with  air  (it  rusts)  but  a  piece  of  white 
phosphorus  bursts  into  flame  when  it  is  exposed 
to  air.  These  are  all  reactions  with  oxygen  from 
the  air  but  they  have  extremely  different  time 
behaviors.  Reactions  proceed  at  different  rates. 
We  care  about  reaction  rate  because  we  must 
understand  how  rapidly  a  reaction  proceeds  and 
what  factors  determine  its  rate  in  order  to  bring 
the  reaction  under  control.* 

Let  us  see  what  the  expression  "the  rate  of  a 
reaction"  means  in  terms  of  an  example — the 
reaction  between  carbon  monoxide  gas,  CO,  and 
nitrogen  dioxide,  N02.  Chemical  tests  show  that 


*  The   study   of  reaction    rates   is  called   chemical 
kinetics. 
124 


the  products  are  carbon  dioxide,  C02,  and  nitric 
oxide,  NO.  The  equation  for  the  reaction  is 


CO  +  N02 


CO.  +  NO 


(') 


Suppose  we  prepare  a  mixture  of  carbon  mon- 
oxide and  nitrogen  dioxide  and  then  heat  it  to 
200°C.  When  the  gas  is  heated,  we  observe  a 
gradual  disappearance  of  the  reddish  brown 
color  of  N02.  Reaction  is  taking  place.  We  can 
find  its  time  behavior  by  measuring  the  change 
in  color  during  a  measured  time  interval.  Since 
the  other  gases  are  colorless,  this  color  change 
indicates  the  number  of  moles  of  N02  that  have 
reacted  during  the  time  interval.  The  quotient  of 
these  two,  moles  reacted  divided  by  the  time 
interval,  is  called  the  rate  of  the  reaction: 


Rate  = 


quantity  NQ2  consumed 


time  interval 
=  quantity  N02  consumed  per  unit  time 

We  can  express  the  rate  of  reaction  (7)  in  terms 
of  the  rate  of  consumption  of  either  CO  or  N02. 
Equally  well,  we  can  express  the  time  behavior 


SEC.    8-1    I    FACTORS    AFFECTING    REACTION    RATES 


125 


of  the  reaction  in  terms  of  the  appearance  of 
either  product,  C02  or  NO.  Which  is  used  de- 
pends upon  convenience  of  measurement.  If  the 
experimenter  prefers  to  measure  the  production 
of  carbon  dioxide,  he  would  express  the  rate  in 
the  form 

_  quantity  CO?  produced 
time  interval 

=  quantity  COj  produced  per  unit  time 


The  quantity  consumed  or  produced  is  conven- 
iently expressed  in  partial  pressure  units  if  the 
substance  is  a  gas.  Concentration  units  are  con- 
venient if  the  reactant  or  product  is  in  solution. 
The  time  measurement  is  also  expressed  in  what- 
ever units  fit  the  reaction:  microseconds  for  the 
explosion  of  household  gas  and  oxygen,  seconds 
or  minutes  for  the  burning  of  a  candle,  days  for 
the  rusting  of  iron,  months  for  the  rotting  of 
wood. 


8-1  FACTORS  AFFECTING  REACTION  RATES 


In  the  laboratory  you  have  observed  the  reaction 
of  ferrous  ion,  Fe+2(aq),  with  permanganate  ion, 
Mn04~  (aq),  and  also  the  reaction  of  oxalate  ion, 
CiQi2(aq),  with  permanganate  ion,  Mn04  (aq). 
These  studies  show  that  the  rate  of  a  reaction 
depends  upon  the  nature  of  the  reacting  substances. 
In  Experiment  14,  the  reaction  between  I03"  and 
HS03~  shows  that  the  rate  of  a  reaction  depends 
upon  concentrations  of  reactants  and  on  the  tem- 
perature. Let  us  examine  these  factors  one  at  a 
time. 

8-1.1    The  Nature  off  the  Reactants 

Compare  the  following  three  reactions,  all  of 
which  occur  in  water  solutions: 

5Q04-  2(aq)  +  2Mn04-  (aq)  +  \6H+(aq)  — *• 

\0CO,(g)  +  2Mn+i(aq)  +  8H,0     slow    (2) 

5Fe+2(aqJ  +  Mn04-  (aq)  +  SH+(aq)  — *- 

5Fe+3(aq)  +  Mn+2(aqJ  +  4H20     very  fast    (J) 

Fe+2(aqJ  +  Ce+YaqJ  — >■ 

Ft+*(aq)  +  Ce+Yaqj    very  fast    (4) 

Both  ferrous  ion,  Fe**(aq)  and  oxalate  ion, 
C204~ 2(aq),  have  the  capability  of  decolorizing 
a  solution  containing  permanganate  ion  at  room 
temperature.  Yet,  there  is  a  great  contrast  in  the 
time  required  for  the  decoloration.  The  differ- 
ence lies  in  specific  characteristics  of  Ft+2(aq) 
and  Ci0^2(aq).  On  the  other  hand,  Fe+2(aq)  is 
also  changed  to  Fe+'6(aq)  by  reacting  either  with 
Mn04"  (aq)  or  with  eerie  ion,  Ct+4(aq).  One  of 
these  reactions  is  simple  and  the  other  involves 


many  molecules.  Yet,  these  are  both  rapid  re- 
actions. 

Here  are  two  reactions  that  take  place  in  the 
gas  phase. 

2NO  +  0>  — >-  2NO,   moderate  at  20°C  (5) 

CH4  +  202  — »-  C02  +  2H>0 

extremely  slow  at  20°C    (<5) 

The  oxidation  of  nitric  oxide,  NO,  is  a  reaction 
involved  in  smog  production.  It  is  moderately 
rapid  at  normal  temperatures.  The  oxidation  of 
methane,  CH4  (household  gas),  however,  occurs 
so  slowly  at  room  temperature  that  we  may  say 
that,  for  all  practical  purposes  it  doesn't  react  at 
all.  Again,  the  difference  in  the  reaction  rates 
must  depend  upon  specific  characteristics  of  the 
reactants,  NO  and  CH4. 

The  determination  of  the  molecular  character- 
istics which  are  important  in  rate  behavior  is  an 
interesting  frontier  of  chemistry.  It  seems  that 
chemical  reactions  which  involve  the  breaking  of 
several  chemical  bonds  and  the  formation  of  new 
chemical  bonds  tend  to  proceed  slowly  at  room 
temperature.  Reaction  (2)  is  of  this  type— there 
are  many  bonds  which  must  be  broken  in  the 
five  C204-2  ions  and  the  two  Mn04~  ions  to  form 
the  10CO2  and  2Mn+2.  This  reaction  might  be 
expected  to  proceed  slowly  (as  it  does).  Reac- 
tion (6)  also  involves  breaking  of  bonds  and 
forming  of  new  bonds  and  it  is  slow  at  room 
temperature.  In  contrast,  reaction  (3)  is  very 
rapid  though  it  involves  breaking  of  chemical 
bonds  and  it  might  be  expected  to  proceed  slowly. 


126 


THE    RATES   OF    CHEMICAL    REACTIONS   I   CHAP.    8 


We  see  that  we  cannot  be  certain  of  a  prediction 
that  a  reaction  might  be  slow.  Reaction  (4)  ap- 
parently does  not  require  bond  breaking  or  bond 
formation.  It  can  be  expected  to  be  rapid  (as  it 
is).  A  prediction  of  this  type  is  usually  reliable. 
Reaction  (5)  requires  breaking  of  but  one  bond 
and  the  formation  of  two.  It  has  a  moderate 
reaction  rate,  rapid  at  high  pressures  and  slow 
at  low  pressures. 

These  and  other  examples  lead  to  the  follow- 
ing rules: 

(1)  Reactions  that  do  not  involve  bond  rear- 
rangements are  usually  rapid  at  room  tem- 
perature. 

(2)  Reactions  in  which  bonds  are  broken  tend 
to  be  slow  at  room  temperature. 

We  can  say  little  more  about  how  the  nature 
of  the  reactants  determines  the  reaction  rate 
until  we  consider  in  detail  how  some  reactions 
take  place.  For  the  time  being,  it  will  suffice  to 
observe  that  this  is  an  active  field  of  study  and 
much  remains  to  be  learned. 


EXERCISE  8-1 

Are  any  of  the  following  three  reactions  likely 
to  be  extremely  rapid  at  room  temperature?  Are 
any  likely  to  be  extremely  slow  at  room  tempera- 
ture? Explain. 

(a)  C^(aq)  +  Fe+3(aq)  — ►- 

Ci+3(aq)  +  Fe+i(aq) 

(b)  3Fe+2(agJ  +  ^O^(aq)  +  4H+(aq)  — ►- 

3Fc+3(aq)  +  NOfeJ  +  2H20 

(c)  C8HU  +  \2h02(g)  — k  %CO,(g)  +  9H2OfeJ 

liquid 
gasoline 


8-1.2    Effect  of  Concentration: 
Collision  Theory 

Henceforth  we  shall  concentrate  our  attention  on 
one  reaction  at  a  time.  The  nature  of  the  reac- 
tants will  be  held  constant  while  the  other  factors 
that  affect  rates  are  considered.  The  first  of  these 
factors  is  concentration. 
Chemists  have  learned  that,  for  many  reac- 


tions, raising  the  concentration  of  a  reactant 
increases  the  reaction  rate.  Not  infrequently, 
though,  there  will  be  no  effect.  In  this  section  we 
shall  consider  how  a  rate  increase  with  rising 
reactant  concentration  is  explained.  In  Section 
8-1.3  we  shall  explore  why  some  reactions  pro- 
ceed at  a  rate  independent  of  the  concentration 
of  one  or  more  reactants.  Both  explanations  are 
based  upon  a  model  of  the  way  chemical  reac- 
tions take  place  on  the  molecular  scale. 

In  the  molecular  view  of  matter,  it  is  natural 
to  assume  that  two  molecules  must  come  close 
together  in  order  to  react.  Therefore,  we  postu- 
late that  chemical  reactions  depend  upon  colli- 
sions between  the  reacting  particles — atoms, 
molecules,  or  ions.  This  model  of  reaction  rate 
behavior  is  called  the  collision  theory.  It  pro- 
vides a  successful  basis  for  understanding  the 
effect  of  concentration.  Just  as  an  increase  of  the 
number  of  cars  in  motion  on  a  highway  leads 
to  a  higher  rate  of  formation  of  dented  fenders, 
increasing  the  number  of  particles  in  a  given 
volume  gives  more  frequent  molecular  collisions. 
The  higher  frequency  of  collisions  results  in  a 
higher  rate  of  reaction. 

Consider  a  homogeneous  system — one  in 
which  all  components  are  in  the  same  phase. 
According  to  the  collision  theory,  we  can  expect 
that  increasing  the  concentration  of  one  or  more 
reactants  will  result  in  an  increase  in  the  rate  of 
the  reaction.  Lowering  the  concentration  has  the 
opposite  effect.  This  is  exactly  the  behavior  found 
in  the  reaction  between  HS03"  (aq)  and  I03"  (aq) 
when  the  concentrations  are  varied  by  adding  or 
removing  reactants  or  solvent  (Experiment  14). 
In  gases,  also  homogeneous  systems,  the  con- 
centration of  an  individual  reactant  can  be 
raised  by  admitting  more  of  that  substance  into 
the  mixture.  The  concentrations  of  all  gaseous 
components  can  be  raised  simultaneously  by  de- 
creasing the  volume  occupied  by  the  mixture. 
Decreasing  the  volume  by  compressing  the  gas 
raises  the  concentration  of  all  reactants,  hence 
increases  the  rates  of  reactions  taking  place.  In- 
creasing the  volume  by  expanding  the  gas  has 
the  opposite  effect  on  concentrations,  hence  de- 
creases reaction  rates. 

In  a  heterogeneous  reaction  system,  the  com- 


SEC.    8-1    I    FACTORS    AFFECTING    REACTION    RATES 


127 


Fig.  8-1.  The  number  of  collisions  per  second  depends 
upon  concentration. 


ponents  are  in  two  or  more  different  phases.  As 
an  example,  consider 

burning 

wood  (solid)  +  oxygen  (gas) >- 

carbon  dioxide  (gas)  +  water  (gas) 

In  a  system  of  this  sort,  the  rate  of  the  reaction 
depends  upon  the  amount  of  interface  between 
the  phases,  or,  in  other  words,  the  area  of  con- 
tact between  them.  For  example,  a  log  burns  in 
air  at  a  relatively  slow  rate.  If  the  amount  of 
exposed  surface  of  the  wood  is  increased  by  re- 
ducing the  log  to  splinters,  the  burning  is  much 
more  rapid.  If,  further,  the  wood  is  reduced  to 
fine  sawdust  and  the  latter  is  suspended  in  a 
current  of  air,  the  combustion  takes  place  ex- 
plosively. Where  one  of  the  reactants  is  a  gas, 
such  as  in  the  above  example,  the  concentration 
of  the  gas  is  also  a  factor.  A  piece  of  wood  burns 
much  more  rapidly  in  pure  oxygen  than  it  does 
in  ordinary  air,  in  which  the  oxygen  makes  up 
only  about  20%  of  the  mixture. 

We  see  that  the  collision  theory  provides  a 
good  explanation  of  reaction  rate  behavior.  It  is 
quite  reasonable  that  the  reaction  rate  should 
depend  upon  collisions  among  the  reactant  mole- 
cules. In  fact,  it  is  so  reasonable  that  we  are  left 
wondering  why  the  concentrations  of  some  re- 
actants in  some  reactions  do  not  affect  the  rate. 


The  explanation  is  found  in  the  detailed  steps 
by  which  the  reaction  takes  place,  the  reaction 
mechanism. 

8-1.3    Reaction  Mechanism 

As  has  been  proposed,  in  order  for  a  chemical 
reaction  to  occur,  particles  must  collide.  The 
particles  may  be  atoms,  molecules,  or  ions.  As 
a  result  of  collisions,  there  can  be  rearrangements 
of  atoms,  electrons,  and  chemical  bonds,  with 
the  resultant  production  of  new  species.  As  an 
example,  let  us  take  another  look  at  the  reaction 
between  Fe+2  and  MnO*-  in  acid  solution: 

5Ft+2(aq)  +  Mn04"  (aq)  +  SH+(aq)  — >- 

5Fe+3(aq)  +  Mn+2(aqJ  +  4H20    (7) 

The  equation  indicates  that  one  Mn04"  ion,  five 
Fe+2  ions,  and  eight  H+  ions  (a  total  of  four- 
teen ions)  must  react  with  each  other.  If  this 
reaction  were  to  take  place  in  a  single  step,  these 
fourteen  ions  would  have  to  collide  with  each 
other  simultaneously.  The  probability  of  such  an 
event  occurring  is  extremely  small — so  small  that 
a  reaction  which  depended  upon  such  a  collision 
would  proceed  at  a  rate  immeasurably  slow. 
Since  the  reaction  occurs  at  an  easily  measured 
rate,  it  must  proceed  by  some  sequence  of  steps, 
none  of  which  involves  such  an  improbable  col- 
lision. 

As  a  matter  of  fact,  chemists  regard  the  colli- 
sion of  even  four  molecules  as  an  extremely 
improbable  event  if  they  are  at  low  concentration 
or  if  they  are  in  the  gas  phase.  We  conclude  that 


128 


THE  RATES  OF  CHEMICAL  REACTIONS  I  CHAP.  8 


a  complex  chemical  reaction  which  proceeds  at 
a  measurable  rate  probably  takes  place  in  a 
series  of  simpler  steps.  The  series  of  reaction 
steps  is  called  the  reaction  mechanism. 

Consider  the  oxidation  of  gaseous  hydrogen 
bromide,  HBr,  a  reaction  that  is  reasonably 
rapid  in  the  temperature  range  from  400  to 
600°C: 

4HBrfgJ  +  02(g)  — ►  2H20(g)  +  2Br2(g)    (8) 

By  the  collision  theory,  we  expect  that  increasing 
the  partial  pressure  (and  thus,  the  concentration) 
of  either  the  HBr  or  02  will  speed  up  the  reaction. 
Experiments  show  this  is  the  case.  Quantitative 
studies  of  the  rate  of  reaction  (8)  at  various  pres- 
sures and  with  various  mixtures  show  that  oxy- 
gen and  hydrogen  bromide  are  equally  effective 
in  changing  the  reaction  rate.  However,  this  re- 
sult raises  a  question.  Since  reaction  (8)  requires 
four  molecules  of  HBr  for  every  one  molecule 
of  02,  why  does  a  change  in  the  HBr  pressure 
have  just  the  same  effect  as  an  equal  change  in 
the  02  pressure? 

The  explanation  is  found  by  considering  the 
details  of  the  process  by  which  reaction  (8)  oc- 
curs. The  overall  reaction  brings  together  five 
molecules,  four  of  HBr  and  one  of  02.  However, 
the  chance  that  five  gaseous  molecules  will  col- 
lide simultaneously  is  practically  zero.  The  reac- 
tion must  occur  in  a  sequence  of  simpler  steps. 

All  of  the  studies  of  reaction  (8)  are  explained 
by  the  following  series  of  reactions: 


HBr  +  02 

HOOBr  +  HBr 

HOBr  +  HBr 


HOOBr         slow      (9) 
2HOBr  fast      (70) 

H20  +  Br2    fast      (77) 


First,  observe  that  adding  reactions  (9)  and  (70) 
plus  twice  reaction  (11)  gives  the  overall  reac- 
tion (8).  Next,  we  see  that  each  step  in  the  se- 
quence requires  only  two  molecules  to  collide. 
Finally,  the  proposal  that  reaction  (9)  is  slow 
whereas  reactions  (10)  and  (//)  are  fast  explains 
why  HBr  and  02  have  the  same  effect  on  the 
reaction  rate. 

Reaction  (9)  is  a  "bottle-neck"  in  the  oxida- 
tion of  hydrogen  bromide.  As  fast  as  HOOBr  is 
formed  by  this  slow  reaction  it  is  consumed  in 
the  rapid  reaction  (10).  But  no  matter  how  rapid 


reactions  (70)  and  (77)  are,  they  can  produce 
H20  and  Br2  only  as  fast  as  the  slowest  reaction 
in  the  sequence.  Hence,  the  factors  that  deter- 
mine the  rate  of  reaction  (9)  determine  the  rate 
of  the  overall  process. 

The  sequence  of  reactions  (9),  (70),  and  (77) 
is  called  the  reaction  mechanism  of  the  overall 
reaction  (8).  Because  it  is  the  slowest  reaction  in 
the  mechanism,  reaction  (9)  is  the  step  that  fixes 
the  rate.  The  slowest  reaction  in  a  reaction  mecha- 
nism is  called  the  rate  determining  step. 

There  are  two  features  of  this  example  that 
are  rather  common.  First,  none  of  the  steps  in 
the  reaction  mechanism  requires  the  collision  of 
more  than  two  particles.  Most  chemical  reactions 
proceed  by  sequences  of  steps,  each  involving  only 
two-particle  collisions.  Second,  the  overall  or  net 
reaction  does  not  show  the  mechanism.  In  gen- 
eral, the  mechanism  of  a  reaction  cannot  be  de- 
duced from  the  net  equation  for  the  reaction;  the 
various  steps  by  which  atoms  are  rearranged  and 
recombined  must  be  determined  through  experi- 
ment. 


EXERCISE  8-2 

Imagine  five  people  working  together  to  wash  a 
stack  of  very  greasy  dishes.  The  first  two  clear 
the  table  and  hand  the  dishes  to  the  third  person 
who  washes  them  and  hands  them  on.  The  last 
two  persons  dry  and  stack  them.  Which  step  is 
likely  to  be  the  rate  determining  step?  In  the 
light  of  your  answer,  discuss  how  the  rate  of  the 
overall  process  would  be  affected  if  a  sixth  person 
joined  the  group  (a)  as  a  table  clearer;  (b)  as  a 
second  dishwasher;  (c)  as  a  dish  dryer. 


8-1.4    The  Quantitative  Effect  of  Concentration 

The  reaction  mechanism  is  deduced  from  quantitative 
studies  of  the  dependence  of  the  rate  upon  the  concentra- 
tions or  pressures  of  the  various  reactants.  To  interpret 
such  studies,  we  need  to  develop  our  collision  theory 
model. 

Consider  the  reaction  between  gaseous  hydrogen,  H2, 
and  gaseous  iodine,  I»: 


Ht(g)  +  U(g) 


201(g) 


(12) 


SEC.    8-1    I    FACTORS    AFFECTING    REACTION    RATES 


129 


Each  time  a  molecule  of  H2  collides  with  an  iodine  mole- 
cule, reaction  may  occur.  The  frequency  of  these  en- 
counters, for  a  particular  H2  molecule,  is  determined  by 
how  many  b  molecules  are  present.  Doubling  the  number 
of  I2  molecules  per  unit  volume  would  just  double  the 
collisions.  Tripling  the  number  of  h  molecules  per  unit 
volume  would  triple  the  collisions.  Since  the  iodine  partial 
pressure  fixes  the  iodine  concentration,  the  rate  of  the 
reaction  is  proportional  to  the  iodine  partial  pressure: 


(rate)  is  proportional  to 


iodine 

partial 

_pressure. 


(13) 


In  the  same  way,  a  particular  iodine  molecule  must  find 
a  hydrogen  molecule  to  react.  The  rate  of  the  reaction  is 
proportional  to  the  partial  pressure  of  the  hydrogen: 


(rate)  is  proportional  to 


hydrogen 
partial 
_  pressure 


(14) 


In  view  of  (13)  and  (14),  the  rate  must  be  proportional 
to  the  product  of  the  partial  pressure  of  iodine  and 
hydrogen : 


(rate)  is  proportional  to 


"hydrogen 

partial 
_  pressure 

X 

"iodine 
partial 
pressure 

In  symbols,  we  can  write 

rate  = 

=  k[pHi]  X 

>lJ 

U5) 


(16) 


8-1.5     Effect  of  Temperature:  Collision  Theory 

In  Experiments  12  and  14,  you  discovered  that 
temperature  has  a  marked  effect  upon  the  rate 
of  chemical  reactions.  Thus,  raising  the  tempera- 
ture speeded  up  the  reaction  between  103"  and 
HSO3".  That  is  the  same  effect,  qualitatively,  that 
is  observed  in  the  reaction  of  a  candle  with  air. 
The  match  "lighted"  the  candle  by  raising  its 
temperature  (at  the  wick).  Once  started,  the  re- 
action of  combustion  releases  enough  heat  to 
keep  the  temperature  high,  thus  keeping  the  re- 
action going  at  a  reasonable  rate.  Raising  the 
temperature  speeded  up  the  reaction.  The  same 
type  of  explanation  applies  to  the  explosion  of 
a  kitchen  full  of  household  gas  and  air  when  a 
cigarette  is  brought  into  the  room.  Around  the 
glowing  tip  of  the  cigarette  the  gas  temperature 
is  raised.  At  this  locale,  the  reaction  speeds  up, 
liberating  heat.  This  heat  warms  the  nearby  re- 
gion even  more  and  the  reaction  goes  somewhat 


faster.  This  acceleration  continues  until  finally 
(in  a  millisecond  or  so)  it  reaches  an  ex- 
plosive rate — the  most  rapid  reaction  permitted 
by  the  collisional  properties  of  the  gas.  Raising 
the  temperature  started  it  all  by  speeding  up  the 
reaction. 

In  all  of  these  reactions  (and  in  almost  all 
others),  increasing  the  temperature  has  a  very 
pronounced  effect,  always  speeding  up  the  reac- 
tion. Two  questions  come  to  mind.  "Why  does 
a  temperature  rise  speed  up  a  reaction?"  and 
"Why  does  a  temperature  rise  have  such  a  large 
effect?"  To  answer  these  questions,  we  return  to 
our  collision  theory. 

From  what  we  know  about  molecular  sizes, 
we  can  calculate  that  a  particular  CH4  molecule 
collides  with  an  oxygen  molecule  about  once 
every  one-thousandth  of  a  microsecond  (10-9 
seconds)  in  a  mixture  of  household  gas  (methane, 
formula  CH4)  and  air  under  normal  conditions. 
This  means  that  every  second  this  methane  mole- 
cule encounters  109  oxygen  molecules!  Yet  the 
reaction  does  not  proceed  noticeably.  We  can 
conclude  either  that  most  of  the  collisions  are 
ineffective  or  that  the  collision  theory  is  not  a 
good  explanation.  We  shall  see  that  the  former 
is  the  case — we  can  understand  why  most  colli- 
sions might  be  ineffective  in  terms  of  ideas  that 
are  consistent  with  the  collision  theory. 

Chemists  have  learned  that  chemical  reactions 
occur  when  collisions  occur  but  only  when  the 
collision  involves  more  than  a  certain  amount  of 
energy.  We  can  understand  this  by  returning  to 
our  analogy  of  cars  bumping  each  other  on  a 
highway.  In  a  line  of  heavy  traffic  one  frequently 
receives  gentle  bumps  from  the  car  in  front  or 
the  car  behind.  No  damage  is  done  to  the  ears- 
only to  tempers.  But  occasionally  a  high  speed 
collision  occurs.  If  this  occurs  with  enough  en- 
ergy, a  bumper  may  be  knocked  off  a  car  and  a 
fender  may  be  dented.  It  is  the  high  energy  col- 
lisions which  cause  the  auto  damage  and  it  is 
high  energy  molecular  collisions  which  cause  the 
"molecular  damage"  that  we  call  a  chemical  re- 
action. Just  as  a  certain  amdunt  of  energy  is 
required  to  break  loose  a  bumper,  a  certair 
amount  of  energy  is  required  to  cause  a  chemical 
reaction.  In  either  instance,  if  there  is  more  than 
this  "threshold  energy,"  the  reaction  can  occur 
and  if  there  is  less,  it  cannot. 


130 


THE    RATES    OF    CHEMICAL    REACTIONS    I    CHAP.    8 


Collimator 


Tin  vapor 
Molten  ti 


Oven,   at 

controlled 

temperature 


Fig.  8-2.  Rotating   disc   jur  measurement  of   atomic  velocities. 


SEC.    8-1    I    FACTORS    AFFECTING    REACTION    RATES 


131 


w 


9  3  7  6  5 

Pie    slice    numher 


Disc  D2    after  many   revolutions' 


Increasing  velocity   of  tin    atoms 


Fig.  8-3.  Distribution  of  atomic  (or  molecular)  veloc- 
ities from  the  rotating  disc. 


DISTRIBUTION    OF    KINETIC    ENERGIES 

This  discussion  of  threshold  energy  causes  us  to 
wonder  what  energies  are  possessed  by  molecules 
at  a  given  temperature.  We  have  already  com- 
pared the  molecules  of  a  gas  with  billiard  balls 
rebounding  on  a  billiard  table.  When  billiard 
balls  bounce  around,  colliding  with  each  other, 
some  of  them  move  rapidly  and  some  slowly.  Do 
molecules  behave  this  way?  Experiment  provides 
the  answer. 

Figure  8-2  shows  a  device  for  measuring  the 
distribution  of  atomic  or  molecular  velocities. 
It  consists  of  two  discs,  A  and  A,  rotating 
rapidly  on  a  common  axle.  They  rotate  in  a 
vacuum  chamber  in  front  of  an  oven  containing 
molten  tin  and  held  at  a  controlled  temperature. 
Vapor  streams  out  of  the  small  opening  in  the 
oven  and  strikes  the  rotating  disc,  A.  When  the 
disc  has  rotated  to  the  position  shown  in  Figure 
8-2B,  a  small  amount  of  gas  has  passed  through 
the  slot  in  disc  A-  A  short  time  later,  shown  in 
Figure  8-2C,  the  atoms  of  tin  have  traveled  part 
of  the  way  toward  the  second  rotating  disc.  The 
fastest  moving  atoms  have  traveled  farther  than 
the  others — they  are  leading  the  way.  The  slowest 
moving  atoms  are  beginning  to  lag  behind.  Still 
later,  Figure  8-2D,  the  atoms  have  spread  out  in 
space  even  more,  and  the  fastest  atoms  have 
already  reached  the  second  rotating  disc.  Now 
as  the  atoms  reach  disc  A,  they  condense  on  it. 
The  position  where  a  given  atom  condenses  on 
disc  A  depends  upon  how  long  that  atom  took  to 


travel  from  A  to  A  and  how  fast  A  is  rotating. 

As  the  slotted  disc  A  lets  through  burst  after 
burst  of  tin  atoms,  a  layer  of  tin  builds  up  on  the 
surface  of  disc  A-  The  pattern  of  this  layer  is 
determined  by  the  distribution  of  velocities  of 
the  atoms  escaping  from  the  oven  at  tempera- 
ture T.  Figure  8-3  shows  the  disc  A  divided  into 
sections,  like  slices  of  a  pie.  The  fastest  moving 
atoms  are  condensed  on  pie  slices  3,  4,  and  5. 
The  slowest  moving  atoms  are  condensed  on  the 
pie  slices  10  and  1 1 .  If  the  disc  is  cut  up  and  each 
slice  weighed,  the  amount  of  tin  can  be  deter- 
mined. A  plot  of  the  weight  of  tin  against  the 
pie  slice  number  indicates  the  distribution  of 
atomic  velocities. 

Clearly,  the  plot  of  Figure  8-3  contains  infor- 
mation about  the  distribution  of  kinetic  energies. 
From  the  rate  of  rotation  of  the  discs  and  the 
distance  between  them  we  can  calculate  the  ve- 
locity an  atom  must  have  to  condense  on  a 
particular  pie  slice.  From  the  atomic  mass  and 
its  velocity,  we  learn  the  atom's  kinetic  energy. 
Figure  8-4  shows  the  result.  At  a  temperature  Ti 


Fig.  8-4.  Effect  of  temperature  on  atomic  {or  molec- 
ular) kinetic  energy  distribution. 


Tj  greater  than  7} 


Kinetic   energy 


132 


THE    RATES    OF    CHEMICAL    REACTIONS       CHAP. 


a  few  atoms  have  very  low  kinetic  energies  and 
some  have  very  high  kinetic  energies.  Most 
of  them  have  intermediate  kinetic  energies,  as 
shown  by  the  solid  curve.  At  a  higher  tempera- 
ture, r2,  the  energy  distribution  is  altered  to  that 
shown  as  a  dashed  curve.  As  can  be  seen,  increas- 
ing the  temperature  causes  a  general  shift  of  the 
distribution  toward  one  with  more  molecules 
having  high  kinetic  energies.  Moreover,  in  going 
from  Tx  to  T2  there  is  a  large  increase  in  the 
number  of  molecules  having  kinetic  energies 
above  a  certain  value. 

THRESHOLD  ENERGY  AND  REACTION  RATE 

We  can  apply  these  curves  to  our  reaction  rate 
problem.  Suppose  a  reaction  can  proceed  only  if 
two  molecules  collide  with  kinetic  energy  exceed- 
ing a  certain  threshold  energy,  E.  Figure  8-4 
shows  us  a  typical  situation.  At  Tx  the  darkly 
shaded  area  is  proportional  to  the  number  of  the 
molecules  which  possess  this  energy  or  more. 
Since  only  a  small  number  of  molecules  have  as 


much  as  this  energy,  few  collisions  are  effective, 
and  the  reaction  is  slow.  But  if  we  raise  the  tem- 
perature to  T2,  the  number  of  molecules  with 
energy  E  or  greater  is  raised  in  proportion  to  the 
diagonally  shaded  area.  Only  a  small  tempera- 
ture change  is  needed  to  make  a  large  change  in 
the  area  out  on  the  tail  of  the  energy  distribution 
curve.  Consequently,  the  reaction  rate  is  very 
sensitive  to  change  in  temperature. 

This  argument  is  based  upon  the  "typical"  situation  in 
which  E  is  well  out  on  the  tail  of  the  curve.  Suppose  it  is 
not;  suppose  E  is  near  the  maximum  of  the  curve  at  7*i, 
or  is  even  to  the  left  of  it.  Then  a  large  number  of  the 
molecules  have  the  requisite  energy,  even  at  the  lower 
temperature,  Tt.  Since  collisions  occur  so  rapidly  (re- 
member, one  every  10~9  second  or  so),  the  reaction  is 
over  in  a  blink  of  the  eye.  This  reaction  would  be  called 
"instantaneous."  The  circumstances  shown  in  Figure  8-4 
are  "typical"  only  of  a  slow  reaction. 

It  should  be  remarked  that  raising  the  temperature  also 
increases  the  rate  by  increasing  the  frequency  of  the 
collisions.  This  is,  however,  a  very  small  effect  compared 
with  that  caused  by  the  increase  in  the  number  of  mole- 
cules with  sufficient  energy  to  cause  reaction. 


8-2  THE  ROLE  OF  ENERGY  IN  REACTION  RATES 


Now  that  we  have  this  evidence  about  the  mo- 
lecular velocity  distribution,  we  can  see  how 
temperature  changes  the  fraction  of  the  molecu- 
lar collisions  involving  an  energy  exceeding  the 
threshold  energy  E.  Now  our  understanding  ot 
the  role  of  energy  in  fixing  reaction  rates  can  be 
expanded. 

8-2.1    Activation  Energy 

Suppose  you  wished  to  make  an  automobile  trip 
from  Los  Angeles  to  San  Francisco.  This  trip  re- 
quires that  the  Tehachapi  Mountains  be  crossed. 
As  shown  in  Figure  8-5,  this  can  be  accomplished 
by  taking  the  highway  through  the  Tejon  Pass. 
Of  course,  the  high  altitude  of  this  pass,  4200 
feet,  makes  this  part  of  the  trip  the  slowest  and 
the  biggest  test  of  your  automobile.  This  is  the 
point  in  the  trip  where  radiator  fluids  boil  and 
engine  troubles  develop.  This  is  the  point  where 
the  older  cars  turn  back  and  limp  home.  Yet, 
this  is  the  lowest  pass  in  these  mountains,  hence 
the  most  favorable  route.  If  your  car  will  sur- 


mount a  4200  foot  pass,  it  will  undoubtedly 
make  the  trip  to  San  Francisco. 

Chemical  reactions  are  similar.  As  molecules 
collide  and  reaction  takes  place,  the  atoms  must 
take  up,  momentarily,  bonding  arrangements 
that  are  less  stable  than  either  reactants  or  prod- 
ucts. These  high  energy  molecular  arrangements 
are  like  the  mountain  pass — they  place  an  energy 
barrier  between  reactants  and  products.  Only  if 
the  colliding  molecules  have  enough  energy  to 
surmount  the  barrier  imposed  by  the  unstable 
arrangements  can  reaction  take  place.  This  bar- 
rier determines  the  "threshold  energy"  or  mini- 
mum energy  necessary  to  permit  a  reaction  to 
occur.  It  is  called  the  activation  energy. 

This  barrier  can  be  shown  graphically  by  am- 
plifying Figure  7-1  (p.  110),  which  showed  the 
relative  energies  of  reactants  and  products.  In 
Figure  8-6  we  see  that  the  diagram  becomes  the 
equivalent  of  the  road  map  in  our  automobile 
trip  analogy.  This  diagram  applies  to  the  reac- 
tion between  carbon  monoxide,  CO,  and  nitro- 
gen dioxide,  N02,  to  produce  carbon  dioxide, 


SEC.    8-2    I    THE    ROLE    OF    ENERGY    IN    REACTION    RATES 


133 


SAN 


LOS  ANGELES 


4200 


LA 


Travel     coordirvate 
-»-        Ter'on    Pass       


SF 


Fig.  8-5.  Crossing  the  mountains  between  Los  Angeles 
and  San  Francisco. 


C02,  and  nitric  oxide,  NO.  The  horizontal  axis 
of  the  diagram,  called  the  reaction  coordinate, 
shows  the  progress  of  the  reaction.  Proceeding 


from  left  to  right  along  this  reaction  coordinate 
signifies  the  CO  and  N02  molecules  approaching 
each  other,  colliding,  and  going  through  inter- 
mediate processes  of  reaction  which  result  in  the 
formation  of  C02  and  NO,  and,  finally,  the  sepa- 
ration of  the  latter  two  molecules.  The  vertical 
axis  represents  the  total  potential  energy  of  the 


134 


THE    RATES    OF    CHEMICAL    REACTIONS    I    CHAP.    8 


Potential 

energy, 

kcaly/mole 


+SO 

j^ZcnWTED  COMPLEX 

- 

Activation 
energy 

/S^ 

Q 

— 

CO  -r  tfOz 

Net  energy  released 

-50 

sv^cq2  *  no 

Reaction    coordinate 


Reactants 


Activated   complex. 


-*-        Products 


Fig.  8-6.  Potential   energy   diagram   for  the  reaction 
CO  +  N02  — >-  C02  +  NO. 


system.  Thus,  the  curve  provides  a  history  of  the 
potential  energy  change  during  a  collision  which 
results  in  a  reaction.  The  energy  required  to  sur- 
mount the  potential  energy  "barrier"  to  reaction 
usually  is  provided  by  the  kinetic  energies  of  the 
colliding  particles,  as  fixed  by  the  temperature. 
Let  us  move  from  left  to  right  along  this  curve 
and  describe  the  events  which  occur.  Along  the 
flat  region  at  the  left,  CO  and  N02  are  approach- 
ing each  other.  In  this  region,  they  possess  kinetic 
energy  and  their  total  potential  energy  shows  no 
change.  The  beginning  of  the  rise  in  the  curve 
signifies  that  the  two  molecules  have  come  suffi- 
ciently close  to  have  an  effect  on  each  other. 
During  this  approach,  the  molecules  slow  down 
as  their  kinetic  energies  furnish  the  potential 
energy  to  climb  the  curve.  If  they  have  sufficient 
kinetic  energy,  they  can  ascend  the  left  side  of 
the  "barrier"  all  the  way  up  to  the  summit.  At- 
taining this  point  is  interpreted  as  follows:  CO 
and  N02  had  sufficient  kinetic  energy  to  over- 
come the  mutually  repulsive  forces  of  their  nuclei 
and  negative  electron  clouds  and  thus  come  very 
close  to  each  other.  Here  at  the  summit  the 
molecular  cluster  is  unstable  with  respect  to 
either  the  forward  reaction  (to  give  C02  and  NO) 
or  the  reverse  reaction  (to  restore  the  molecules 
of  CO  and  N02).  This  transitory  arrangement  is 
of  key  importance  (its  potential  energy  fixes  the 


activation  energy);  it  is  called   the  activated 
complex. 

Now  there  are  two  possibilities:  (1)  the  acti- 
vated complex  may  separate  into  the  two  original 
CO  and  N02  molecules,  which  would  then  re- 
trace their  former  path  on  the  curve,  or,  (2)  the 
activated  complex  may  separate  into  C02  and 
NO  molecules.  The  latter  possibility  is  repre- 
sented by  moving  down  the  right  side  of  the 
"barrier."  In  the  flat  region  at  the  right,  C02  and 
NO  have  separated  beyond  the  point  of  having 
any  effect  on  each  other  and  the  potential  energy 
of  the  activated  complex  has  become  kinetic  en- 
ergy again. 

In  the  event  that  the  CO  and  N02  molecules 
do  not  have  sufficient  energy  to  attain  the  sum- 
mit, they  reach  a  point  only  part  way  up  the  left 
side  of  the  "barrier."  Then,  repelling  one  an- 
other, they  separate  again,  going  downhill  to  the 
left. 

We  have  labeled  the  difference  between  the 
high  potential  energy  at  the  activated  complex 
and  the  lower  energy  of  the  reactants  as  the  acti- 
vation energy.  The  activation  energy  is  the 
energy  necessary  to  transform  the  reactants  into 
the  activated  complex.  This  may  involve  weaken- 
ing or  breaking  bonds,  forcing  reactants  close 
together  in  opposition  to  repulsive  forces,  or 
storing  energy  in  a  vibrating  molecule  so  that  it 
reacts  on  collision.  Increasing  the  temperature 
affects  reaction  rate  by  increasing  the  number  of 
molecular  collisions  that  involve  sufficient  energy 
to  form  this  activated  complex.  The  magnitude 


SBC.  8-2  I  THB  ROLB  OF  ENERGY  IN  REACTION  RATES 


135 


of  ihe  activation  energy  for  a  reaction  can  be 
determined  by  measuring  experimentally  the 
change  in  reaction  rate  with  temperature. 

8-2.2     Heat  of  Reaction 

We  can  deduce  the  heat  of  reaction  from  Figure 
8-6.  In  our  example,  the  reactants  are  at  a  higher 
total  energy  than  are  the  products.  This  means 
that  in  the  course  of  the  reaction  there  will  be  a 
net  release  of  energy.  This  reaction  is  exothermic. 
Figure  8-6  shows  that  the  reaction  releases  54 
kcal  of  heat  per  mole  of  carbon  monoxide  con- 
sumed. Notice  that  the  height  of  the  energy 
barrier  between  reactants  and  products  has  no 
effect  on  the  net  heat  release.  We  must  put  in  an 
amount  of  energy  equal  to  the  activation  energy 
to  get  to  the  top  of  the  barrier  but  we  get  it  all 
back  on  the  way  down  the  other  side. 

Now  let  us  consider  the  reverse  reaction.  We 
need  not  draw  another  reaction  diagram,  since 
Figure  8-6  will  suffice.  Now  we  are  interested  in 
the  reaction  between  C02  and  NO  to  produce 
CO  and  N02: 

CO  +  N02  -<—  C02  +  NO  (17) 

This  reaction  begins  at  the  lower  energy  appro- 
priate to  the  chemical  stability  of  C02  +  NO 
(at  the  right  side  of  Figure  8-6)  and  ends  at  the 
higher  energy  appropriate  to  the  chemical  sta- 
bility of  CO  +  N02  (at  the  left  side  of  Figure 
8-6).  The  difference  in  energy,  the  heat  of  this 
reaction,  is  just  equal  to  that  of  the  reverse 
reaction  but  is  opposite  in  sign.  This  reaction 
absorbs  54  kcal  of  heat  per  mole  of  carbon 
monoxide  produced.  It  is  endothermic. 

Figure  8-6  contains  one  other  very  interesting 
piece  of  information  concerning  the  rate  of  the 
reverse  reaction,  (77).  This  reaction  rate  is  con- 
trolled by  the  energy  barrier  confronting  the 
colliding  molecules  of  C02  and  NO.  We  see  from 
the  diagram  that  the  activation  energy  for  this 
reaction  is  higher  than  that  for  the  reaction  we 
studied  earlier.  Further,  it  is  higher  by  exactly 
the  heat  of  reaction.  We  conclude  that  the  reac- 
tion between  C02  and  NO  will  be  slower,  at  any 
given  temperature,  than  the  reverse  reaction  be- 
tween CO  and  N02,  if  the  rates  are  compared 
at  the  same  partial  pressures. 


The  relationship  between  activation  energies 
for  the  forward  and  reverse  reactions  can  be  ex- 
pressed mathematically.  The  activation  energy  is 
denoted  by  the  symbol  AHX  (read  "delta-//- 
cross")  and  the  heat  of  the  reaction  by  AH. 
Hence  we  may  write: 

AH}{  =  AHl  +  AH  (18) 

AH\i  =  activation  energy  for  reaction  pro- 
ceeding to  right  (always  endothermic) 
AHi  =  activation  energy  for  reaction  pro- 
ceeding to  left  (always  endothermic) 
AH  =  heat  absorbed  during  reaction  pro- 
ceeding left  to  right  (either  endother- 
mic or  exothermic). 

The  heat  of  reaction,  AH,  is  positive  if  heat  is 
absorbed  as  the  reaction  proceeds,  left  to  right. 
It  is  negative  if  heat  is  evolved.  In  our  example, 
CO  +  N02  — >-  C02  +  NO, 

A//fe  =  +32  kcal/mole 
AHi  =  +86  kcal/mole 
AH    =  —  54  kcal/mole 


We  see  that 


(32)  =  (86) +  (-54) 


which  is  in  accordance  with  equation  (18).  This 
relationship  is  important  because  it  implies  that 
we  need  only  two  of  the  three  quantities,  AHlR, 
AH[,  and  AH  and  then  we  can  calculate  the 
third.  For  example,  if  we  measure  the  heat  of  the 
reaction,  AH,  and  also  measure  the  rate  of  the 
reaction  between  CO  and  N02  in  order  to  deter- 
mine AH)t,  then  we  can  calculate  AH[  by  equa- 
tion (18).  From  AHXL  we  learn  something  about 
the  rate  of  reaction  of  C02  +  NO. 

8-2.3    Action  off  Catalysts 

Many  reactions  proceed  quite  slowly  when  the 
reactants  are  mixed  alone  but  can  be  made  to 
take  place  much  more  rapidly  by  the  introduc- 
tion of  other  substances.  These  latter  substances, 
called  catalysts,  are  not  used  up  in  the  reaction. 
The  process  of  increasing  the  rate  of  a  reaction 
through  the  use  of  a  catalyst  is  referred  to  as 
catalysis.  You  have  seen  at  least  one  example  of 
catalytic  action,  the  effect  of  Mn+2(aq)  in  speed- 
ing up  the  reaction  between  CiO;2(aq)  and 
Mn04"  (aq). 


136 


THE    RATES    OF    CHEMICAL    REACTIONS    I    CHAP. 


SAM  FRANCISCO 


TEJON  PASS 
A  ttitude     4,  2.  OOft.  I 


LOS  ANGELES 


Travel     coor-din,ate 


Fig.  8-7.  An  alternate,  easier  route  through  the  moun- 
tains between  Los  Angeles  and  San  Fran- 
cisco. 


The  action  of  a  catalyst  can  be  explained  in 
terms  of  our  mountain  pass  analogy.  In  Figure 
8-5  we  see  a  formidable  mountain  pass  obstruct- 


ing travel  between  Los  Angeles  and  San  Fran- 
cisco. Many  people  who  would  like  to  take  this 
trip  cannot  because  their  automobile  isn't  up  to 
this  much  of  a  climb.  For  this  reason,  an  alter- 
nate route  was  improved.  The  coast  route,  shown 
in  Figure  8-7  is  somewhat  longer  but  it  is  easier 
because  the  highest  pass,  Gaviota  Pass,  is  only 


SEC.  8-2  I  THE  ROLE  OF  ENERGY  IN  REACTION  RATES 


137 


Potential 
energy 


Reaction   coordinates 


Fig.  8-8.  Effect  of  a  catalyst  on  a  reaction  and  its 
reverse. 

900  feet  above  sea  level.  Some  travelers  still  use 
Tejon  Pass  but,  in  addition,  the  trip  can  be  made 
via  the  easier  Gaviota  Pass.  The  net  result  is  that 
more  people  per  day  are  able  to  make  the  trip 
from  Los  Angeles  to  San  Francisco.  Of  course, 
the  lower  pass  serves  to  increase  the  rate  of  re- 
turn travel  as  well. 

Figure  8-8  shows  the  analogous  situation  for 
a  chemical  reaction.  The  solid  curve  shows  the 
activation  energy  barrier  which  must  be  sur- 
mounted for  reaction  to  take  place.  When  a 
catalyst  is  added,  a  new  reaction  path  is  provided 
with  a  different  activation  energy  barrier,  as  sug- 
gested by  the  dashed  curve.  This  new  reaction 
path  corresponds  to  a  new  reaction  mechanism 
that  permits  the  reaction  to  occur  via  a  different 
activated  complex.  Hence,  more  particles  can  get 
over  the  new,  lower  energy  barrier  and  the  rate 
of  the  reaction  is  increased.  Note  that  the  activa- 
tion energy  for  the  reverse  reaction  is  lowered 
exactly  the  same  amount  as  for  the  forward  re- 
action. This  accounts  for  the  experimental  fact 
that  a  catalyst  for  a  reaction  has  an  equal  effect 
on  the  reverse  reaction;  that  is,  both  reactions 
are  speeded  up  by  the  same  factor.  If  a  catalyst 
doubles  the  rate  in  one  direction,  it  also  doubles 
the  rate  in  the  reverse  direction. 

8-2.4    Examples  of  Catalysts 

In  all  cases  of  catalysis,  the  catalyst  acts  by  inserting 
intermediate  steps  in  a  reaction — steps  that  would  not 


occur  without  the  catalyst.  The  catalyst  itself  must  be 
regenerated  in  a  subsequent  step.  (An  added  substance 
which  is  permanently  used  by  reaction  is  a  reactant,  not 
a  catalyst.)  An  example  is  provided  by  the  catalytic  action 
of  acid  on  the  decomposition  of  formic  acid,  HCOOH. 
A  model  of  formic  acid  is  shown  in  Figure  8-9.  The 
carbon  atom  has  attached  to  it  a  hydrogen  atom,  an 
oxygen  atom,  and  an  OH  group. 

Figure  8-10  shows  how  this  molecule  might  decompose. 
If  the  hydrogen  atom  attached  to  carbon  migrates  over 
to  the  OH  group,  the  carbon-oxygen  bond  can  break  to 
give  a  molecule  of  water  and  a  molecule  of  carbon  mon- 
oxide. This  migration,  shown  in  the  center  drawing,  re- 
quires a  large  amount  of  energy.  This  means  there  is  a 
high  activation  energy.  Hence,  the  reaction  occurs  very 
slowly : 

HCOOH  — >■  H,0  +  CO  (19) 

Fig.  8-9.  A  model  of  formic  acid. 


Ball -and- Spring  Model 


Space.  -FiltiriQ  Afodel 


138 


THE    RATES    OF    CHEMICAL    REACTIONS   I    CHAP.    8 


"React-ion.    coordinate,    uncatalyzed    decomposition 


Fig. 


8-10.  Potential  energy  diagram  for  the  uncata- 
lyzed  decomposition  of  formic  acid. 


If  sulfuric  acid,  H2SO4,  is  added  to  an  aqueous  solution 
of  formic  acid,  carbon  monoxide  bubbles  out  rapidly. 
This  also  occurs  if  phosphoric  acid,  H3PO,,  is  added  in- 
stead. The  common  factor  is  that  both  of  these  acids 
release  hydrogen  ions,  H+.  Yet,  careful  analysis  shows 
that  the  concentration  of  hydrogen  ion  is  constant  during 
the  rapid  decomposition  of  formic  acid.  Evidently,  hy- 
drogen ion  acts  as  a  catalyst  in  the  decomposition  of 
formic  acid. 

Chemists  have  a  rather  clear  picture  of  how  H+  cata- 
lyzes reaction  (79).  The  availability  of  H+  in  the  solution 
makes  a  new  reaction  path  available.  The  new  reaction 
mechanism  begins  with  the  addition  of  a  hydrogen  ion  to 
formic  acid,  as  shown  in  Figure  8-11.  Thus,  the  catalyst 
is  consumed  at  first,  forming  a  new  species,  (HCOOH2)+. 
In  this  species  one  of  the  carbon-oxygen  bonds  is  weak- 
ened. With  only  a  small  expenditure  of  energy,  the  next 
reaction  shown  in  Figure  8-11  can  occur,  producing 
(HCO)+  and  H20.  Finally,  (HCO)+  decomposes  to  pro- 
duce carbon  monoxide,  CO,  and  H+.  This  last  reaction 
of  the  sequence  regenerates  the  catalyst,  H+. 

Each  of  the  steps  in  this  new  reaction  mechanism  is 
governed  by  the  same  principles  that  govern  a  simple  re- 
action. Each  reaction  has  an  activation  energy.  The  over- 
all reaction  has  a  potential  energy  diagram  that  is  merely 
a  composite  of  the  simple  energy  curves  of  the  succeeding 
steps. 


The  highest  energy  required  in  this  new  reaction  path 
is  only  18  kcal,  much  lower  than  the  activation  energy 
shown  in  Figure  8-10  for  the  uncatalyzed  reaction.  Hence 
the  rate  of  decomposition  is  much  faster  when  acid  is 
present. 

Notice  that  catalytic  action  does  not  cause  the  reaction. 
A  catalyst  speeds  up  a  reaction  that  might  take  place  in 
its  absence  but  at  a  much  lower  rate. 

In  some  cases,  the  catalyst  is  a  solid  substance  on  whose 
surface  a  reactant  molecule  can  be  held  (adsorbed)  in  a 
position  favorable  for  reaction  until  a  molecule  of  another 
reactant  reaches  the  same  point  on  the  solid.  Metals  such 
as  iron,  nickel,  platinum  and  palladium  seem  to  act  in 
this  way  in  reactions  involving  gases.  There  is  evidence 
that  in  some  cases  of  surface  adsorption,  bonds  of  re- 
actant particles  are  weakened  or  actually  broken,  thus 
aiding  reaction  with  another  reactant  particle. 

A  very  large  number  of  catalysts,  called  enzymes,  are 
found  in  living  tissues.  Among  the  best  known  examples 
of  these  are  the  digestive  enzymes,  such  as  the  ptyalin  in 
saliva  and  the  pepsin  in  gastric  juice.  A  common  func- 
tion of  these  two  enzymes  is  to  hasten  the  breakdown  of 
large  molecules,  such  as  starch  and  protein,  into  simpler 
molecules  which  can  be  utilized  by  body  cells.  In  addition 
to  the  relatively  small  number  of  digestive  enzymes,  there 
are  many  other  enzymes  involved  in  biochemical  proc- 
esses. Enzymes  are  considered  again  in  Chapter  24. 

The  specific  methods  by  which  catalysts  work  are  not 
clearly  understood  in  most  cases.  Finding  a  catalyst  suit- 
able for  a  given  reaction  usually  requires  a  long  period 


QUESTIONS    AND    PROBLEMS 


139 


Reaction    coordinat-e-,     catatyzed    decomjposi-Hon. 


of  laboratory  experimentation.  Yet,  we  can  look  forward 
to  the  time  when  a  catalyst  can  be  tailor-made  to  fit  a 
particular  need.  This  exciting  prospect  accounts  for  the 
great  activity  on  this  chemical  frontier. 


Fig.  8-11.  Potential  energy  diagram  of  the  catalyzed 
decomposition  of  formic  acid. 


QUESTIONS  AND  PROBLEMS 

1 .  The  rate  of  movement  of  an  automobile  can  be 
expressed  in  the  units  miles  per  hour.  In  what 
units  would  you  discuss  the  rate  of: 

(a)  movement  of  movie  film  through  a  pro- 
jector; 

(b)  rotation  of  a  motor  shaft; 

(c)  gain  of  altitude; 

(d)  consumption  of  milk  by  a  family; 

(e)  production  of  automobiles  by  an  auto  as- 
sembly plant. 

2.  Pick  the  member  of  each  pair  having  the  greater 
reaction  rate.  Assume  similar  conditions  within 
each  pair. 

(a)  Iron  rusting  or  copper  tarnishing. 

(b)  Wax  burning  or  paper  burning. 

(c)  Evaporation  of  gasoline  or  evaporation  of 
water. 


3.  Describe  two  homogeneous  and  two  heteroge- 
neous systems  that  are  not  described  in  the  text. 

4.  Explain  why  there  is  danger  of  explosion  where 
a  large  amount  of  dry,  powdered,  combustible 
material  is  produced. 

5.  Explain  (at  the  molecular  level)  why  an  increase 
in  concentration  of  a  reactant  may  cause  an  in- 
crease in  rate  of  reaction. 

6.  Consider  two  gases  A  and  B  in  a  container  at 
room  temperature.  What  effect  will  the  following 
changes  have  on  the  rate  of  the  reaction  between 
these  gases? 

(a)  The  pressure  is  doubled. 

(b)  The    number    of   molecules   of  gas    A    is 
doubled. 

(c)  The  temperature  is  decreased  at  constant 
volume. 


140 


THE    RATES    OF    CHEMICAL    REACTIONS       CHAP.    8 


7.  In  an  important  industrial  process  for  producing 
ammonia  (the  Haber  Process)  the  overall  re- 
action is 

N2fgJ  +  3H2(g)  — ►-  2NH3(g)  +  24,000  calories 

A  yield  of  approximately  98  %  can  be  obtained 
at  200°C  and  1000  atmospheres  of  pressure.  The 
process  makes  use  of  a  catalyst  which  is  usually 
finely  divided,  mixed  iron  oxides  containing 
small  amounts  of  potassium  oxide,  K20,  and 
aluminum  oxide,  A1203. 

(a)  Is  this  reaction  exothermic  or  endothermic? 

(b)  Suggest  a  reason  for  the  fact  that  this  reac- 
tion is  generally  carried  out  at  a  temperature 
of  500°C  and  350  atmospheres  in  spite  of  the 
fact  that  the  yield  under  these  circumstances 
is  only  about  30%. 

(c)  What  is  the  AH  for  the  reaction  in  kilo- 
calories  per  mole  of  NH3(gJ? 

(d)  How  many  grams  of  hydrogen  must  react  to 
form  1.60  moles  of  ammonia? 

8.  Give  several  ways  by  which  the  rate  of  combus- 
tion in  a  candle  flame  might  be  increased.  State 
why  the  rate  would  be  increased. 

9.  State  three  methods  by  which  the  pressure  of  a 
gaseous  system  can  be  increased. 

10.  Do  you  expect  the  reaction 

CtlUg)  +  302(g)  — >-  2C02(g)  +  2H-20(g) 

to  represent  the  mechanism  by  which  ethylene, 
C2H4,  burns?  Why? 

11.  A  group  of  students  is  preparing  a  ten-page 
directory.  The  pages  have  been  printed  and  are 
stacked  in  ten  piles,  page  by  page.  The  pages 
must  be:  (1)  assembled  in  order,  (2)  straightened, 
and  (3)  stapled  in  sets.  If  three  students  work 
together,  each  performing  a  different  operation, 
which  might  be  the  rate-controlling  step?  What 
would  be  the  effect  on  the  overall  rate  if  the  first 
step  were  changed  by  ten  helpers  joining  the 
individual  assembling  the  sheets?  What  if  these 
ten  helpers  joined  the  student  working  on  the 
second  step?  The  third  step? 

1 2.  Describe  the  life  and  death  of  an  ordinary,  empty 
water  glass,  Utilize  the  concept  "threshold  en- 
ergy." 


13.  An  increase  in  temperature  of  10°C  rarely  dou- 
bles the  kinetic  energy  of  particles  and  hence  the 
number  of  collisions  is  not  doubled.  Yet,  this 
temperature  increase  may  be  enough  to  double 
the  rate  of  a  slow  reaction.  How  can  this  be 
explained? 

14.  In  a  collision  of  particles,  what  is  the  primary 
factor  that  determines  whether  a  reaction  will 
occur? 

15.  In  Figure  8-6,  why  is  kinetic  energy  decreasing  as 
NO>  and  CO  go  up  the  left  side  of  the  barrier 
and  why  is  kinetic  energy  increasing  as  C02  and 
NO  go  down  the  right  side?  Explain  in  terms  of 
conservation  of  energy  and  also  in  terms  of  what 
is  occurring  to  the  various  particles  in  relation 
to  each  other. 

16.  Phosphorus,  P4,  exposed  to  air  burns  spontane- 
ously to  give  P4Oi0;  the  AH  of  this  reaction  is 
—  712  kcal/mole  P4. 

(a)  Draw  an  energy  diagram  for  the  net  reaction, 
explaining  the  critical  parts  of  the  curve. 

(b)  How  much   heat   is   produced   when    12.4 
grams  of  phosphorus  burn? 

17.  Considering  that  so  little  energy  is  required  to 
convert  graphite  to  diamond  (recall  Problem  8, 
Chapter  7),  how  do  you  account  for  the  great 
difficulty  found  in  the  industrial  process  for  ac- 
complishing this? 

18.  Why  does  a  burning  match  light  a  candle? 

19.  Draw  an  energy  diagram  for  the  reaction 


C(s)  +  Q2(g) 


CO-2(g) 


(a)  when  the  C  is  in  large  chunks  of  coal. 

(b)  Is  the  curve  changed  if  very  fine  carbon 
powder  is  used? 

20.  Sketch  a  potential  energy  diagram  which  might 
represent  an  endothermic  reaction.  (Label  parts 
of  curve  representing  activated  complex,  activa- 
tion energy,  net  energy  absorbed.) 

21.  Why  is  it  difficult  to  "hardboil"  an  egg  at  the 
top  of  Pike's  Peak?  Is  it  also  difficult  to  cook 
scrambled  eggs  there?  Explain. 

22.  Give  two  factors  that  would  increase  the  rate 
of  a  reaction  and  explain  why  these  do  increase 
the  rate. 


HENRY    EYRING,    1901    - 

Henry  Eyring  is  one  of  the  most  active  and  honored  chem- 
ists of  our  time.  His  advancement  of  the  theory  of  reaction 
rates  benefits  practically  every  field  of  chemistry  and  chemi- 
cal technology.  His  300  published  papers  and  five  books 
range  through  chemistry,  physics,  metallurgy,  and  biology. 

Henry  was  born  in  Chihuahua,  Mexico,  on  a  cattle 
ranch.  The  Eyring  family  was  forced  to  abandon  this  home 
eleven  years  later  under  threat  by  the  revolutionist  Salazar. 
With  other  displaced  American  colonists,  the  family  moved 
to  Texas,  then  to  Arizona.  By  distinguishing  himself  in 
high  school,  Henry  Eyring  earned  a  scholarship  to  the 
University  of  Arizona.  Years  later,  this  university  gave 
him  their  Distinguished  Alumnus  Award,  proud  that  he  had 
earned  a  B.S.  in  mining  engineering  and  an  M.S.  in 
metallurgy. 

After  graduation  he  was  highly  successful  but  not  satis- 
fied as  an  engineer  in  a  flotation  mill.  He  turned  to  the 
University  of  California  where  he  received  the  Ph.D.  in 
physical  chemistry  and  where  he  felt  the  inspiration  of  the 
great  G.  N.  Lewis.  Two  years  of  teaching  at  the  University 
of  Wisconsin,  a  year  of  study  in  Germany,  and  a  year  as 
lecturer  at  the  University  of  California  won  him  a  faculty 
position  at  Princeton  University.  In  1946  he  went  to  the 
University  of  Utah  as  Chairman  of  the  Chemistry  Depart- 
ment and  Dean  of  the  Graduate  School,  carrying  with  him 
world  recognition  in  chemistry. 

Henry  Eyring'' s  research  has  been  original  and  frequently 
unorthodox.  He  was  one  of  the  first  chemists  to  apply 
quantum  mechanics  in  chemistry.  He  unleashed  a  revolu- 
tion in  the  treatment  of  reaction  rates  by  use  of  detailed 
thermodynamic  reasoning.  Having  formulated  the  idea  of 
the  activated  complex,  Eyring  proc  ceded  to  find  a  myriad 
of  fruitful  applications — to  viscous  flow  of  liquids,  to  dif- 
fusion in  liquids,  to  conductance,  to  adsorption,  to  catalysis. 

Anyone  who  hears  Eyring  speak  on  chemistry  leaves 
convinced  that  clarifying  chemical  behavior  is  exhilarating 
fun.  Ey ring's  enjoyment  of  science  is  as  obvious  as  is  the 
importance  of  his  fundamental  contributions.  There  remains 
the  noteworthy  facet  of  this  great  scientist  that  he  is  deeply 
religious  and  gives  generously  of  his  time  and  energy  to  his 
church.  He  shows  deep  concern  about  the  political,  social, 
and  ethical  implications  of  science  and  is  always  ready  to 
discuss  them.  Thus  it  can  be  said  that,  in  the  broadest 
sense,  Henry  Eyring  acts  as  a  catalyst  of  men's  minds. 


CHAPTER 


9 

Equilibrium  in 
Chemical  Reactions 


•  •  •  by  •  •  •  equilibrium,  we  mean  a  state  in  which  the  properties  of 
a  system,  as  experimentally  measured,  would  suffer  no  further  observ- 
able change  even  after  the  lapse  of  an  indefinite  period  of  time.  It  is  not 
intimated  that  the  individual  particles  are  unchanging. 

G.  n.  lewis  and  m.  randall,  1923 


In  Chapter  8  we  discussed  the  rate  of  the  reaction 
between  CO  and  N02, 

CO(g)  +  NCVgj  — >-  C02(g)  +  HO(g)     (7) 

Then,  later  in  that  chapter,  we  turned  to  the 
question  of  the  rate  of  the  reaction  that  is  the 
reverse  of  (7), 

C02(g)  +  NO(g)  — »-  CO(g)  +  N02(g)     (2) 

Indeed!  Can't  this  reaction  make  up  its  mind? 
If  we  mix  CO(g)  and  N02(g),  reaction  (7)  be- 
gins. But  as  soon  as  it  does,  C02(g)  and  NO(g) 
are  formed.  As  these  products  accumulate,  re- 
action (2)  becomes  possible,  undoing  reaction 


(7).  Which  wins  out? 

By  direct  observation  of  the  reddish-brown 
color  of  N02  we  can  see  the  progress  of  reaction 
(7).  The  N02  is  consumed  at  first,  but  after  a 
time  the  color  stops  changing.  When  changes 
no  longer  occur  in  a  reacting  chemical  system, 
we  say  the  system  has  reached  a  state  of  equi- 
librium. The  equilibrium  situation  raises  many 
interesting  questions.  How  do  we  recognize  equi- 
librium? What  are  the  molecules  doing  at  the 
state  of  equilibrium?  What  factors  change  the 
state  of  equilibrium?  What  is  the  composition  of 
the  gas  mixture  at  equilibrium?  In  this  chapter 
we  shall  seek  answers  to  these  questions. 


9-1     QUALITATIVE  ASPECTS  OF  EQUILIBRIUM 


We  have  encountered  equilibrium  before — in  our 
consideration  of  phase  changes.  In  Section  5-1.2 
142 


we  considered  the  liquid-gas  equilibrium  that 
fixes  the  vapor  pressure  of  a  liquid,  and  in  Sec- 


SEC.    9-1    I   QUALITATIVE    ASPECTS   OF   EQUILIBRIUM 


143 


Solid  begins 
to  dissolve 


Solid  stilt  dissolving, 
color  deepening 


JVb  -morQ  changes  , 
equilibrium  exists 


Fig. 


9-1.  Iodine  dissolving  in  an  alcohol-water  mix- 
ture. Equilibrium  is  recognized  by  constant 
color  of  the  solution. 


tion  5-2.4  we  considered  the  solid-liquid  equilib- 
rium that  fixes  the  solubility  of  a  solid  in  a  liquid. 
With  this  as  background,  let  us  consider  the  first 
question  about  equilibrium:  How  do  we  recog- 
nize it? 


9-1.1    Recognizing  Equilibrium 

Figure  9-1  shows  the  addition  of  solid  iodine  to 
a  mixture  of  water  and  alcohol.  At  first  the 
liquid  is  colorless  but  very  quickly  a  reddish 
color  appears  near  the  solid.  Stirring  the  liquid 
causes  swirls  of  the  reddish  color  to  move  out — 
solid  iodine  is  dissolving  to  become  part  of  the 
liquid.  Changes  are  evident:  the  liquid  takes  on 
an  increasing  color  and  the  pieces  of  solid  iodine 
diminish  in  size  as  time  passes.  Finally,  however, 
the  color  stops  changing  (see  Figure  9-1).  Solid 
is  still  present  but  the  pieces  of  iodine  no  longer 
diminish  in  size.  Since  we  can  detect  no  more 
evidence  of  change,  we  say  that  the  system  is  at 
equilibrium.  Equilibrium  is  characterized  by  con- 
stancy of  macroscopic  properties* 

Calcium  carbonate,  CaC03,  decomposes  upon 


*  Remember  that  the  word  macroscopic  was  defined  in 
Chapter  7.  It  means  a  large  amount  of  material — enough 
to  see  and  weigh. 


heating  to  form  carbon  dioxide  gas,  C02,  and 
calcium  oxide  (lime),  CaO:f 

CaCOi(s)  ^S-  CaO(s)  +  C02(g)  (5) 

temp. 

Figure  9-2  shows  the  result  of  heating  solid 
CaC03,  initially  under  a  vacuum,  to  800°C  (part 
A).  Decomposition  begins  according  to  reaction 
(3)  and  the  gas  pressure  rises  (part  B).  The  pres- 
sure continues  to  rise  until  it  reaches  190  mm 
(part  C).  Thereafter,  no  further  change  is  evident. 
Since  we  can  detect  no  more  evidence  of  change, 
we  say  that  the  system  is  at  equilibrium.  Equi- 
librium is  characterized  by  constancy  of  macro- 
scopic properties. 

Though  a  system  at  equilibrium  is  constant  in 
properties,  constancy  is  not  the  only  require- 
ment. Consider  a  laboratory  burner  flame.  There 
is  a  well-defined  structure  to  the  flame — an  inner 
cone  surrounded  by  a  luminous  region  whose 
appearance  does  not  change.  A  temperature 
measurement  made  at  a  particular  place  in  the 
flame  shows  that  the  temperature  at  that  spot 
is  constant.  At  another  place  in  the  flame  the 
temperature  might  be  different  but,  again,  it 
would  be  constant,  not  changing  with  time.  A 
measurement  of  the  gas  flow  rate  shows  a  con- 
stant movement  of  gas  into  the  flame.  Yet  a 
laboratory  burner  flame  is  not  at  equilibrium  be- 


f  Reaction  (3)  is  used  for  the  manufacture  of  millions 
of  tons  of  lime  every  year  in  the  United  States,  for  use, 
principally,  in  plaster. 


144 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    I    CHAP.    9 


CaCO,  (s) 


Vacuum 


Ca  COj  (S)  -h  CaO  (S) 


Solid  firstr  lieatad. 
PresPitra  ~  O 


JP-K&ssura    t-i&incj. 

Calciurn   ox.ide  and 

carhon.    dioxide  forming. 


Pressure.    —190  -vnm. 
2>/o    -more    chances. 
Equilibrium    exis-ts. 


Fig.  9-2.  The  thermal  decomposition  of  calcium  car- 
bonate, CaCO:,(.y)  *±  CaO(i)  +  CO  An). 


cause  chemical  change  is  occurring.  Methane, 
CH4,  and  oxygen,  Oj,  are  continuously  fed  into 
the  flame  and  carbon  dioxide,  C02,  and  water, 
H20,  are  continuously  leaving.  Substances  are 
entering  and  leaving  at  all  times.  Such  a  system 
is  called  an  open  system.  Furthermore,  the  tem- 
perature is  not  uniform  throughout  the  system. 
Equilibrium  can  exist  only  in  a  closed  system — 
a  system  containing  a  constant  amount  of  matter 
with  all  of  this  matter  at  the  same  temperature. 
The  laboratory  burner  flame  is  called  a  steady 
state  to  indicate  that  some  of  its  properties  are 
constant  but  equilibrium  does  not  exist. 

Now  we  can  give  a  complete  statement  about 
recognizing  equilibrium:  equilibrium  is  rec- 
ognized by  the  constancy  of  macroscopic 
properties  in  a  closed  system  at  a  uniform 
temperature. 

EXERCISE  9-1 

Which  of  the  following  systems  constitute  steady 
state  situations,  and  which  are  at  equilibrium? 
For  each,  a  constant  property  is  indicated. 

(a)  An  open  pan  of  water  is  boiling  on  a  stove. 
The  temperature  of  the  water  is  constant. 

(b)  A  balloon  contains  air  and  a  few  drops  of 
water.  The  pressure  in  the  balloon  is  con- 
stant. 

(c)  An  ant-hill  follows  its  daily  life.  The  popula- 
tion of  the  ant-hill  is  constant. 


9-1.2    The  Dynamic  Nature  off  Equilibrium 

The  constancy  of  properties  at  equilibrium  refers 
to  macroscopic  measurements.  Now  we  will  con- 
sider what  the  equilibrium  is  like  on  the  molec- 
ular level,  as  chemists  picture  it. 

SOLUBILITY 

Figure  9-1C  shows  a  system  at  equilibrium. 
Solid  iodine  has  dissolved  in  an  alcohol-water 
mixture  until  the  solution  is  saturated.  Then  no 
more  solid  dissolves  and  the  color  of  the  solution 
remains  constant. 

Among  the  molecules,  however,  business  is 
going  on  as  usual.  Iodine  dissolves  by  the  detach- 
ment of  surface  layer  molecules  from  the  iodine 
crystals.  The  rate  at  which  this  process  occurs  is 
fixed  by  the  stability  of  the  crystal  (tending  to 
hold  the  molecules  in  the  surface  layer)  and  the 
temperature  (the  thermal  agitation  tending  to 
dislodge  the  molecules  from  their  lattice  posi- 
tions). As  the  dissolving  continues,  the  concen- 
tration of  iodine  molecules  in  the  solution  in- 
creases. 

Occasionally  a  molecule  moving  about  in  the 
solution  encounters  the  surface  of  an  iodine 
crystal  and  lodges  there.  This  addition  to  the 
crystal  is  called  precipitation,  or  crystallization, 
and  it  occurs  more  and  more  often  as  the  con- 
centration of  iodine  in  solution  rises. 

Here  we  have  two  opposing  processes.  At  a 
given  temperature,  molecules  leave  the  surface 
of  the  crystal  at  a  constant  rate,  tending  to  in- 
crease the  concentration  in  solution.  On  the 
other  hand,  dissolved  molecules  are  continually 
striking  the  surface  and  precipitating,  tending  to 


SEC.    9-1     I    QUALITATIVE    ASPECTS    OF    EQUILIBRIUM 


145 


decrease  the  concentration  of  molecules  in  solu- 
tion. When  enough  material  has  dissolved  so  that 
the  rate  of  return  of  molecules  to  the  surface  of 
the  solid  is  just  equal  to  the  rate  at  which  they 
are  leaving  the  surface,  no  more  net  change  will 
occur.  Even  though  molecules  are  continually 
dissolving  and  others  are  precipitating,  as  long 
as  these  two  processes  are  in  balance,  the  amount 
of  iodine  dissolved  per  unit  volume  of  solution 
will  be  constant.  This  macroscopic  property,  the 
solubility,  is  now  constant:  the  system  is  in  solu- 
bility equilibrium.  But  chemists  interpret  this 
constancy  as  a  balance  between  two  oppos- 
ing processes  which  continue  at  equilibrium. 
At  equilibrium,  microscopic  processes  con- 
tinue but  in  a  balance  that  yields  no  macro- 
scopic changes. 

VAPOR    PRESSURE 

Consideration  of  the  dissolving  of  iodine  in  an 
alcohol-water  mixture  on  the  molecular  level 
reveals  the  dynamic  nature  of  the  equilibrium 
state.  The  same  type  of  argument  is  applicable 
to  vapor  pressure. 

We  have  already  noted  that  if  we  place  liquid 
water  in  a  flask  at  20°C  and  seal  the  flask,  some 
water  molecules  leave  the  liquid  and  enter  the 
gas  phase.  The  partial  pressure  rises  as  more  and 
more  water  molecules  become  part  of  the  gas. 
Finally,  however,  the  pressure  stops  rising  and 
the  partial  pressure  of  water  becomes  constant. 
This  partial  pressure  is  the  vapor  pressure  and 
equilibrium  now  exists. 

Yet,  it  is  reasonable  to  suppose  that  water 
molecules  from  the  liquid  are  still  evaporating, 
even  at  equilibrium.  Molecules  in  the  liquid  have 
no  way  of  "knowing"  that  the  partial  pressure 
of  the  vapor  is  equal  to  the  vapor  pressure.  In 
the  gas  phase,  the  randomly  moving  molecules 
continue  to  strike  the  surface  of  the  liquid  and 
condense.  Equilibrium  corresponds  to  a  perfect 
balance  between  this  continuing  evaporation  and 
condensation.  Then  no  net  changes  can  be  de- 
tected.* 

*  When  the  partial  pressure  of  the  water  equals  the 
vapor  pressure,  the  gas  above  the  liquid  is  said  to  be 
saturated.  The  word  "saturated"  has  the  same  meaning 
as  it  did  relative  to  solubility:  the  gas  phase  contains  as 
much  water  vapor  as  it  can  hold  at  equilibrium. 


»1 

OB    nnw 


A 

B 

C 

Unsaturated 

Saturated 

Supers  a  tu  rated 

vapor" 

vapor, 
EQUILIBRIUM 

Vapor 

Fig.  9-3.  Exchange  of  molecules  between  liquid  and 

gas  (A)  when  the  partial  pressure  is  below 
the  vapor  pressure;  (B)  at  equilibrium;  (C) 
when  the  partial  pressure  is  above  the  vapor 
pressure. 


Figure  9-3  shows  this  schematically.  If  the 
partial  pressure  of  the  vapor  is  less  than  the 
equilibrium  value  (as  in  Figure  9-3A),  the  rate  of 
evaporation  exceeds  the  rate  of  condensation 
until  the  partial  pressure  of  the  vapor  equals  the 
equilibrium  vapor  pressure.  If  we  inject  an  excess 
of  vapor  into  the  bottle  (as  in  Figure  9-3C),  con- 
densation will  proceed  faster  than  evaporation 
until  the  excess  of  vapor  has  condensed.  The 
equilibrium  vapor  pressure  corresponds  to  that 
concentration  of  water  vapor  at  which  condensa- 
tion and  evaporation  occur  at  exactly  the  same 
rate  (as  in  Figure  9-3 B).  At  equilibrium,  micro- 
scopic processes  continue  but  in  a  balance  that 
yields  no  macroscopic  changes. 

CHEMICAL    REACTIONS 

Let  us  examine  a  chemical  reaction  to  see  if  these 
same  conditions  apply.  Suppose  we  fill  two  iden- 
tical bulbs  to  equal  pressures  of  nitrogen  dioxide. 
Now  immerse  the  first  bulb  (bulb  A)  in  an  ice 
bath  and  the  second  bulb  (bulb  B)  in  boiling 
water,  as  in  Figure  9-4.  The  gas  in  bulb  A  at 
0°C  is  almost  colorless;  the  gas  in  bulb  B  at 
100°C  is  reddish-brown.  The  predominant  mo- 
lecular species  in  the  cold  bulb  must  be  different 
from  that  in  the  hot  bulb.  A  variety  of  experi- 
ments shows  that  the  cold  bulb  contains  mostly 
N2O4  molecules.  These  same  experiments  show 
that  the  hot  bulb  contains  mostly  N02  molecules. 
The  N2O4  molecules  absorb  no  visible  light,  so 


EQUILIB 


RIUM    IN    CHEMICAL    REACTIONS   |    CHAP.    9 


r  -  0°C 


t^2S°C 


t=100°C 


Fig  9-4.  Nitrogen  dioxide  gas  at  different  tempera- 
tures. Bulb  A:  At  0°C:  N20,  {almost  color- 
less). Bulb  B:  At  100°C:  N02  (reddish- 
brown). 


the  cold  gas  is  almost  colorless.  The  N02  mole- 
cules do  absorb  some  visible  light,  so  the  hot  gas 
is  reddish-brown. 

Now  let  us  transfer  these  two  bulbs  to  a  bath 
at  room  temperature,  as  shown  in  Figure  9-5. 
Immediately  the  color  begins  to  intensify  in  bulb 
A.  The  color  shows  that  a  chemical  change  has 
occurred,  forming  N02  molecules  from  N204: 
In  bulb  A        N204(gj  — ►-  2N02(g,>        00 

At  the  same  time,  the  color  in  bulb  B  begins  to 
pale,  showing  that  a  chemical  change  has  oc- 
curred in  this  bulb  as  well,  forming  N204  mole- 
cules from  N02: 

In  bulb  B       2N02(gj  — *■  WJg)        (5) 

In  each  bulb  the  colors  continue  to  change,  bulb 
A  becoming  darker  and  bulb  B  becoming  more 


Fig.  9-5.  Nitrogen  dioxide  gas  at  room  temperature. 

Bulb  A  and  bulb  B  after  transfer  to  water 
bath  at  25°C. 


pale.  Finally,  as  the  two  bulbs  approach  the  same 
temperature,  the  colors  stop  changing.  A  close 
examination  shows  that  the  two  colors  are  now 

the  same! 

By  direct  visual  observation  we  can  watch  the 
contents  of  these  two  bulbs  approach  the  con- 
stancy of  macroscopic  properties  (in  this  case, 
color)  that  indicates  equilibrium.  In  bulb  A  equi- 
librium was  approached  by  the  dissociation  of 
N204,  reaction  (4);  in  bulb  B  it  was  approached 
by  the  opposite  reaction,  reaction  (5).  Here  it  is 
clear  why  the  color  of  each  bulb  stopped  chang- 
ing at  the  particular  hue  characteristic  of  the 
equilibrium  state  at  25°C.  The  reaction  between 
N02  and  N204  can  proceed  in  both  directions: 

N204(gj  — *■  2N02(gj  00 

N204(gj  ■<—  2N02(g>)  (5) 

Since  N204  molecules  can  dissociate  in  bulb  A, 
they  must  also  be  able  to  dissociate  in  bulb  B. 
Surely  an  N204  molecule  doesn't  act  differently 
in  bulb  A  (at  25°C)  than  it  does  in  bulb  B  (at 
25°C).  The  same  sort  of  statements  must  apply 
to  the  combination  of  two  N02  molecules.  If  the 
reaction  occurs  in  bulb  B,  then  it  must  also  occur 
in  bulb  A.  The  net  change  we  see  (by  observing 
the  changing  N02  color)  represents,  then,  the 
difference  in  the  rate  of  production  of  N02  by 
reaction  (4)  and  the  rate  of  loss  of  N02  by  reac- 
tion (5).  Changes  will  cease  when  these  two  rates 
are  exactly  equal.  If  we  approach  equilibrium 
from  a  lower  temperature  (which  favors  N204), 
then  reaction  (4)  predominates  at  first.  But  as 
more  and  more  NOz  is  produced,  reaction  (5) 
becomes  faster  and  faster.  When  reaction  (5)  be- 
comes just  as  fast  as  reaction  (4),  then  equilib- 
rium has  been  reached:  macroscopic  properties 
no  longer  change  even  though  both  reactions  still 
proceed  in  a  state  of  balance.  At  this  time  we 
replace  the  single  arrow  in  reaction  (4)  by  a 
double  arrow  (+±)  or  an  equals  sign  (=)  to 
show  that  equilibrium  prevails:* 

NtOt(g)  q=fc  2NOt(g)  (*) 

NjCVsJ    =    THOjg)  «*) 


*=25°C 


♦These  alternative  notations,  +±.  and  =,  are  used 
by  chemists  interchangeably  in  equations  for  chemical 
reactions.  Both  notations  will  be  seen. 


SEC.    9-1    I    QUALITATIVE    ASPECTS    OF    EQUILIBRIUM 


147 


In  bulb  B  we  approached  equilibrium  from  a 
higher  temperature  (which  favors  N02);  then  re- 
action (5)  predominated  at  first.  Using  the  same 
sort  of  argument  we  applied  to  bulb  A,  we  see 
that  as  time  progresses,  reaction  (4)  becomes 
more  and  more  rapid  (as  N204  is  produced)  and 
reaction  (5)  becomes  slower  (as  N02  is  used  up). 
Finally,  when  the  rates  become  equal,  equilib- 
rium is  reached  and  the  equilibrium  expression 
(6)  is  applicable  in  bulb  B. 

For  chemical  reactions,  just  as  for  phase 
changes,  at  equilibrium,  microscopic  processes 
continue  but  in  a  balance  which  gives  no  macro- 
scopic changes. 

9-1.3    The  State  of  Equilibrium 

It  is  most  important  to  note  that  in  our  descrip- 
tion of  the  equilibrium  state  we  have  not  implied 
that  at  equilibrium  the  number  of  moles  of  N204 
remaining  is  the  same  as  the  number  of  moles  of 
N02  produced.  Equation  (<5)  gives  us  no  informa- 
tion concerning  the  fraction  of  the  nitrogen  di- 
oxide present  as  NOi  at  equilibrium.  This  is 
easily  verified  by  heating  the  water  surrounding 
bulbs  A  and  B  about  10°.  The  colors  of  the  gases 
in  both  bulbs  change  to  a  new  equilibrium  color 
(one  corresponding  to  the  presence  of  more 
N02).  Yet  the  same  expression  is  applicable: 
N204(gj  =«=*:  2HOJg)  (6) 

What  does  equation  (<5)  tell  us,  then?  First,  it 
tells  us  that  equilibrium  prevails  (ihe  q=^  sign 
tells  us  that).  Next,  it  tells  us  that  there  are  two 
types  of  molecules  present,  N204  molecules  and 
N02  molecules.  Finally,  it  tells  us  that  during  the 
approach  to  equilibrium,  two  molecules  of  N02 
are  produced  (or  consumed)  for  every  one  mole- 
cule of  N204  dissociated  (or  formed).  It  does  not 
tell  us  whether  at  equilibrium  there  will  be  much 
or  little  N02  compared  with  the  amount  of  N204. 
To  emphasize  this  point,  let  us  consider  an- 
other familiar  reaction, 

HMg)  qj=fc  H2(g)  +  \02(g)  (7) 

Until  we  are  given  the  necessary  information  we 
have  no  idea  how  complete  the  decomposition  of 
water  is  at  equilibrium.  All  we  know  is  that  for 
every  mole  of  water  which  decomposes  we  will 


obtain  1  mole  of  hydrogen  and  \  mole  of  oxygen. 

It  has  been  determined  that  in  a  closed  vessel 
at  2273°K  and  a  total  pressure  equal  to  one 
atmosphere  at  the  time  equilibrium  is  attained, 
0.6%  of  the  water  has  dissociated.  If  we  started 
with  one  mole  of  water,  0.6%  =  0.6  X  ^hs  = 
0.006  mole  would  decompose.  There  would  be 
left  1  -  0.006  =  0.994  mole  of  undecomposed 
water.  There  would  be  formed  0.006  mole  of  H2 
and  0.003  mole  of  oxygen.  We  can  summarize 
this  as  follows: 

H&g )  q=b  H2(g)  +  \02(g) 

Initial  moles  1  0  0 


Moles  present 
at  equilibrium 


0.994 


0.006        0.003 


In  other  words,  if  we  start  with  water,  not  much 
of  it  has  decomposed  when  the  equilibrium  state 
is  attained  at  2273°K. 

Now  what  about  approaching  the  equilibrium 
state  by  starting  with  hydrogen  and  oxygen?  Let 
us  start  with  1  mole  of  hydrogen  and  \  mole  of 
oxygen  and  allow  the  reaction  to  attain  equilib- 
rium at  2273°K  and  a  total  pressure  equal  to  one 
atmosphere.  At  equilibrium  we  find  present  0.994 
mole  of  water,  0.006  mole  of  Hj,  and  0.003  mole 
of  02.  This  can  be  summarized  as  follows: 


Initial  moles 

Moles  present 
at  equilibrium 


Wg)  +  \0«(g) 
1  0.5 

0.006        0.003 


U£>(g) 
0 

0.994 


If  we  start  with  hydrogen  and  oxygen,  equilib- 
rium is  attained  after  most  of  the  hydrogen  and 
oxygen  have  united  to  form  water.  More  im- 
portant, though,  the  partial  pressures  at  equilib- 
rium are  the  same  as  those  obtained  beginning 
with  pure  H20.  The  equilibrium  pressures  are 
fixed  by  the  temperature,  the  composition,  and 
the  total  pressure;  they  do  not  depend  upon  the 
direction  from  which  equilibrium  is  approached. 
The  balanced  equation  does  not  indicate  the 
concentrations  (or  partial  pressures)  at  equilib- 
rium. 

9-1.4    Altering  the  State  of  Equilibrium 

We  have  seen  that,  qualitatively,  the  state  of 
equilibrium  for  a  system  is  characterized  by  the 


148 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    I    CHAP.    9 


relative  amounts  of  products  and  reactants  pres- 
ent. With  reference  to  the  decomposition  of 
water,  any  change  in  conditions  which  would 
cause  more  than  0.6%  of  the  water  to  dissociate 
at  equilibrium  would  be  said  to  change  the  state 
of  equilibrium  for  the  reaction, 


H,0(g)  +±  U2(g)  +  \Q2(g) 


(7) 


in  favor  of  the  formation  of  more  hydrogen  and 
oxygen. 

What  conditions  might  alter  the  equilibrium 
state?  Concentration  and  temperature!  These 
are  factors  that  affect  the  rate  of  reaction.  Equi- 
librium is  attained  when  the  rates  of  opposing 
reactions  become  equal.  Any  condition  that 
changes  the  rate  of  one  of  the  reactions  involved 
in  the  equilibrium  may  affect  the  conditions  at 
equilibrium. 

CONCENTRATION 

Consider  the  reaction  you  encountered  in  the 
laboratory — that  between  ferric  ion  (Fe+3)  and 
thiocyanate  ion  (SCN-): 


Fe+YaqJ  +  SCN-(aq) 


FeSCN+2(aqj    (5) 


Again  we  have  visual  evidence  of  concentration 
at  equilibrium  since  the  intensity  of  the  color  is 
fixed  by  the  concentration  of  the  FeSCN+2  ion. 
The  addition  of  either  more  ferric  ion  [by  adding 


Fig.  9-6.  Equilibrium  conditions  are  affected  by  the 
reactant  concentrations. 


a  soluble  salt  such  as  ferric  nitrate,  Fe(N03)3]  or 
more  thiocyanate  ion  (by  adding,  say,  sodium 
thiocyanate)  changes  the  concentration  of  one  of 
the  reactants  in  (8).  Immediately  the  color  of 
the  solution  darkens,  showing  that  there  is  an 
increase  in  the  amount  of  the  colored  ion, 
FeSCN+2.  The  equilibrium  concentrations  are  af- 
fected if  the  concentrations  of  reactants  {or  prod- 
ucts) are  altered. 

TEMPERATURE 

We  have  already  considered  an  example  of  the 
change  of  equilibrium  concentrations  as  the  tem- 
perature is  altered.  The  relative  amounts  of  N02 
and  N204  are  readily  and  obviously  affected  by 
a  temperature  change.  The  equilibrium  concen- 
trations are  affected  if  the  temperature  is  altered. 

CATALYSTS 

Catalysts  increase  the  rate  of  reactions.  It  is 
found  experimentally  that  addition  of  a  catalyst 
to  a  system  at  equilibrium  does  not  alter  the 
equilibrium  state.  Hence  it  must  be  true  that  any 
catalyst  has  the  same  effect  on  the  rates  of  the 
forward  and  reverse  reactions.  You  will  recall 
that  the  effect  of  a  catalyst  on  reaction  rates  can 
be  discussed  in  terms  of  lowering  the  activation 
energy.  This  lowering  is  effective  in  increasing 
the  rate  in  both  directions,  forward  and  reverse. 
Thus,  a  catalyst  produces  no  net  change  in  the 
equilibrium  concentrations  even  though  the  sys- 
tem may  reach  equilibrium  much  more  rapidly 
than  it  did  without  the  catalyst. 

9-1.5    Attainment  of  Equilibrium 

The  equilibrium  state  is  not  always  attained  in 
chemical  reactions.  Consider  reaction  (7): 

H20(g)  q=fc  H2(g)  +  \02(g) 

AH  =  +57.8  kcal    (7) 

A  large  amount  of  heat  is  absorbed  in  this  reac- 
tion, 57.8  kcal/mole  of  water  decomposed.  If  the 
temperature  is  lowered,  the  state  of  equilibrium 
is  even  more  favorable  to  the  production  of 
water  at  room  temperature  than  it  is  at  2273°K. 
Yet  a  mixture  of  hydrogen  and  oxygen  can  re- 
main at  room  temperature  for  a  long  period 
without  apparent  reaction.  Equilibrium  is  not 


SEC.    9-1     I    QUALITATIVE    ASPECTS    OF    EQUILIBRIUM 


149 


attained  in  this  system  because  the  rate  of  the 
reaction  between  hydrogen  and  oxygen  at  room 
temperature  is  too  low.  This  explanation  is  easily 
verified  by  speeding  up  the  reaction  slightly.  If  a 
mixture  of  H2  and  02  is  disturbed  with  a  small 
spark,  reaction  begins  and  it  enthusiastically  (and 
explosively)  continues  until  most  of  the  gases 
have  been  converted  to  water. 

This  distinction  between  the  conditions  in  a 
chemical  system  at  equilibrium  and  the  rate  at 
which  these  conditions  are  attained  is  very  im- 
portant in  chemistry.  By  arguments  that  we  shall 
consider  a  chemist  can  decide  with  confidence 
whether  equilibrium  favors  reactants  or  products 
or  neither.  He  cannot  predict,  however,  how 
rapidly  the  system  will  approach  the  equilibrium 
conditions.  That  is  a  matter  of  reaction  rates, 
and  the  chemist  must  perform  separate  experi- 
ments to  learn  whether  a  given  rate  is  rapid  or 
not. 

9-1.6    Predicting  New  Equilibrium 

Concentrations:  Le  C  ha  teller's  Principle 

We  are  not  satisfied  with  the  conclusion  that  this 
change  or  that  change  affects  the  equilibrium 
concentrations.  We  would  also  like  to  predict 
the  direction  of  the  effect  (does  it  favor  products 
or  reactants?)  and  the  magnitude  of  the  effect 
(how  much  does  it  favor  products  or  reactants?). 
The  first  desire,  to  know  the  qualitative  effects, 
is  answered  by  a  generalization  first  proposed  by 
a  French  chemist,  Henry  Louis  Le  Chatelier,  and 
now  called  Le  Chatelier's  Principle. 

Le  Chatelier  sought  regularities  among  a  large 
amount  of  experimental  data  concerning  equi- 
libria. To  summarize  the  regularities  he  found, 
he  made  this  generalization:  If  an  equilibrium 
system  is  subjected  to  a  change,  processes 
occur  that  tend  to  counteract  partially  the 
imposed  change.  This  generalization  has  been 
found  to  be  applicable  to  such  a  large  number 
of  systems  that  it  is  now  called  a  principle.  Let 
us  see  how  it  applies  to  our  examples. 

CONCENTRATION    AND    LE    CHATELIER'S 
PRINCIPLE 

If  a  soluble  thiocyanate  salt  is  added  to  an  equi- 
librium solution  containing  both  ¥e+3(aq)  and 


SCN-(aq),  the  color  of  the  complex  ion  in- 
creases : 

Fe+3(aq)  +  SCN-faqj  3=fc  FeSCN+2( aq)    (8) 

A  new  state  of  equilibrium  is  then  attained  in 
which  more  FeSCN+2  is  present  than  was  there 
before  the  addition  of  SCN-.  Increasing  the  con- 
centration of  SCN-  has  increased  the  concentra- 
tion of  the  FeSCN+2  ion.  This  is  in  accord  with 
Le  Chatelier's  Principle.  The  change  imposed 
on  the  system  was  an  increase  in  the  concentra- 
tion of  SCN-.  This  change  can  be  counteracted 
in  part  by  some  Fe+3  and  SCN-  ions  reacting  to 
form  more  FeSCN+2.  The  same  argument  applies 
to  an  addition  of  ferric  ion  from  a  soluble  ferric 
salt.  In  each  case,  the  formation  of  FeSCN+2  uses 
up  a  portion  of  the  added  reactant,  partially 
counteracting  the  change. 

PRESSURE    AND    LE    CHATELIER'S 
PRINCIPLE 

Instead  of  altering  the  concentration  of  one  in- 
dividual component  in  an  equilibrium  system, 
we  can  alter  the  concentration  of  all  gaseous 
components  by  changing  the  pressure  at  which 
the  system  is  confined.  Let  us  start  with  the  sys- 
tem represented  by  equation  (7)  and  double  the 
total  pressure.  The  system  now  occupies  a  much 
smaller  volume  than  it  did  previously.  The  total 
number  of  moles  present  per  unit  volume  is 
greater  than  it  was  under  the  original  equilibrium 
conditions.  This  change  can  be  counteracted  in 
part  if  some  hydrogen  and  oxygen  combine  to 
form  gaseous  water.  Then  the  total  number  of 
moles  present  is  reduced  (1$  moles  unite  to  form 
1  mole).  Hence,  we  can  predict  that  increasing 
the  concentration  of  all  components  by  increas- 
ing the  pressure  will  shift  the  state  of  equilibrium 
in  favor  of  the  formation  of  gaseous  water.  This 
is  in  accord  with  experiment. 

A  change  in  total  pressure  does  not  always 
shift  equilibrium.  The  first  reaction  mentioned 
in  this  chapter  exemplifies  this: 

CO(g)  +  NCVgj  +±  CO,(g)  +  NO(g)     (9) 

If  we  increase  the  pressure  on  a  mixture  of  these 
four  gases  at  equilibrium,  the  gases  are  com- 
pressed to  a  smaller  volume.  Once  again  the  con- 


150 


EQUILIBRIUM   IN    CHEMICAL   REACTIONS   I   CHAP.    9 


centrations  are  all  increased.  What  does  Le 
Chatelier's  Principle  tell  us  here?  If  the  equilib- 
rium state  is  altered  to  favor  products,  some  CO 
and  N02  molecules  react  to  form  an  exactly 
equal  number  of  molecules  of  C02  and  NO. 
Since  there  is  no  change  in  the  total  number  of 
moles,  the  proposed  change  in  the  equilibrium 
does  not  partially  reduce  the  pressure  change. 
Le  Chatelier's  Principle  tells  us  that  processes 
occur  so  as  to  "counteract  partially  the  imposed 
change."  Here  neither  a  change  favoring  the 
reactants  nor  a  change  favoring  products  will 
"counteract"  the  imposed  pressure  change. 
Hence,  Le  Chatelier's  Principle  leads  us  to  expect 
no  change  of  the  equilibrium  state  for  reaction 
(9)  when  the  pressure  is  altered.  Experimentally 
we  find  that  no  change  is  observed.  The  equilib- 
rium state  is  not  affected  by  a  pressure  change 
for  any  equilibrium  gas  mixture  where  the  num- 
ber of  reactant  molecules  is  the  same  as  the 
number  of  product  molecules  in  the  balanced 
reaction. 


EXERCISE  9-2 

Does  Le  Chatelier's  Principle  predict  a  change  of 
equilibrium  concentrations  for  the  following  re- 
actions if  the  gas  mixture  is  compressed?  If  so, 
does  the  change  favor  reactants  or  products? 

(a)  N20<fej  q=b  2N02(gj 

(b)  H2(g)  +  Ug)  +±  2Hl(g) 

(c)  N2(g)  +  3H2(g)  +±  2NH3(g) 


TEMPERATURE   AND    LE   CHATELIER'S 
PRINCIPLE 

Let  us  add  to  reaction  (4)  the  information  that 
the  decomposition  of  N204  is  endothermic: 

WJg)  +±:  2NOi(g)       AH  =  +14.1  kcal    (4) 

Our  experimental  observations  indicated  that 
warming  a  bulb  containing  N02  and  N204  caused 
a  shift  of  the  equilibrium  state  in  favor  of  the 
formation  of  N02  (the  reddish-brown  color 
deepened).  It  is  easy  to  see  that  this  is  in  accord 
with  Le  Chatelier's  Principle.  A  rise  in  tempera- 
ture is  caused  by  an  input  of  heat.  At  the  higher 


temperature,  the  equilibrium  is  changed  to  form 
more  N02.  The  formation  of  N02  absorbs  a  por- 
tion of  the  heat  that  caused  the  temperature  rise. 
Raising  the  temperature  of  liquid  water  raises 
its  vapor  pressure.  This  is  in  accord  with  Le 
Chatelier's  Principle  since  heat  is  absorbed  as  the 
liquid  vaporizes.  This  absorption  of  heat,  which 
accompanies  the  change  to  the  new  equilibrium 
conditions,  partially  counteracts  the  temperature 
rise  which  caused  the  change. 

9-1.7    Application  of  Equilibrium  Principles: 
The  Haber  Process 

Knowledge  of  chemical  principles  pays  rewards 
in  technological  progress.  Control  of  chemical 
reactions  is  the  key.  The  large  scale  commercial 
production  of  nitrogen  compounds  provides  a 
practical  example  of  the  beneficial  application  of 
Le  Chatelier's  Principle. 

The  most  difficult  step  in  the  process  for  the 
conversion  of  the  inert  nitrogen  of  the  atmos- 
phere into  important  commercial  compounds 
such  as  fertilizers  and  explosives  involves  the 
reaction 


N2(g)  +  3Hi(gj  +±z  2NH3(g)  +  22  kcal    (10a) 


or 


NJg)  +  3H2(g) 


2NH3(g) 

AH  =  -22  kcal    (Wb) 


Can  we  predict  the  optimum  conditions  for  a 
high  yield  of  NH3?  Should  the  system  be  allowed 
to  attain  equilibrium  at  a  low  or  a  high  tempera- 
ture? Application  of  Le  Chatelier's  Principle  sug- 
gests that  the  lower  the  temperature  the  more 
the  equilibrium  state  will  favor  the  production 
of  NH3.  Should  we  use  a  low  or  a  high  pressure? 
The  production  of  NH3  represents  a  decrease 
in  total  moles  present  from  4  to  2.  Again  Le 
Chatelier's  Principle  suggests  use  of  pressure  to 
increase  concentration.  But  what  about  prac- 
ticality? At  low  temperatures  reaction  rates  are 
slow.  Therefore  a  compromise  is  necessary.  Low 
temperature  is  required  for  a  desirable  equilib- 
rium state  and  high  temperature  is  necessary  for 
a  satisfactory  rate.  The  compromise  used  in- 
dustrially involves  ?"  intermediate  temperature 
around  500°C  and  even  then  the  success  of  the 


SEC.    9-2    I    QUANTITATIVE    ASPECTS    OF    EQUILIBRIUM 


151 


process  depends  upon  the  presence  of  a  suitable 
catalyst  to  achieve  a  reasonable  reaction  rate. 
With  regard  to  pressure,  another  compromise 
is  needed.  It  is  expensive  to  build  high  pressure 
equipment.  A  pressure  of  about  350  atmospheres 
is  actually  used.  Under  these  conditions,  350 
atmospheres  and  500°C,  only  about  30%  of  the 
reactants  are  converted  to  NH3.  The  NH3  is  re- 
moved from  the  mixture  by  liquefying  it  under 
conditions  at  which  N2  and  H2  remain  as  gases. 


The  unreacted  N2  and  H2  are  then  recycled  until 
the  total  percent  conversion  to  ammonia  is  very 
high. 

Prior  to  World  War  I  the  principal  sources  of 
nitrogen  compounds  were  some  nitrate  deposits 
in  Chile.  Fritz  Haber,  a  German  chemist,  suc- 
cessfully developed  the  process  we  have  just 
described,  thus  allowing  chemists  to  use  the 
almost  unlimited  supply  of  nitrogen  in  the  at- 
mosphere as  a  source  of  nitrogen  compounds. 


9-2     QUANTITATIVE  ASPECTS  OF  EQUILIBRIUM 


Le  Chatelier's  Principle  permits  the  chemist  to 
make  qualitative  predictions  about  the  equilib- 
rium state.  Despite  the  usefulness  of  such  predic- 
tions, they  represent  far  less  than  we  wish  to 
know.  It  is  a  help  to  know  that  raising  the  pres- 
sure will  favor  production  of  NH3  in  reaction 
(10a).  But  how  much  will  the  pressure  change 
favor  NH3  production?  Will  the  yield  change  by 
a  factor  of  ten  or  by  one-tenth  of  a  percent?  To 
control  a  reaction,  we  need  quantitative  informa- 
tion about  equilibrium.  Experiments  show  that 
quantitative  predictions  are  possible  and  they 
can  be  explained  in  terms  of  our  view  of  equilib- 
rium on  the  molecular  level. 


9-2.1    The  Equilibrium  Constant 

By  means  of  colorimetric  determination  in  the 
laboratory  you  measured  the  concentration  of 
FeSCN+2,  which  we  shall  designate  [FeSCN+2],* 
in  solutions  containing  ferric  and  thiocyanate 
ions,  Fe+3  and  SCN~.  The  reaction  is 

Fe+YaqJ  +  SCN-(aq)  +±  FeSCN+*faqj     (8) 

From  [FeSCN+2]  and  the  initial  values  of  [Fe+3] 
and  [SCN-]  you  calculated  the  values  of  [Fe+3] 
and  [SCN-]  at  equilibrium.  You  then  made  cal- 
culations for  various  combinations  of  these  val- 


*  Hereafter,  we  shall  regularly  use  the  square  brackets 
notation,  [  ],  to  indicate  concentration.  Thus,  we  read 
[Fe+I]  as  "ferric  ion  concentration." 


ues.  Many  experiments  just  like  these  show  that 
the  ratio 

[FeSCN+2] 


[Fe+3][SCN-] 


(//) 


comes  closest  to  being  a  fixed  value.  Note  that 
this  ratio  is  the  quotient  of  the  equilibrium  con- 
centration of  the  single  substance  produced  in 
the  reaction  divided  by  the  product  of  the  equi- 
librium concentrations  of  the  reactants. 

Colorimetric  analysis  based  on  visual  estima- 
tion is  not  very  exact.  Some  more  accurate  data 
on  the  H2,  I2,  HI  system  at  equilibrium  are 
shown  in  Table  9-1.  The  reaction  is 


2Hl(g) 


H2(g)  +  h(g) 


(12) 


The  data  have  been  expressed  in  concentrations, 
although  pressure  units  are  more  usual  for  a 
reaction  involving  gases. 


EXERCISE  9-3 

For  the  last  two  experiments  in  Table  9-1,  num- 
bers 4  and  5,  why  is  [H2]  =  [I2]?  For  experiment 
1,  what  were  the  initial  concentrations  of  H2  and 
I2  before  the  reaction  occurred  to  form  HI  ? 


Let's  work  with  these  data.  Heartened  by  your 
results  from  your  own  laboratory  data,  let  us 
compute  the  value  of  the  ratio 


[H«][l«] 
[HI] 


(13) 


152 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    |    CHAP.    9 


Table  9-1.    equilibrium  concentration  at  698. e  k  of  hydrogen, 

IODINE,    AND     HYDROGEN     IODIDE 


EXPT. 
NO. 


[H2] 
(moles/liter) 


(moles/liter) 


[HI] 

(moles/liter) 


1 

1.8313  X  10"3 

3.1292  X  10"3 

17.671  X  10"3 

2 

2.9070  X  10"3 

1.7069  X  10-3 

16.482  X  10"3 

3 

4.5647  X  10-3 

0.7378  X  10-3 

13.544  X  10-3 

* 

4 

0.4789  X  10-3 

0.4789  X  10-3 

3.531  X  10-3 

5 

1.1409  X  10"3 

1.1409  X  10-3 

8.410  X  10"» 

*  Values  above  the  line  were  obtained  by  heating  hydrogen  and  iodine 
together;  values  below  the  line,  by  heating  pure  hydrogen  iodide. 


We  obtain  the  numbers  in  Table  9-II.  In  view  of 
the  precision  of  the  data  from  which  these  ratios 
are  derived,  the  ratios  are  far  from  constant. 
Now  let  us  try  the  ratio 


rH2][i2] 

[HI]2 


(14) 


These  calculations  are  summarized  in  Table 
MIL 

The  results  are  most  encouraging  and  imply 
that  with  a  fair  degree  of  accuracy  we  can  write 

[H2HI2]  f    , 

1  rHI]2  J  =  a  constant 

=  1.835  X  10"2  at  698.6°K        (75) 
Look  at  this  ratio  in  terms  of  reaction  (12): 

2Hl(g)  ^±  H2(g)  +  h(g)  (12) 

The  ratio  (75)  is  the  product  of  the  equilibrium 
concentrations  of  the  substances  produced  in  the 
reaction,  [H2]  X  [h],  divided  by  the  square  of 


the  concentration  of  the  reacting  substance, 
[HI]2.  In  this  ratio,  the  power  to  which  we  raise 
the  concentration  of  each  substance  is  equal  to 
its  coefficient  in  reaction  (12). 

9-2.2    The  Law  of  Chemical  Equilibrium 

Let  us  summarize  what  we  have  learned.  For  the 
reaction 

Ft+'(aq)  +  SCN-(aq)  +=h  FeSCN+2(aqj    (5) 

we  found  that  the  concentrations  of  the  mole- 
cules involved  have  a  simple  relationship. 

[FeSCN+2]  ,     , 

— ■*—  =  a  constant 


(76) 

[Fe+3][SCN-] 

Then  we  considered  precise  equilibrium  data  for 
the  reaction 


2Hl(g) 


Wg)  +  h(g) 


(12) 


Table  9-11 

VALUES    OF 
OF    TABLE    9 


[HtHlt] 
[HI] 


FOR    DATA 


EXPT. 
NO. 


Table  9-111 

VALUES    OF 

l&ffi    FOR    DATA 

OF    TABLE    9 

-1 

[H.][l.l 
[HI] 

EXPT.                                   TH2][I2] 
NO.                                [HI]2 

1 

32.429  X  10-* 

2 

30.105  X  10-6 

3 

24.866  X  10"s 

4 

6.495  X  10~6 

5 

15.477  X  10-* 

1.8351  X  10"2 
1.8265  X  10-2 
1.8359  X  10"2 
1.8390  X  10"2 
1.8403  X  10-2 

Average        1.835    X  lO"2 


SEC.    9-2    |    QUANTITATIVE    ASPECTS    OF    EQUILIBRIUM 


153 


The  concentrations  of  the  molecules  appearing 
in  reaction  (12)  were  found  to  have  a  simple 
relationship, 


[H«J[I»3  = 
[HI? 


=  a  constant 


(14) 


In  each  of  our  simple  relationships,  (76)  and  (74), 
the  concentrations  of  the  products  appear  in  the 
numerator.  In  each  relationship  the  concentra- 
tions of  reactants  appear  in  the  denominator.  In 
reaction  (12),  two  molecules  of  hydrogen  iodide 
react.  This  influences  expression  (14)  because  it 
is  necessary  to  square  the  concentration  of  hy- 
drogen iodide,  [HI],  in  order  to  obtain  a  con- 
stant ratio. 

These  observations  and  many  others  like  them 
lead  to  the  generalization  known  as  the  Law  of 
Chemical  Equilibrium.  For  a  reaction 


aA  +  bB  +±:  eE  +  fF 


(17) 


when  equilibrium  exists,  there  will  be  a  simple 
relation  between  the  concentrations  of  products, 
[E]  and  [F],  and  the  concentrations  of  reactants, 

[A]  and  [B] : 


\EY\FV 

■  /-.  rniK  =  K  =  a  constant  at 

constant  temperature 


(18) 


In  this  generalized  equation,  (18),  we  see  that 
again  the  numerator  is  the  product  of  the  equi- 
librium concentrations  of  the  substances  formed, 
each  raised  to  the  power  equal  to  the  number  of 
moles  of  that  substance  in  the  chemical  equation. 
The  denominator  is  again  the  product  of  the 
equilibrium  concentrations  of  the  reacting  sub- 
stances, each  raised  to  a  power  equal  to  the 
number  of  moles  of  the  substance  in  the  chemi- 
cal equation.  The  quotient  of  these  two  remains 
constant.  The  constant  K  is  called  the  equilib- 
rium constant.  This  generalization  is  one  of  the 
most  useful  in  all  of  chemistry.  From  the  equa- 
tion for  any  chemical  reaction  one  can  imme- 
diately write  an  expression,  in  terms  of  the  con- 
centrations of  reactants  and  products,  that  will 
be  constant  at  any  given  temperature.  If  this 
constant  is  measured  (by  measuring  all  of  the 
concentrations  in  a  particular  equilibrium  solu- 
tion), then  it  can  be  used  in  calculations  for  any 
other  equilibrium  solution  at  that  same  tempera- 
ture. 

In  Table  9-IV  are  listed  some  reactions  along 
with  the  equilibrium  law  relation  of  concentra- 
tions and  the  numerical  values  of  the  equilibrium 
constants.  First,  let's  verify  the  forms  of  the 
equilibrium  law  relation  among  the  concentra- 


Table  9-IV.     some  equilibrium  constants 


Cu(s)  +  2Ag+(aq)  +±  Cu+i(aq)  +  2Ag(s) 
Ag+(aq)  +  2NH3(aq)  +±  Ag(UH3h+(aq) 
N2Ot(g)  +±  2N02(g) 
2Hl(g)  :<=£  U2(g)  +  \2(g) 


HSOr(aq) 

CH3COOHfa</J 

AgC\(s) 
H20 

Ag\(s) 


H+(aq)  +  SOrHaq) 

H+(aq)  +  CH.COO-  (aq) 

Ag+(aq)  +  C\-(aq) 
H+(aq)  +  OH-(aq) 
Ag+(aq)  +  l-(aq) 


EQUILIBRIUM  LAW 
RELATION 


K  = 


[Cu+; 
[Ag+] 


[Ag(NFWl 

[Ag+][NH3p 

[N20<] 


K  = 


[H8][l«] 
[HI? 


fH+irsor'1 

[HSCV] 

[H+][CH;,COO-l 
[CHjCOOHJ 

K  =  [Ag+][C1-] 

A:  =  [H+][OH] 

K  =  [Ag+][1-] 


K  AT  STATED  TEMP. 


2  X  1015  at  25°C 

1.7  X  107at  25°C 
0.87  at  55°C 
0.018  at  423°C 
0.013  at  25°C 

1.8  X  10"5at  25°C 

1.7  X  10-'°  at  25°C 
10-"  at  25°C 
10-16  at  25°C 


154 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    |    CHAP.    9 


tions.  The  very  first  has  an  unexpected  form.  For 
the  reaction, 

Cu(s)  +  2Ag+(aq)  =  Cu+YaqJ  +  2Ag(s)     (19) 

you  do  not  find 

rr\.+2irAoi2 

(20) 


rcu+'irAgi*    K 

[Ag+]'[Cu] 


but  rather,  you  find 


[Cu^] 
[Ag+? 


(21) 


This  is  because  the  concentrations  of  solid  cop- 
per and  solid  silver  are  incorporated  into  the 
equilibrium  constant.  The  concentration  of  solid 
copper  is  fixed  by  the  density  of  the  metal — it 
cannot  be  altered  either  by  the  chemist  or  by  the 
progress  of  the  reaction.  The  same  is  true  of  the 
concentration  of  solid  silver.  Since  neither  of 
these  concentrations  varies,  no  matter  how  much 
solid  is  added,  there  is  no  need  to  write  them 
each  time  an  equilibrium  calculation  is  made. 
Equation  (21)  will  suffice. 


EXERCISE  9-4 

If  we  assign  the  equilibrium  constant  K'  to  ex- 
pression (20)  and  K  to  expression  (21), 

[Ag+ftCu]  [Ag+? 

show  that 

[Ag]2 


Another  K  of  unexpected  form  applies  to  the 
reaction 

H,0  +±:  H+(aq)  +  OH(aq)  (22) 

For  this  reaction  we  might  have  written 


[H20] 


(23) 


Instead,  Table  9-1 V  lists  expression  (24)  as 

[H+][OH]  =  K  (24) 

The  concentration  of  water,  [H20],  does  not 
appear  in  the  denominator  of  expression  (24). 
This  is  usually  done  in  treating  aqueous  reactions 


that  consume  or  produce  water.  It  is  justified 
because  the  variation  in  the  concentration  of 
water  during  reaction  is  so  slight  in  dilute  aque- 
ous solutions.  We  can  treat  [H2OJ  as  a  concen- 
tration that  does  not  vary.  Hence,  it  can  be  in- 
corporated in  the  equilibrium  constant. 


EXERCISE  9-5 

Water  has  a  density  of  one  gram  per  milliliter. 
Calculate  the  concentration  of  water  (expressed 
in  moles  per  liter)  in  pure  water.  Now  calculate 
the  concentration  of  water  in  0.10  M  aqueous 
solution  of  acetic  acid,  CHLCOOH,  assuming 
each  molecule  of  CH3COOH  occupies  the  same 
volume  as  one  molecule  of  H20. 


In  summary,  the  concentrations  of  solids  and 
the  concentrations  of  solvent  (usually  water)  can 
be  and  usually  are  incorporated  in  the  equilib- 
rium constant,  so  they  do  not  appear  in  the 
equilibrium  law  relation. 

Now  look  at  the  numerical  values  of  the  equi- 
librium constants.  The  ICs  listed  range  from  10+15 
to  10-16,  so  we  see  there  is  a  wide  variation.  We 
want  to  acquire  a  sense  of  the  relation  between 
the  size  of  the  equilibrium  constant  and  the  state 
of  equilibrium.  A  large  value  of  K  must  mean 
that  at  equilibrium  there  are  much  larger  con- 
centrations present  of  products  than  of  reactants. 
Remember  that  the  numerator  of  our  equilib- 
rium expression  contains  the  concentrations  of 
the  products  of  the  reaction.  The  value  of 
2  X  1015  for  the  K  for  reaction  (19)  certainly 
indicates  that  if  a  reaction  is  initiated  by  placing 
metallic  copper  in  a  solution  containing  Ag+ 
(for  example,  in  silver  nitrate  solution),  when 
equilibrium  is  finally  reached,  the  concentration 
of  Cu+2  ion,  [Cu+2],  is  very  much  greater  than 
the  square  of  the  silver  ion  concentration,  [Ag+]2. 

A  small  value  of  K  for  a  given  reaction  implies 
that  very  little  of  the  products  have  to  be  formed 
from  the  reactants  before  equilibrium  is  attained. 
The  value  of  K  =  10-16  for  the  reaction 

Agl(s)  +±:  Ag+(aq)  +  \(aq)  (25) 

indicates  that  very  little  solid  Agl  can  dissolve 


SEC.    9-2    |   QUANTITATIVE    ASPECTS    OF    EQUILIBRIUM 


155 


before  equilibrium  is  attained.  Silver  iodide  has 
extremely  low  solubility.  Conversely,  if  0.1  M 
solutions  of  KI  and  AgN03  are  mixed,  the  values 
of  [Ag+]  and  [I~]  are  large  and  the  equilibrium 
state  cannot  be  reached  until  the  [Ag+]  and  [I~] 
have  been  greatly  reduced  by  the  precipitation 
of  Agl. 

9-2.3    The  Law  of  Chemical  Equilibrium  Derived 
from  Rates  off  Opposing  Reactions 

Chemists  picture  equilibrium  as  a  dynamic  balance  be- 
tween opposing  reactions.  An  understanding  of  the  Law 
of  Chemical  Equilibrium  can  be  built  upon  this  basis. 
Consider  the  oxidation  of  nitric  oxide,  NO,  to  nitrogen 
dioxide,  Nd: 

2NOteJ  +  Oi(g)  — »■  mQh(g)  (26) 

The  reaction  to  the  right,  (/?),  proceeds  with  a  rate  that 
is  found  experimentally  to  depend  upon  the  concentra- 
tions of  the  reactants  as  follows : 

(rate)*  =  **[NO]*[0,]  (27) 

The  reverse  reaction  to  the  left,  (L),  has  been  studied  as 
well.  The  rate  of  this  reaction 

2NO(g)  +  Ot(g)  ■«—  2NCVSJ  (28) 

depends  upon  the  concentrations  as  follows: 

(rateU  =  *L[NO,]»  (29) 

Expressions  (27)  and  (29)  show  how  the  rates  of  reaction 
(26)  and  its  reverse,  reaction  (28),  depend  upon  the  con- 
centrations. Now  we  can  apply  our  microscopic  view  of 
the  equilibrium  state.  Chemical  changes  will  cease  (on 
the  macroscopic  scale)  when  the  rate  of  reaction  (26)  is 
exactly  equal  to  that  of  reaction  (28).  When  this  is  so, 
we  can  equate  expressions  (27)  and  (29): 

(rate)*  =  (rate)*,  (30) 

or 

**[NO]s[02]  =  fa[NO,]»  (31) 

By  algebraic  rearrangement,  equation  (31)  can  be  written 

k*  [NO,]'  2 

kL       [NO]*[0,]  v  } 

Since  both  kR  and  kL  are  constants  at  a  given  temperature, 
their  ratio  is  constant.  Hence  (32)  is  the  equilibrium  law 
expression  for  the  equilibrium 

2KO(g)  +  Oz(g)  q=b  2NO,(g) 

and  the  equilibrium  constant 


(33) 


kL 


(34) 


Thus  we  see  that  the  experimental  rate  laws  for  this 
reaction  and  its  reverse  lead  to  the  equilibrium  law.  In 
every  reaction  that  has  been  sufficiently  studied,  this  same 
result  is  obtained.  We  are  led  to  have  confidence  in  the 


molecular  view  of  equilibrium  as  a  dynamic  balance 
between  opposing  reactions. 


9-2.4    The  Factors  Which  Determine 
Equilibrium 

We  have  learned  much  about  equilibrium.  It  is 
characterized  by  constancy  of  macroscopic  prop- 
erties but  with  molecular  processes  continuing 
in  a  state  of  dynamic  balance.  At  equilibrium  we 
can  conclude  that  every  reaction  that  takes  place 
does  so  at  the  same  reaction  rate  as  its  reverse 
reaction. 

We  have  gone  further  and  discovered  that  the 
equilibrium  conditions  imply  a  constant  relation- 
ship among  the  concentrations  of  reactants  and 
products.  This  relationship  is  called  the  Law  of 
Chemical  Equilibrium.  Using  this  law,  we  can 
express  the  conditions  at  equilibrium  in  terms  of 
a  number  K,  called  the  equilibrium  constant. 

Despite  this  detailed  familiarity  with  equilib- 
rium, there  is  one  facet  we  have  not  considered 
at  all.  What  determines  the  equilibrium  con- 
stant? Why  does  one  reaction  favor  reactants 
and  another  reaction  favor  products?  What  fac- 
tors cause  sodium  chloride  to  have  a  large  solu- 
bility in  water  and  silver  chloride  to  have  a  low 
solubility?  Why  does  equilibrium  favor  the  re- 
action of  oxygen  with  iron  to  form  Fe203  (rust) 
but  not  the  reaction  of  oxygen  with  gold?  As 
scientists,  we  cannot  resist  wondering  what  fac- 
tors determine  the  conditions  at  equilibrium. 

This  is  the  activity  of  science  we  called  "Won- 
dering why"  (Section  1-1.3) — we  are  searching 
for  explanation.  An  explanation  is  a  likeness 
which  connects  the  system  under  study  with  a 
model  system  which  is  well  understood.  We 
might  begin  by  considering  Figure  9-7.  Here  we 
see  a  golf  bag  thrown  into  the  rear  of  a  station 
wagon.  Unfortunately,  the  ball  pocket  is  open 
and  all  of  the  golf  balls  have  spilled  out  onto  the 
floor  of  the  station  wagon.  Because  the  floor  has 
a  step  in  it,  the  golf  balls  on  the  upper  level  pos- 
sess some  potential  energy  (energy  of  position). 
The  golf  balls  tend  to  roll  to  the  lower  level 
spontaneously,  as  shown  in  Figure  9-8.  As  a  golf 
ball  does  this,  its  potential  energy  becomes  ki- 
netic energy  (energy  of  motion).  Finally,  this 


156 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    ]    CHAP.    9 


Fig.  9-7.  Golf  balls  rolling  on  the  floor  of  a  station 
wagon. 


kinetic  energy  is  dissipated  into  heat.  Now  the 
golf  balls  lie  at  rest  at  the  lower  floor  level. 

This  situation  has  some  similarities  to  the 
chemical  change  in  a  spontaneous,  exothermic 
reaction.  The  reactants  of  high  heat  content  re- 
act spontaneously  to  form  products  of  lower 
heat  content.  As  each  molecular  reaction  occurs, 
the  excess  heat  content  becomes  kinetic  energy. 
The  product  molecules  separate  from  each  other 
with  high  kinetic  energy.  As  they  collide  with 


other  molecules,  this  energy  is  dissipated  into 
heat.  Figure  9-8  shows  this  through  a  heat  con- 
tent diagram  for  the  chemical  reaction. 

(1)  There  are  two  states  of  each  system: 

Initial  State  Final  State 


Golf  balls:    on  upper  level 
Reaction:      reactants 


on  lower  level 
products 


Fig.  9-8.  Comparison  of  a  chemical  reaction  to  golf 
balls  rolling  downhill. 


Ht3h 

heat 

content 


\ 


Low 

heat 

content 


SEC.    9-2    I    QUANTITATIVE    ASPECTS    OF    EQUILIBRIUM 


157 


(2)  The  potential  energy  of  the  initial  state  is 
higher  than  the  potential  energy  of  the  final 
state: 


Golf  balls: 


Reaction: 


Initial  State 

high  potential 
energy 

high  heat 
content 


Final  State 

low  potential 
energy 

low  heat 
content 


(3)  As  the  change  from  initial  state  to  final  state 
proceeds,  the  form  of  the  energy  changes: 


Initial  State 

Golf  balls:     potential 
energy 

Reaction:      heat  content 


Final  State 

kinetic  energy, 
and  then  heat 

molecular  kinetic 
energy,  and 
then  heat 


a  liquid.  Water  evaporates  spontaneously 
but  it  absorbs  heat  as  it  does  so.  It  is  not 
"rolling  downhill"  energetically.  When  am- 
monium chloride  dissolves  in  water,  the  solu- 
tion becomes  cooler.  Again  heat  is  absorbed 
— yet  the  ammonium  chloride  goes  ahead 
and  dissolves. 
(2)  Another  difficulty  is  that  spontaneous  chemi- 
cal reactions  do  not  go  to  completion.  Even 
if  a  spontaneous  reaction  is  exothermic,  it 
proceeds  only  till  it  reaches  equilibrium.  But 
in  our  golf  ball  analogy,  "equilibrium"  is 
reached  when  all  of  the  golf  balls  are  on  the 
lower  level.  Oui  analogy  would  lead  us  to 
expect  that  an  exothermic  reaction  would 
proceed  until  all  of  the  reactants  are  con- 
verted to  products,  not  to  a  dynamic  equi- 
librium. 


(4)  The  changes  from  initial  to  final  state  pro- 
ceed spontaneously  toward  lowest  potential 
energy,  the  direction  corresponding  to  "roll- 
ing downhill": 


Initial  State 


Final  State 


Golf  balls: 
Reaction: 


spontaneous 
spontaneous 


Having  established  these  similarities,  we  might 
offer  a  possible  generalization: 

Since:  golf  balls  always  roll  downhill  spon- 
taneously 

Perhaps:  reactions  always  proceed  spontane- 
ously in  the  direction  toward  minimum 
energy. 

This  proposal  leads  us  to  expect  that  a  reaction 
will  tend  to  proceed  spontaneously  if  the  prod- 
ucts have  lower  energy  than  the  reactants.  This 
expectation  is  in  accord  with  experience  with 
many  reactions,  especially  for  those  which  re- 
lease a  large  amount  of  heat. 

There  are  two  basic  and  serious  difficulties 
with  our  proposed  explanation.  (Remember 
"Cylindrical  objects  burn"?) 

(1)  Some  endothermic  reactions  proceed  spon- 
taneously. One  example  is  the  evaporation  of 


Because  of  these  failures,  we  need  to  alter  our 
proposed  explanation.  We  must  seek  a  new  anal- 
ogy that  gives  a  better  correspondence  with  the 
behavior  of  chemical  reactions.  How  should  we 
alter  our  golf  ball  analogy  to  bring  it  into  better 
accord  with  experimental  facts?  Here  is  a  pos- 
sible view. 

Consider  how  the  golf  ball  situation  shown  in 
Figure  9-8  will  change  when  the  station  wagon 
is  driven  over  a  bumpy  road.  Now  the  golf  balls 
are  shaken  and  jostled  about;  they  roll  around 
and  collide  with  each  other.  Every  now  and  then 
one  of  the  golf  balls  even  accumulates  enough 
energy  (through  collisions)  to  return  to  the  upper 
level  of  the  station  wagon  floor.  Of  course^  any 
golf  ball  that  is  bounced  up  tends  to  roll  back 
down  to  the  lower  level  a  little  later.  As  this 
bumpy  ride  continues,  a  state  is  reached  in  which 
golf  balls  are  being  jostled  up  to  the  higher  level 
at  the  same  rate  they  are  rolling  back  down  to 
the  lower  level.  Then  "equilibrium"  exists.  Some 
of  the  golf  balls  are  on  the  lower  level  and  some 
on  the  upper  level.  Since  the  rate  of  rolling  up 
equals  the  rate  of  rolling  down,  a  dynamic  bal- 
ance exists. 

This  analogy  solves  the  problems  of  the  sim- 
pler "golf  balls  roll  downhill"  picture.  The 
bumpy  road  model  contains  a  new  feature  that 
gives  a  basis  for  expecting  "reaction"  in  the 


158 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    I    CHAP.    9 


High 

heat 

content- 


\ 


heat 
content 


iiiiih 


*\ 


F/g.  9-9.  Golf  balls  rolling  on  the  floor  of  a  station 
wagon  driving  on  a  bumpy  road. 


endothermic  direction.  Some  golf  balls  roll  uphill 
if  they  are  shaken  hard  enough.  The  tendency  to 
roll  back  down  will  always  keep  them  coming 
back  to  the  lower  level  and,  finally,  equilibrium 
will  be  reached  when  the  rate  of  rolling  down 
equals  the  rate  of  jostling  up. 

What  happens  if  the  road  becomes  smoother? 
The  "jostling  up"  reaction  is  less  favored— the 
equilibrium  conditions  change  in  favor  of  the 
golf  balls  at  the  lower  level. 

Now  turn  to  the  chemical  reaction.  What  fea- 
ture in  a  reacting  chemical  system  corresponds 
to  the  jostling  of  the  bumpy  road  in  our  analogy? 
It  is  the  temperature.  At  any  temperature  except 
absolute  zero  there  is  a  constant  random  jostling 
of  the  molecules.  Some  molecules  have  low  ki- 
netic energies,  some  have  high  kinetic  energies — 
we  looked  at  the  distribution  of  the  energies  in 
Chapter  8  (see  Figure  8-4,  page  131).  Some  of  the 
molecules  will  occasionally  accumulate  enough 
energy  to  "roll  uphill"  to  less  stable  molecular 
forms.  On  the  one  hand,  molecular  changes  will 
take  place  in  the  direction  of  minimum  energy. 


On  the  other  hand,  the  molecular  changes  will 
finally  reach  a  dynamic  equilibrium  when  the 
random  jostlings  or  energy  transfers  at  the  tem- 
perature of  the  system  are  restoring  molecules  to 
the  molecular  forms  of  higher  energy  at  the  same 
rate  as  they  are  "rolling  downhill"  to  the  lower 
energy  forms. 

Now  we  have  an  analogy  that  does  aid  us  in 
understanding  chemical  reactions  and  equilib- 
rium. We  can  see  the  following  features  of  chem- 
ical reactions: 

(1)  Chemical  reactions  proceed  spontaneously 
to  approach  the  equilibrium  state. 

(2)  One  factor  that  fixes  the  equilibrium  state  is 
the  energy.  Equilibrium  tends  to  favor  the 
state  of  the  lowest  energy. 

(3)  The  other  factor  that  fixes  the  equilibrium 
state  is  the  randomness  implied  by  the  tem- 
perature. Equilibrium  tends  to  favor  the 
state  of  greatest  randomness. 

(4)  The  equilibrium  state  is  a  compromise 
between  these  two  factors,  minimum  en- 
ergy and  maximum  randomness.  At  very 
low  temperatures,  energy  tends  to  be  the 
more  important  factor.  Then  equilibrium 
favors  the  molecular  substances  with   the 


SEC.    9-2    |    QUANTITATIVE    ASPECTS    OF    EQUILIBRIUM 


159 


lowest  heat  content.  At  very  high  tempera- 
tures, randomness  becomes  more  important. 
Then  equilibrium  favors  a  random  distribu- 
tion among  reactants  and  products  without 
regard  for  energy  differences. 

Our  analogy  can  be  stretched  one  point  fur- 
ther. We  might  ask  whether  the  relative  area  of 
the  upper  floor  level  compared  with  that  of  the 
lower  floor  level  has  any  bearing  on  the  distribu- 
tion of  golf  balls.  After  all,  if  the  upper  level  area 
is  small,  as  in  Figure  9-7,  few  golf  balls  are  likely 
to  remain  there.  Contrast  Figure  9-10,  in  which 
the  golf  bag  has  been  removed.  Now  the  upper 
floor  level  has  much  more  area.  Golf  balls  which 
reach  the  upper  level  will  now  have  a  great  deal 
of  space.  They  can  roll  around  longer  before 
returning  to  the  lower  level.  The  effect  of  extend- 
ing the  upper  level  will  be  to  increase  the  fraction 
of  the  golf  balls  that  occupy  the  upper  level  at 
"equilibrium." 

This  extension  of  the  analogy  increases  its 
value  in  considering  chemical  reactions.  The 
simplest  example  is  probably  the  vaporization  of 
a  liquid.  It  is  true  that  the  molecules  have  lower 
energy  when  they  cluster  together  tightly  in  the 
liquid  state.  On  the  other  hand,  the  gaseous  state 
provides  a  broad  upper  level.  Every  molecule 
which  vaporizes  has  an  amount  of  space  avail- 


Fig.  9-10.  Golf  balls  rolling  on  the  floor  of  a  station 
wagon.  The  effect  of  extending  the  upper 
level. 


able  to  it  much  larger  than  it  had  in  the  crowded 
liquid.  This  "available  space"  factor,  accom- 
panied by  the  random  jostlings  of  temperature  to 
overcome  the  potential  energy  difference,  aids 
vaporization. 

Now  let  us  review  what  happens  as  we  warm 
a  solid  substance  from  a  very  low  temperature 
to  a  very  high  temperature.  As  the  temperature 
is  raised,  small  energy  differences  become  unim- 
portant. Thus,  if  the  temperature  of  the  solid  is 
raised  too  much,  the  lower  energy  of  the  regular 
solid  becomes  unimportant  compared  with  the 
random  thermal  energies.  The  solid  melts,  sur- 
rendering this  energy  stability  in  return  for  the 
randomness  of  the  liquid  state.  If  the  tempera- 
ture is  raised  still  further,  the  energies  of  attrac- 
tion among  the  molecules  become  unimportant 
compared  with  the  random  thermal  energies. 
Then  the  liquid  vaporizes,  surrendering  the  lower 
potential  energy  afforded  by  the  molecules  re- 
maining close  together  in  favor  of  the  still  higher 
randomness  of  the  gaseous  state.  If  we  raise  the 
temperature  still  further,  the  energies  that  hold 
molecules  together  begin  to  become  unimportant 
compared  with  random  thermal  energies.  Fi- 
nally, at  extremely  high  energies,  molecules  no 
longer  exist — all  is  chaos.  This  is  the  chemical 
situation  within  the  Sun.  Since  at  such  high  en- 
ergies chemical  reactions  become  unimportant, 
all  chemists  on  the  Sun  are  out  of  work.  We'd 
better  return  to  room  temperature  to  apply  our 
knowledge  of  equilibrium  to  chemical  systems 
within  our  interest. 


160 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    |    CHAP.    9 


QUESTIONS  AND  PROBLEMS 

1.  Which  of  the  following  are  equilibrium  situa- 
tions and  which  are  steady  state  situations? 

(a)  A  playing  basketball  team  and  the  bench  of 
reserves.  The  number  of  players  "in  the 
game"  and  the  number  "on  the  bench"  are 
constant. 

(b)  The  liquid  mercury  and  mercury  vapor  in  a 
thermometer.  Temperature  is  constant. 

(c)  Grand  Coulee  dam  and  the  lake  behind  it. 
Water  level  is  constant,  though  a  river  flows 
into  the  lake. 

(d)  A  well-fed  lion  in  a  cage.  The  lion's  weight 
is  constant. 

Answer,  (a)  and  (b)  are  equilibrium  situations. 

2.  Which  of  the  following  are  equilibrium  situa- 
tions and  which  are  steady  state  situations? 

(a)  A  block  of  wood  floating  on  water. 

(b)  During  the  noon  hour,  the  water  fountain 
constantly  has  a  line  of  ten  persons. 

(c)  When  a  capillary  tube  is  dipped  in  water, 
water  rises  in  the  capillary  (because  of  sur- 
face tension)  to  a  height  /;  and  remains  con- 
stant there. 

(d)  The  capillary  system  of  (c)  considered  over 
such  a  long  period  that  evaporation  of  water 
out  the  end  of  the  capillary  cannot  be  neg- 
lected. 

(e)  At  a  particular  point  in  the  reaction  chamber 
of  a  jet  engine,  the  gas  composition  (fuel, 
air,  and  products)  is  constant. 

3.  What,  specifically,  is  "equal"  in  a  chemical  re- 
action that  has  attained  a  state  of  equilibrium? 

4.  One  drop  of  water  may  or  may  not  establish  a 
state  of  vapor  pressure  equilibrium  when  placed 
in  a  closed  bottle.  Explain. 

5.  Why  are  chemical  equilibria  referred  to  as 
"dynamic"? 

6.  What  do  the  following  experiments  (done  at 
25°C)  show  about  the  state  of  equilibrium? 

(a)  One  liter  of  water  is  added  a  few  milliliters 
at  a  time,  to  a  kilogram  of  salt  which  only 
partly  dissolves. 


(b)  A  large  salt  shaker  containing  one  kilogram 
of  salt  is  gradually  emptied  into  one  liter  of 
water.  The  same  amount  of  solid  dissolves 
as  in  (a). 

7.  The  following  chemical  equation  represents  the 
reaction  between  hydrogen  and  chlorine  to  form 
hydrogen  chloride: 

H,(g)  +  C\,(g)  ^±  mC\(g)  +  44.0  kcal 

(a)  List  four  important  pieces  of  information 
conveyed  by  this  equation. 

(b)  What  are  three  important  areas  of  interest 
concerning  this  reaction  for  which  no  infor- 
mation is  indicated? 

8.  How  does  a  catalyst  affect  the  equilibrium  con- 
ditions of  a  chemical  system? 

9.  In  any  discussion  of  chemical  equilibrium  why 
are  concentrations  always  expressed  in  moles, 
rather  than  in  grams,  per  unit  volume? 

10.  If  the  phase  change  represented  by 

Heat  +  H,Of JJ  +±  HjO(g) 
has  reached  equilibrium  in  a  closed  system: 

(a)  What  will  be  the  effect  of  a  reduction  of  vol- 
ume, thus  increasing  the  pressure? 

(b)  What  will  be  the  effect  of  an  increase  in  tem- 
perature? 

(c)  What  will  be  the  effect  of  injecting  some 
steam  into  the  closed  system,  thus  raising 
the  pressure? 

11.  Methanol  (methyl  alcohol)  is  made  according  to 
the  following  net  equation: 


CO(g)  +  2H«(g) 


CH3OH(gJ  +  heat 


Predict  the  effect  on  equilibrium  concentrations 
of  an  increase  in:  (a)  Temperature,  (b)  Pressure 

Answer,  (a)  Decreases  CH3OH 
(b)  Increases  CH3OH. 

12.  Consider  the  reaction: 

4HCl(g)  +  02(g)  +± 

2H2Ofgj  +  2Ch(g)  +  27  kcal 


QUESTIONS    AND    PROBLEMS 


161 


What  effect  would  the  following  changes  have  on 
the  equilibrium  concentration  of  Cl2?  Give  your 
reasons  for  each  answer. 

(a)  Increasing  the  temperature  of  the  reaction 
vessel. 

(b)  Decreasing  the  total  pressure. 

(c)  Increasing  the  concentration  of  02. 

(d)  Increasing  the  volume  of  the  reaction  cham- 
ber. 

(e)  Adding  a  catalyst. 

13.  Write  the  equation  for  the  dissociation  of  Hl(g) 
into  its  elements. 

(a)  Will  HI  dissociate  to  a  greater  or  a  lesser 
extent  as  the  temperature  is  increased? 
AH  =  -6.2  kcal/mole  H\(g). 

(b)  How  many  grams  of  iodine  will  result  if  at 
equilibrium  0.050  mole  of  HI  has  disso- 
ciated ? 

14.  Consider  two  separate  closed  systems,  each  at 
equilibrium: 

(a)  HI  and  the  elements  from  which  it  is  formed, 

(b)  H2S  and  the  elements  from  which  it  is 
formed. 

What  would  happen  in  each  if  the  total  pressure 
were  increased?  Assume  conditions  are  such  that 
all  reactants  and  products  are  gases. 

15.  Each  of  the  following  systems  has  come  to  equi- 
librium. What  would  be  the  effect  on  the  equilib- 
rium concentration  (increase,  decrease,  no 
change)  of  each  substance  in  the  system  when 
the  listed  reagent  is  added? 


REACTION 

REAGENT 

(a)  CtfVf  J  +=±H/£J  +  CHt(g) 

H,(g) 

(b)  Cu-wfaqj  +  4NH3f£j  q=>= 

Cu(NH.dt2(«q) 

CuSO/sj 

(C)   Ag+(aq)  +  C\(aq)  q=>: 

AgClCsJ 

AgC\(s) 

(d)  PbSO/sJ  +  H+(aq)  3=fc 

Pb+2(aq)  +  USO^(aq) 

Pb(NOa), 

solution 

(e)  CO(g)  +  \0,(g)  q=b 

CO>(g)  -f  heat 

heat 

16.  Nitric  oxide,  NO,  releases  13.5  kcal/mole  when 
it  reacts  with  oxygen  to  give  nitrogen  dioxide. 
Write  the  equation  for  this  reaction  and  predict 
the  effect  of  (i)  raising  the  temperature,  and  of 
(ii)  increasing  the  concentration  of  NO  (at  a 
fixed  temperature)  on: 

(a)  the  equilibrium  concentrations; 

(b)  the  numerical  value  of  the  equilibrium  con- 
stant; 

(c)  the  speed  of  formation  of  N02. 


17.  Given: 

S02(g)  +  \Q2(g) 


SQ3(g)  +  23  kcal 


(a)  For  this  reaction  discuss  the  conditions  that 
favor  a  high  equilibrium  concentration  of 
S03. 

(b)  How  many  grams  of  oxygen  gas  are  needed 
to  form  1.00  gram  of  SQ3? 


Answer.  0.200  gram  of  02. 


18.  Given: 


CaCO/sJ  =?=fc  CaO(s)  +  CO,(g) 

(closed  system) 

At  a  fixed  temperature,  what  effect  would  adding 
more  CaC03  have  on  the  concentration  of  C02 
in  the  region  above  the  solid  phase?  Explain. 


19.  Given: 

Wg)  +  h(g) 


2Hl(g)  (closed  system) 


Answer,  (a)  C2H6  (increase),  H2  (increase), 
QH4  (decrease). 


At  450°C,  A:  =  50.0  for  the  above  reaction. 
What  is  the  equilibrium  constant  of  the  reverse 
reaction  at  450°C? 

20.  Write  the  expression  indicating  the  equilibrium 
law  relations  for  the  following  reactions. 

(a)  Wg)  +  3H2f*J  *±  2NH3(g) 

(b)  CO(g)  +  N02(g)  ^^  CQ2(g)  +  NO(g) 

(c)  Zn(s)  +  2Ag+f«q;  +±  Zn^(aq)  +  2Ag(s) 

(d)  Pbh(s)  +±  Pb+*(aq)  +  2l~(aq) 

(e)  CN-(aq)  +  H20(/;  +=± 

HCNfaqJ  +  OH-faqJ 

21.  Equilibrium  constants  are  given  for  several  sys- 
tems below.  In  which  case  does  the  reaction  as 
written  occur  to  the  greatest  extent  ? 


162 


EQUILIBRIUM    IN    CHEMICAL    REACTIONS    I    CHAP.    9 


REACTION 


(a)  CHzCOOHdiq)  +± 

H+(aq)  +  CH3COO(aq) 

(b)  CdS(s)  +±: 

Cd^(aq)  +  S^(aq) 

(C)  H+(aq)  +  HS(aq)  +± 

HS(aq) 


1.8  X  10-& 
7.1  X  10-28 
1      X  107 


22.  In  the  reaction 
2Hl(g) 


Wg)  +  h(g) 


at  448°C  the  partial  pressures  of  the  gas  at 
equilibrium  are  as  follows: 

[HI]  =  4  X  10-3atm; 
[H,]  =  7.5  X  10-3atm; 
[I,]    =  4.3  X  10-B  atm. 

What  is  the  equilibrium  constant  for  this  re- 
action? 

23.  Reactants  A  and  B  are  mixed,  each  at  a  concen- 
tration of  0.80  mole/liter.  They  react  slowly,  pro- 
ducing C  and  D:  A  +  B  +±  C  +  D.  When 
equilibrium  is  reached,  the  concentration  of  C 
is  measured  and  found  to  be  0.60  mole/liter. 
Calculate  the  value  of  the  equilibrium  constant. 

Answer.  K  =  9.0. 

24.  The  water  gas  reaction 

C02(g)  +  H2(g)  *=±  CO(g)  +  H20(g) 

was  carried  out  at  900°C  with  the  following 
results. 


PARTIAL  PRESSURE, 
ATM  AT  EQUILIBRIUM 


TRIAL 

NO. 
1 

2 
3 


CO 

0.352 
0.266 
0.186 


H.Q 

0.352 
0.266 
0.686 


CO, 

0.648 
0.234 
0.314 


H- 


0.148 
0.234 
0.314 


(a)  Write  the  equilibrium  constant  expression. 

(b)  Verify  that  the  expression  in  (a)  is  a  con- 
stant, using  the  data  given. 

25.  Select  from  each  of  the  following  pairs  the  more 
random  system. 

(a)  A  brand  new  deck  of  cards  arranged  accord- 
ing to  suit  and  number. 

(a')  The  same  deck  of  cards  after  shuffling. 

(b)  A  box  full  of  sugar  cubes. 

(b')  Sugar  cubes  thrown  on  the  floor. 

(c)  A  hay  stack. 

(c')  Stacked  fire  wood. 

26.  For  each  of  the  following  reactions,  state:  (i) 
whether  tendency  toward  minimum  energy  fa- 
vors reactants  or  products,  (ii)  whether  tendency 
toward  maximum  randomness  favors  reactants 
or  products. 

(a)  H20(l)  +±  H20(s) 

(b)  H20(l)  +±:  H20(g) 

(c)  CaCCVsj  +  43  kcal 

(d)  h(s)  +  1.6  kcal  +± 


AH  =  -1.4  kcal 
AH  =  +10  kcal 

^±  CaO(s)  +  C02(g) 

I2  (in  alcohol) 


(e)  4Fe(s)  +  3Q2(g)  +±  2Fe-,03(s)  +  400  kcal 


CHAPTER 


10 


Solubility 
Equilibria 


•  •    •   solubility   •    •    •   depends  fundamentally  upon  the  ease  with  which 

•  •  •  two  molecular  species  are  able  to  mix,  and  if  the  two  species  dis- 
play a  certain  hostility  toward  mixing,  •  •  •  saturation  [is]  attained  at 
smaller  concentration   •    •    • 

JOEL    H.    HILDEBRAND,    1936 


The  principles  of  equilibrium  have  wide  applica- 
bility and  great  utility.  For  example,  they  aid  us 
in  understanding  and  controlling  the  solubility 
of  solids  and  gases  in  liquids.  We  shall  consider, 
first,  the  solubility  of  a  molecular  solid  in  a  liq- 


uid, then  the  solubility  of  a  gas  in  a  liquid.  The 
usefulness  of  equilibrium  principles  is  even 
greater  in  treating  the  solubilities  of  ionic  solids 
in  water.  Much  of  this  chapter  will  be  devoted  to 
aqueous  solutions  of  ionic  solids. 


10-1    SOLUBILITY:  A  CASE  OF  EQUILIBRIUM 


The  starting  point  in  any  quantitative  equilib- 
rium calculation  is  the  Equilibrium  Law.  For  a 
generalized  reaction, 


aA  +  bB  + 


eE  +  fF  + 


(/) 


equilibrium  exists  when  the  concentrations  sat 
isfy  the  relation 


K  =  a  constant 


(2) 


[A]'[B?  •  •  • 
First,  we  shall  apply  expression  (2)  to  the  solu 


tion  system  of  solid  iodine  dissolving  in  liquid 
ethyl  alcohol. 

10-1.1    The  Solubility  of  Iodine 
in  Ethyl  Alcohol 

As  a  solid  dissolves  in  a  liquid,  atoms  or  mole- 
cules leave  the  solid  and  become  part  of  the 
liquid.  These  atoms  or  molecules  may  carry  no 
charge  (then  they  are  electrically  neutral)  or  they 
may  be  ions.  The  iodine-alcohol  system  is  of  the 

163 


164 


SOLUBILITY    EQUILIBRIA    I    CHAP.    10 


former  kind.  As  iodine  dissolves,  neutral  mole- 
cules of  iodine,  I2,  leave  the  regular  crystal  lattice 
and  these  molecules  become  part  of  the  liquid 
phase.  At  equilibrium,  excess  solid  must  remain 
and  a  fixed  concentration  of  iodine  is  present  in 
solution.  This  fixed  concentration  is  called  the 
solubility. 
For  this  system  at  equilibrium,  the  reaction  is 

I2(solid)  =  I2(alcohol  solution)  (3) 

The  Equilibrium  Law  applied  to  this  reaction 
gives 

K  =  a  constant  =  [concentration  I2  in  alcohol] 

K  =  [I,]  (4) 


crystal  than  in  the  solution.  The  potential  energy 
must  rise  as  a  molecule  leaves  the  crystal  and  the 
principles  that  govern  rates  of  reaction  are  op- 
erative. Presumably  there  is  an  activated  complex 
for  the  process.  The  rate  at  which  molecules  leave 
a  square  centimeter  of  surface,  passing  over  the 
energy  barrier,  is  determined  by  the  height  of  the 
barrier  and  by  the  temperature.  We  can  call  this 
rate  kd.  Changing  the  temperature  does  not  affect 
the  activated  complex,  but  the  molecular  energy 
distribution  is  altered  (see  Figure  8-3,  p.  131). 
Hence,  kd  is  a  function  of  temperature.  These 
two  factors  determine  completely  the  rate  of  dis- 
solving: 

/rate  molecules  leave \ 


/  surface  \       /rate  molecules  leave \ 
(rate  of  dissolving)  =  (  area      )  X  I  a  square  centimeter  j 

\  of  crystal  surface      / 


-        (A) 
(rate  of  dissolving)  =  A  X  kd 

10-1.2    The  Dynamic  Nature  of  Solubility 
Equilibrium 

The  simple  form  of  the  equilibrium  expression 
(4)  follows  directly  from  the  dynamic  nature  of 
the  solubility  equilibrium.  There  must  be  a  dy- 
namic balance  between  the  rate  that  iodine  mole- 
cules leave  the  crystal  and  the  rate  that  iodine 
molecules  return  to  the  crystal.  To  understand 
this  dynamic  balance,  we  must  consider  the  fac- 
tors that  determine  these  two  rates. 

RATE   OF   DISSOLVING 

One  of  the  factors  that  influences  the  rate  of  dis- 
solving of  solid  is  the  area,  A,  of  the  crystal 
surface  that  contacts  the  liquid.  If  many  crystals 
(with  large  A)  are  dissolving  simultaneously,  the 
rate  of  dissolving  is  faster  than  if  only  a  few 
crystals  (with  small  A)  are  in  the  solvent.  The 
rate  of  dissolving  is  proportional  to  this  liquid- 
solid  surface  area,  A. 
A  molecule  of  iodine  is  more  stable  in  the 


X 


(*-> 


(5) 


RATE   OF   PRECIPITATION 


The  rate  of  precipitation  is  the  rate  at  which 
molecules  return  to  the  surface  and  fit  into  the 
crystal  lattice.  To  do  this,  the  molecules  in  solu- 
tion must  first  strike  the  crystal  surface.  Again, 
the  more  surface,  the  more  frequently  will  dis- 
solved molecules  encounter  a  piece  of  crystal. 
The  rate  of  precipitation  is  proportional  to  A. 
In  addition,  the  rate  that  molecules  strike  the 
surface  depends  upon  how  many  molecules  there 
are  per  unit  volume  of  solution.  As  the  concen- 
tration rises,  more  and  more  molecules  strike 
the  surface  per  unit  time.  The  rate  of  precipita- 
tion is  proportional  to  the  iodine  concentration, 

[hi 

The  last  factor  is,  again,  the  rate  that  mole- 
cules can  pass  over  the  energy  barrier — the  acti- 
vated complex  for  precipitation.  Again  there  is 
a  rate  constant,  kp,  that  is  determined  by  tem- 
perature and  the  height  of  the  energy  barrier  to 
precipitation. 

We  have,  then,  three  factors  that  determine 
the  rate  of  precipitation: 


(rate  of  precipitation)  =  f  surfaceVconcentrati°n  ofWrate  dissolved  molecules  pass\ 
\area      /\dissolved  I2         /\over  activation  energy  barrier/ 

=        (A)     X  [I2]  X  (*,) 

(rate  of  precipitation)  =  A  X  kp  X  [I2] 


(6) 


SEC.    10-1   l  solubility:   a  case  of  equilibrium 


165 


THE    DYNAMIC    NATURE 
OF    EQUILIBRIUM 

At  equilibrium,  we  can  equate  (5)  and  (6): 

(rate  of  dissolving)  =  (rate  of  precipitation) 

A  X  kd  =  A  X  kp  X  [I2]  (7) 

The  area  of  contact,  A,  appears  both  on  the  left 
and  on  the  right  of  expression  (7).  Hence,  it 
cancels  out.  Dividing  both  sides  of  (7)  by  kp,  we 
obtain 

^  =  [I2]  (8) 

Since  kd  and  kp  each  depend  upon  temperature, 
their  ratio  depends  upon  temperature.  Other- 


Fig.  10-1.  Solubility  equilibrium  is  dynamic. 

THE  DISSOLVING    PROCESS 


wise,  however,  each  is  constant.  We  can  write 
K  =  [h]  00 


where 


K~  k„ 


Thus,  by  expressing  the  dynamic  balance  be- 
tween the  rates  of  dissolving  and  precipitation, 
we  obtain  (4).  The  concentration  of  I2  at  equi- 
librium is  a  constant,  fixed  by  the  temperature. 
This  constant  equals  the  solubility. 

10-1.3    The  Factors  That  Fix  Solubility 
of  a  Solid 

All  of  the  discussion  we  have  just  applied  to  the 
dissolving  of  iodine  in  ethyl  alcohol  applies 
equally  well  to  the  dissolving  of  iodine  in  carbon 

THE  PRECIPITATION  PROCESS 


Area    A 


\ 


Solid 


Solution 


So  lut ion 


Solid 


fd*sfol£j  =  *d  *  ra~4  (/r^itLon)  =    *J>  f~1>  (concentration) 

At  equilibrium,       {j****  °f    )  =  /  Rate .  °f         ) 
I  dissolving/         [precipitation/ 


166 


SOLUBILITY    EQUILIBRIA    |    CHAP.    10 


tetrachloride,  CC14.  Iodine  at  room  temperature 
dissolves  in  carbon  tetrachloride  at  a  certain  rate 
that,  at  equilibrium,  exactly  equals  the  rate  of 
precipitation.  Again  we  reach  the  simple  equi- 
librium expression 


K=jf=  M 


(4) 


Despite  this  qualitative  similarity,  the  solu- 
bility of  iodine  in  CCU  is  very  different  from  its 
solubility  in  alcohol.  One  liter  of  alcohol  dis- 
solves 0.84  mole  of  iodine,  whereas  one  liter  of 
CCU  dissolves  only  0.12  mole: 

^alcohol  =  0.84  mole/liter  (9) 

Kccu  =  0.12  mole/liter  (10) 

Why  are  these  constants  so  different?  To  see 
why,  we  must  turn  to  the  two  factors  that  control 
every  equilibrium,  tendency  toward  minimum 
energy  and  tendency  toward  maximum  random- 
ness. 

THE    EFFECT    OF    RANDOMNESS 

In  either  solvent,  alcohol  or  carbon  tetrachloride, 
the  dissolving  process  destroys  the  regular  crys- 
tal lattice  of  iodine  and  forms  the  disordered 
solution.  The  dissolving  process  increases  ran- 


Fig.  10-2.  A    large    energy   difference    between    solid 
and  solution  lowers  the  solubility. 


domness.  The  tendency  toward  maximum  ran- 
domness tends  to  cause  solids  to  dissolve. 

THE    EFFECT    OF    ENERGY 

Experiment  shows  that  heat  is  absorbed  as  iodine 
dissolves.  The  regular,  ideally  packed  iodine 
crystal  gives  an  iodine  molecule  a  lower  potential 
energy  than  does  the  random  and  loosely  packed 
solvent  environment.  We  see  that  the  second 
factor,  tendency  toward  minimum  energy,  favors 
precipitation  and  growth  of  the  crystal. 

Now  we  see  the  opposing  factors  at  equilib- 
rium. To  increase  randomness,  solid  tends  to  dis- 
solve. To  lower  energy,  solid  tends  to  precipitate. 
Equilibrium  is  reached  when  the  concentration 
is  such  that  these  two  tendencies  are  equal. 

How  much  the  energy  factor  favors  the  crystal 
depends  upon  the  change  in  heat  content  as  a 
mole  of  solid  dissolves.  This  change  is  called  the 
heat  of  solution.  The  heats  of  solution  of  iodine 
in  these  two  solvents  have  been  measured;  they 
are  as  follows: 


h(s)  +  1.6  kcal 
h(s)  +  5.8  kcal 


I2(in  alcohol) 
I2(in  CCU) 


(11) 
(12) 


We  see  that  there  is  a  much  greater  energy  rise 
when  a  mole  of  I2  dissolves  in  CC14  than  when 
a  mole  of  I2  dissolves  in  alcohol.  Thus  the  energy 
factor  (favoring  the  crystal)  that  opposes  the 
randomness  factor  (favoring  solution)  is  much 


CCL 


Solid 


I-    iri    alcohol 


Solid    Iz 


/ 


A    large    energy    effect 
opposing    dissolving 


.A    small  energy  effect 
opposing    dissolving 


SEC.   10-1   I  solubility:   a  case  of  equilibrium 


167 


larger  for  CC14  than  for  alcohol.  The  solubility 
of  iodine  in  CC14  is  not  as  high  as  it  is  in  alcohol. 
This  energy  rise  establishes,  for  this  case,  the 
"hostility"  toward  mixing  referred  to  in  the  quo- 
tation at  the  beginning  of  the  chapter.  The  larger 
the  "hostility,"  as  measured  by  heat  absorbed 
on  mixing,  the  lower  will  be  the  solubility. 

THE    EFFECT    OF    TEMPERATURE 

Raising  the  temperature  always  tends  to  favor 
the  more  random  state.  For  these  solvents,  this 
means  that  more  solid  will  dissolve,  since  the 
liquid  solution  is  more  random  than  the  solid. 
The  solubility  of  iodine  increases  as  temperature 
is  raised,  both  in  alcohol  and  in  carbon  tetra- 
chloride. 


EXERCISE  10-1 

The  heat  of  solution  of  iodine  in  benzene  is  +4.2 
kcal/mole  (heat  is  absorbed).  Assuming  the  in- 
crease in  randomness  is  the  same  when  iodine 
dissolves  in  liquid  benzene  as  it  is  in  ethyl  alco- 
hol and  in  CC14,  justify  the  prediction  that  the 
solubility  of  I2  in  benzene  is  higher  than  in  CC14 
but  lower  than  in  alcohol. 


10-1.4    Solubility  of  a  Gas  in  a  Liquid 

Gases,  too,  dissolve  in  liquids.  Let  us  apply  our 
understanding  of  equilibrium  to  this  type  of  sys- 
tem. 

THE    EFFECT   OF    RANDOMNESS 

The  gaseous  state  is  more  random  than  the  liquid 
state  since  the  molecules  move  freely  through  a 
much  larger  space  as  a  gas.  Hence  randomness 
decreases  as  a  gas  dissolves  in  a  liquid.  In  this 
case,  unlike  solids,  the  tendency  toward  maxi- 
mum randomness  favors  the  gas  phase  and  op- 
poses the  dissolving  process. 

THE    EFFECT    OF    ENERGY 

In  a  gas  the  molecules  are  far  apart  and  they 
interact  very  weakly.  As  a  gas  molecule  enters 


the  liquid,  it  comes  close  to  the  solvent  molecules 
and  they  attract  each  other,  lowering  the  poten- 
tial energy.  Again  we  find  a  contrast  to  the 
behavior  of  solids.  When  a  gas  dissolves  in  a 
liquid,  heat  is  evolved.  The  tendency  toward 
minimum  energy  favors  the  dissolving  process. 

Thus  we  see  that  the  equilibrium  solubility  of 
a  gas  again  involves  a  balance  between  random- 
ness and  energy  as  it  does  for  a  solid,  but  the 
effects  are  opposite.  For  a  gas,  the  tendency 
toward  maximum  randomness  favors  the  gas 
phase,  opposing  dissolving.  The  tendency  toward 
minimum  energy  favors  the  liquid  state,  hence 
favors  dissolving. 

As  an  example,  consider  the  solubilities  of  the 
two  gases,  oxygen,  02,  and  nitrous  oxide,  N20, 
in  water.  The  heats  of  solution  have  been  meas- 
ured and  are  as  follows: 

02(g)  +±  02(aq)  +  3.0  kcal/mole  02  (75) 

KiO(g)  +±l  H20(aq)  +  4.8  kcal/mole  N20    (14) 

Assuming  the  randomness  factor  is  about  the 
same,  the  gas  with  the  larger  heat  effect  (favoring 
dissolving)  should  have  the  higher  solubility.  The 
measured  solubilities  at  one  atmosphere  pressure 
and  20°C  of  oxygen  and  nitrous  oxide  in  water 
are,  respectively,  02,  1.4  X  10~3  mole/liter  and 
N20,  27  X  10~3  mole/liter,  consistent  with  our 
prediction. 

THE    EFFECT   OF   TEMPERATURE 

Raising  the  temperature  always  tends  to  favor 
the  more  random  state.  This  means  that  less  gas 
will  dissolve,  since  the  gas  is  more  random  than 
the  liquid.  The  solubility  of  a  gas  decreases  as 
temperature  is  raised. 


EXERCISE  10-2 

From  the  heat  of  solution  of  chlorine  in  water, 
—6.0  kcal/mole  (heat  evolved),  how  do  you  ex- 
pect the  solubility  of  chlorine  at  one  atmosphere 
pressure  and  20°C  to  compare  with  that  of  oxy- 
gen and  of  nitrous  oxide,  N20? 


168 


SOLUBILITY    EQUILIBRIA    I    CHAP.     10 


SOLIDS 


Tendency    to 
minimum    energy 


Solid 


Opposes    dissolving 


Te-n.de-n.cy  -to 
maximunt     randomness 


Favors     dissolving 

Raising    T  favors     solution- 
solubility    rises    as   T  rises. 


GASES 


s 


Gas 


Solution 


Favors    dissolving 

Fig.  10-3.  Maximum    randomness    versus    minimum 
energy — solubility  of  solids  and  gases. 


Solution 


Opposes   dissolving 

Raising    T  favors  gas : 
Solubility    drops    as   T  rises. 


10-2    AQUEOUS  SOLUTIONS 


Expression  (4)  is  applicable  to  the  solubility  equi- 
libria of  some  substances  in  water,  but  not  of  all. 
Contrast,  for  example,  water  solutions  of  sugar, 
salt,  and  hydrochloric  acid.  Sugar  forms  a  mo- 
lecular solid  and,  as  it  dissolves  in  water,  the 
sugar  molecules  remain  intact.  These  molecules 


leave  the  crystal  and  become  a  part  of  the  liquid. 
This  is  exactly  the  situation  we  described  for 
iodine  in  alcohol,  and  expression  (4)  is  applicable 
to  aqueous  sugar  solutions.  But  sodium  chloride, 
NaCl,  behaves  quite  differently.  As  salt  dissolves, 
positively  charged  sodium  ions  and  negatively 


SEC.    10-2    |    AQUEOUS    SOLUTIONS 


169 


charged  chloride  ions  enter  the  solution  and 
these  ions  behave  quite  independently: 

NaClfsj  — >-  Na+(aq)  +  Cl~(  aq)         (75) 

Hydrochloric  acid,  HC1,  is  similar.  This  sub- 
stance is  a  gas  at  normal  conditions.  At  very  low 
temperatures  it  condenses  to  a  molecular  solid. 
When  HC1  dissolves  in  water,  positively  charged 
hydrogen  ions  and  negatively  charged  chloride 
ions  are  found  in  the  solution.  As  with  sodium 
chloride,  a  conducting  solution  containing  ions 
is  formed: 


HCl(g)  — >-  H+(aq)  +  C\-(aq) 


(76) 


Substances  like  NaClfsj  and  HC\(g)  that  dis- 
solve in  water  to  form  conducting  solutions  are 
called  electrolytes.  The  conduction  involves 
movement  of  the  ions  through  the  solution,  posi- 
tive ions  moving  in  one  direction  and  negative 
ions  in  the  other.  This  shows  that  the  positive 
and  negative  ions  behave  independently.  In  view 
of  this  independence  of  the  ions,  solubility  be- 
havior in  an  electrolyte  solution  is  more  com- 
plicated than  that  given  by  expression  (4).  We 
shall  find  that  equilibrium  principles  are  corre- 
spondingly more  important. 

10-2.1    Types  off  Compounds  That 
Are  Electrolytes 

The  ions  in  an  electrolyte  solution  can  arise  in 
two  major  ways.  They  may  already  be  present 
in  the  pure  compound,  as  in  ionic  solids.  When 
such  a  solid  is  placed  in  water,  the  ions  separate 
and  move  throughout  the  solution.*  However, 
some  compounds  that  form  ions  in  water  are  not 
considered  to  contain  ions  when  pure,  whether  in 
the  solid,  liquid,  or  gas  phase.  Hydrochloric  acid, 
HC1,  and  sulfuric  acid,  H2S04,  are  good  exam- 
ples of  the  second  type  of  compound.  They  form 
molecular  liquids  (or  solids,  if  cold  enough).  But 
in  water  they  form  ions:  HC1  gives  hydrogen 
ion,  H+(aq),  and  chloride  ion,  C\-(aq);  H2S04 

*  Liquids  that  form  conducting  solutions  are  called 
ionizing  solvents.  A  few  other  compounds  (ammonia, 
NH3,  sulfur  dioxide,  S02,  sulfuric  acid.  H2SO,,  etc.)  are 
"ionizing  solvents"  but  water  is  by  far  the  most  impor- 
tant. We  will  discuss  water  exclusively  but  the  same  ideas 
apply  to  the  other  solvents  in  which  ions  form. 


gives  hydrogen  ion,  H+(aq),  hydrogen  sulfate  or 
bisulfate  ion,  HS04~,  and  sulfate  ion,  S04~2: 

HCl(g)  +  water  — >-  H+(aq)  +  C\-(aq)         (76) 

H2S04frJ  +  water  — *-  H+(aq)  +  HSCV (aq)    (77) 

HS04"  (aq)  ^±  H+(aq)  +  SOf  2faqj      (7S) 

However  they  are  formed,  and  from  whatever 
source,  aqueous  ions  are  individual  species  with 
properties  not  possessed  by  the  materials  from 
which  they  came.  Furthermore,  the  properties  of 
a  particular  kind  of  ion  are  independent  of  the 
source.  Chloride  ions  from  sodium  chloride, 
Na.C\(s),  have  the  same  properties  as  chloride 
ions  in  an  aqueous  solution  of  hydrochloric  acid, 
HC1.  In  a  mixture  of  the  two,  all  of  the  chloride 
ions  act  alike;  none  "remembers"  whether  it 
entered  the  solution  from  an  ionic  NaCl  lattice 
or  from  a  gaseous  HC1  molecule. 

Since  the  properties  of  an  ionic  solution  (that 
is,  a  solution  containing  ions)  differ  in  important 
ways  from  those  of  nonconducting  solutions,  it 
is  important  to  be  able  to  predict  which  sub- 
stances are  likely  to  form  ionic  solutions  in 
water.  The  periodic  table  guides  us. 

In  Chapter  6  we  saw  that  the  chemistry  of 
sodium  can  be  understood  in  terms  of  the  special 
stability  of  the  inert  gas  electron  population  of 
neon.  An  electron  can  be  pulled  away  from  a 
sodium  atom  relatively  easily  to  form  a  sodium 
ion,  Na+.  Chlorine,  on  the  other  hand,  readily 
accepts  an  electron  to  form  chloride  ion,  Cl~, 
achieving  the  inert  gas  population  of  argon. 
When  sodium  and  chlorine  react,  the  product, 
sodium  chloride,  is  an  ionic  solid,  made  up  of 
Na+  ions  and  Cl~  ions  packed  in  a  regular  lat- 
tice. Sodium  chloride  dissolves  in  water  to  give 
Na+(aq)  and  C\~(aq)  ions.  Sodium  chloride  is 
an  electrolyte;  it  forms  a  conducting  solution  in 
water. 

This  example  illustrates  the  guiding  principles. 
Sodium  is  a  metal — electrons  can  be  pulled  away 
from  sodium  relatively  easily  to  form  positive 
ions.  Chlorine  is  a  nonmetal — it  tends  to  accept 
electrons  readily  to  form  negative  ions.  When  a 
metallic  element  reacts  with  a  nonmetallic  ele- 
ment, the  resulting  compound  usually  forms  a 
conducting  solution  when  dissolved  in  water. 


170 


SOLUBILITY    EQUILIBRIA    |    CHAP.    10 


H* 

Positive  ions     -t) 

tair  form,    s-oluble 
■A  ALL   anions 

compounds  yvii 

He 

Li  + 

Be*2 

B 

co;2 

/no; 

OH' 

F~ 

Ne 

Not 

M9» 

At'3 

Si 

po;3 

/£o+ 

cr 

Ar 

K  + 

Ca'2 

5c'3 

Ti 

V 

Cr+3 

Mrt2 

Fe'2/ 

Co+2 

Ni'2 

Cut/ 

/ttt2 

Zn+Z 

Oct* 

Ge 

As*3/ 

4** 

Se 

Br~ 

Kr 

Rb+ 

sS2 

Y+3 

Zr 

Nb 

Mo 

Tc 

Ru 

Rh 

Pd 

A$+ 

Cd'z 

if3 

St/ 
/t+ 

5b'3/ 
/h's 

Te 

r 

Xe 

Cs+ 

Bet2 

It3 

&R.E. 

Hf 

Ta 

W 

Re 

Os 

Ir 

Pt 

Au 

H<3z7 
/Hg 

n  + 

Pb'2 

Bi+3 

Po 

At~ 

Rn 

Fr+ 

Rt2 

The  metals  are  found  toward  the  left  side  of  the 
periodic  table  and  the  nonmetals  are  at  the  right 
side.  A  compound  containing  elements  from  the 
opposite  sides  of  the  periodic  table  can  be  expected 
to  form  a  conducting  solution  when  dissolved  in 
water.  Notice  from  our  examples  that  hydrogen 
reacts  with  nonmetals  to  form  compounds  that 
give  conducting  solutions  in  water.  In  this  sense, 
hydrogen  acts  like  a  metallic  element. 


EXERCISE  10-3 

Using  the  periodic  table  as  a  guide,  predict  which 
of  the  following  compounds  form  ionic  solutions 
in  water:  silicon  carbide,  SiC;  magnesium  bro- 
mide, MgBr2;  carbon  tetrabromide,  CBr4;  chro- 
mic chloride,  CrCl3. 


10-2.2    A  Qualitative  View  of  Aqueous 
Solubilities 

Hereafter  in  this  chapter  we  shall  be  concerned 
exclusively  with  substances  that  form  ionic  solu- 
tions in  water.  Since  each  substance  is  electrically 
neutral  before  it  dissolves,  it  must  form  ions  of 
positive  charge  and,  as  well,  ions  of  negative 
charge.  Ions  with  positive  charges  are  called 
cations.  Ions  with  negative  charges  are  called 
anions.  A  conducting  solution  is  electrically 
neutral;  it  contains  both  anions  and  cations. 

First,  let  us  consider  substances  with  high  solu- 
bility. As  was  stated  on  p.  73,  chemists  consider 


Fig.  10-4.  Almost  all  compounds  of  the  alkalies,  hy- 
drogen ion,  and  ammonium  ion  are  soluble 
in  water. 


a  substance  to  be  soluble  if  it  dissolves  to  a 
concentration  exceeding  one-tenth  of  a  mole  per 
liter  (0.1  M)  at  room  temperature.  Using  this 
meaning  of  the  word  soluble,  we  can  say  that 
some  cations  (positive  ions)  form  soluble  com- 
pounds with  all  anions  (negative  ions).  These 
cations  are  the  hydrogen  ion,  H+(aq),  ammo- 
nium ion,  NH^,  and  the  alkali  ions,  Li+,  Na+, 
K+,  Rb+,  Cs+,  Fr+.  Figure  10-4  shows  the  place- 
ment of  these  ions  in  the  periodic  table. 

The  same  sort  of  remark  can  be  made  about 
two  anions  (negative  ions).  Almost  all  com- 
pounds involving  nitrate  ion,  N03_,  and  acetate 
ion,  CH3COO-,  are  soluble  in  water.* 

Other  anions  (negative  ions)  form  compounds 
of  high  solubility  in  water  with  some  metal 
cations  (positive  ions)  and  compounds  of  low 
solubilities  with  others.  Figure  10-5  indicates,  for 
five  anions,  the  metal  ions  that  form  compounds 
of  low  solubilities.  Figure  10-5 A  refers  to  chlo- 
rides, bromides,  and  iodides,  CI-,  Br-,  and  I-; 
Figure  10-5B  refers  to  sulfates,  S04-2;  Figure 
10-5C  refers  to  sulfides,  S-2.  Notice  the  difference 
between  Figure  10-4  and  Figure  10-5.  The  cross- 


*  There  are  a  few  compounds  of  alkalies,  nitrate,  and 
acetate  that  have  low  solubilities,  but  most  of  them  are 
quite  complex  in  composition.  For  example,  sodium 
uranyl  acetate,  NaU02(CH3COO)3  has  low  solubility. 
Silver  acetate  is  an  exception  but  its  solubility  is  moderate. 


SEC.    10-2    I    AQUEOUS    SOLUTIONS 


171 


hatching  in  Figure  10-4  identifies  metal  ions  that 
form  soluble  compounds.  Figure  10-5  identifies 
the  ones  with  low  solubility. 

Figure  10-6  continues  this  pictorial  presenta- 
tion of  solubilities.  Figure  10-6A  shows  the  posi- 
tive ions  that  form  hydroxides  of  low  solubility. 
Figure  10-6B  shows  the  positive  ions  that  have 
low  solubility  when  combined  with  phosphate 
ion,  P(V3,  carbonate  ion,  C03-2,  and  sulfite  ion, 

so3-2. 

These  figures  furnish  a  handy  summary  of 
solubility  behavior.  We  see  from  Figure  10-5 A 
that  few  chlorides  have  low  solubilities.  The  few 
that  do  contain  cations  of  metals  clustered  to- 
ward the  right  side  of  the  periodic  table  (silver 
ion,  Ag+,  cuprous  ion,  Cu+,  mercurous  ion, 
Hg2+2,  and  lead  ion,  Pb+2)  but  they  do  not  fall 
in  a  single  column.  This  irregularity  is  not  un- 


usual in  solubility  behavior  and  is  seen  again  in 
Figures  10-5B  and  10-6A.  In  these  two  figures, 
the  elements  in  the  second  column  (the  alkaline 
earths)  show  a  trend  in  behavior.  Thus,  beryllium 
and  magnesium  ions  (Be+2  and  Mg+2)  form  solu- 
ble sulfates.  The  others — calcium,  strontium, 
barium,  and  radium  ions  (Ca+2,  Sr+2,  Ba+2,  and 
Ra+2) — form  sulfates  with  low  solubilities.  Just 
the  opposite  behavior  is  seen  in  Figure  10-6A  for 
the  compounds  of  these  same  elements  with  hy- 
droxide ion,  OH-.  As  for  the  elements  in  the 
middle  of  the  periodic  table,  we  see  that  they 
form  compounds  of  low  solubilities  with  the 
ions:  sulfide,  S-2,  hydroxide,  OH~,  phosphate, 
P04-3,  carbonate,  C03"2,  and  sulfite,  S03-2. 

The  information  contained  in  Figures  10-4, 
10-5,  and  10-6  is  collected  for  reference  in  Table 
10-1. 


Table  10-1.     solubility  of  common  compounds  in  water 

+ 


NEGATIVE  IONS 

(Anions) 


POSITIVE  IONS 

(Cations) 


FORM 


compounds  with 
solubility: 


All 

Alkali  ions 

Li+,  Na+,  K+,  Rb+,  Cs+,  Fr+ 

Soluble 

All 

Hydrogen  ion,  H+(aq) 

Soluble 

All 

Ammonium  ion,  NH4+ 

Soluble 

Nitrate,  N03_ 

All 

Soluble 

Acetate,  CH3COO" 

All 

Soluble 

Chloride,  CI"] 
Bromide,  Br-  > 
Iodide,  I"       J 

Ag+,  Pb+2,  Hg2+2,  Cu+ 
All  others 

Low  Solubility 
Soluble 

Sulfate,  S04-* 

Ag\  Ca+2,  Sr+2,  Ba+2,  Pb+2 
All  others 

Low  Solubility 
Soluble 

Sulfide,  S-* 

Alkali  ions,  H+(aq),  NH4\  Be+2, 

Mg+2,  Ca+2,  Sr+2,  Ba« 
All  others 

Soluble 
Soluble 

Low  Solubility 

Hydroxide,  OH" 

Alkali  ions,  H+(aq),  NH«\  Sr+2,  Ba"" 
All  others 

Soluble 

Low  Solubility 

Phosphate,  PCV8  i 
Carbonate,  COa  2 
Sulfite,  SOr2        J 

Alkali  ions,  H+(aq),  NH4+ 
All  others 

Soluble 

Low  Solubility 

172 


SOLUBILITY    EQUILIBRIA    I    CHAP.     10 


H* 

A 

He 

Li* 

Be*2 

B 

cor2 

vuy 

/no; 

OH' 

F~ 

Ne 

with 

CIZ  Br-T  I'\ 

Na* 

Mo*2 

Al*3 

Si 

po;3 

s-y 

/so* 

cr 

Br~ 
I" 

Ar 

K* 

Ca*2 

5c*3 

Ti 

V 

Cr*3 

Mn*2 

Fe+V 

/Pe+i 

Co*2 

*-«k 

Ztt2 

OS3 

G-e 

As*3/ 

A*5 

Se 

Kr 

Rb* 

Sr*2 

Y+3 

Zr 

Nb 

Mo 

Tc 

Ru 

Rh 

Pd    1        1  Cd*2 

In*3 

5n  y 

/  +4 
/on 

sb*y 
/b*s 

Te 

Xe 

Cs* 

Ba*2 

La*3 
&R.E. 

Hf 

To. 

W 

Re 

Os 

lr 

Pt 

Y^9 

ti*\    9 Br3 

Po 

At~ 

Rn 

Fr* 

Ra*2 

H* 

B 

He 

Positive  ions 

1    -that  form. 

Li* 

Be*2 

compounds  oJ~  jLUW  soLtLbilit-y 
with  p^S 

B 

CO'/ 

/no; 

OH' 

F~ 

Ne 

Na* 

M9*2 

Al*3 

Si 

po;3 

s'2A 

cr 

Ar 

- 

Ca*2 
Sr*2 
Ba*2 

5c*3 

Ti 

V 

Cr*3 

Mn*2 

Fe*2/ 
/Fe+i 

Co*2 

Nt*2 

Cu/ 
/tJ2 

Zn*2 

aa*3 

Ge 

As*3/ 
/s+S 

Se 

Br~ 

Kr 

Rb* 

y+3 

Zr 

Nb 

Mo 

Tc 

Ru 

Rh 

Pd 

A«+ 

Cd*Z 

In*3 

Sn*V 

sb*y 
/b*s 

Te 

r 

Xe 

~ 

la*3 
&R.E. 

Hf 

Ta 

vV 

Re 

Os 

lr 

Pt 

Au 

Hsiy 

M1 

Tl*\      \Bi*3 

Po 

At~ 

Rn 

Fr* 

Ra*2 

H  + 

c 

■thxiir  form 

He 

1  Positive  ions 

Li* 

Be*2 

compotinas  oj~  jluw    solubility 
with   |  Q 

B 

CO'/ 

vuy 

/no; 

OH' 

F- 

Ne 

Na* 

Mo*2 

M"m 

po;3 

1 

cr 

Ar 

K* 

Ca*2\ 

Sc*3    Ti        V      Cr*3  Mn*2  **1*3  Co*2  Ni+2  ** \,  Z*t+Z 

Fe*3                          Cu2 

Ga+3   &e    A**J+s 
As5 

Se 

Br~ 

Kr 

Rb* 

Sr*2] 

Rh      Pd     Ag*   Cd*2 

T  +3  Sn*2     5b*3 
ln        rS+    si*5 

Te 

r 

Xe 

Cs* 

Ba+2% 

Wk 

TI*  Pb*2  Bi*3 

Po 

At~ 

Rn 

Fr* 

Ra*2 

Fig.  10-5.  Positive  ions  forming  compounds  of  low  solubilities  with  various  anions. 


SEC.     10-2    |    AQUEOUS    SOLUTIONS 


173 


H* 

Lim 

Na+ti 

K*\ 

sS3 

Rb  + 

Sr+Z 

y+3 

Cs* 

Bct2\ 

La*3 

Fr  + 

Ra*2 

Positive  ions 


that  form 
compounds  of  LOW  solubility 
with 


hydroxtde  ton,    OH 


+3M*?2Fe 


,*2   Ni*2  Cu"„  Zn*2   Ga*3 


Mo      Tc      Ru     Rh      Pd     Ao*    Cd*z  In 


Jr       Pt      At 


Zn~   Go"    Ge 


*Z    t+3  Sn 


H* 

Pi 

£J 

t  form 

He 

H  Positive  ions   | 

Lt  + 

Be*2\ 

1  PO/3,   CO/2,  and  SO/2 ■ 

B 

CO/2 

my 
/do; 

OH' 

F~ 

Ne 

No* 

m  1*3 

cr 

Ar 

Mo*2m 

At 

Si 

po4j 

so/ 

K* 

Ca*2 

SS> 

V      Cr      Mn 

hmUmI  ^■■t 

Ge 

As*3 

A/s 

Se 

Br~ 

Kr 

Rb  + 

C  +2 

*r+3 

Nb     Mo       Tc 

i 

In*3 

Sn*2 
in*4 

Sb*3 
Sb*5 

Te 

r 

Xe 

or         i 

" 

tin       ra.      j\y       t-a 

Cs* 

**z  til  v 

w  +2 
Ta       W      Re       Os       Ir       Pt      Au    "**  +2 

Eg 

TV* 

Pb*z 

Bi*3 

Po 

At" 

Rn 

Fr* 

Ra+Z 

Fig.  10-6.  More  positive  ions  forming  compounds  of 
low  solubilities  with  various  anions. 


EXERCISE  10-4 

Use  Figures  10-4,  10-5,  and  10-6  to  decide  the 
solubility  of  each  of  the  compounds  listed  below. 
Write  "sol"  if  the  compound  is  soluble  and 
"low"  if  it  has  low  solubility. 

Mg(N03)2,    MgCl2,    MgS04,    Mg(OH)2,    MgC03, 

Ca(N03)2,     CaCl*     CaS04,     Ca(OH)2,     CaC03, 

Sr(N03)2,      SrCl2,      SrS04,      Sr(OH)2,      SrC03. 

EXERCISE  10-5 

Decide  the  formula  of  each  of  the  following  com- 
pounds and  indicate  those  with  low  solubility  in 


water.  Silver  carbonate;  aluminum  chloride; 
aluminum  hydroxide;  cuprous  chloride  (the 
chloride  of  Cu+);  cupric  chloride  (the  chloride 
of  Cu+2);  ammonium  bromide. 


10-2.3    The  Equilibrium  Law 

Table  10-1  and  the  schematic  presentations  of 
Figures  10-4,  10-5,  and  10-6  are  useful  for  a 
quick  and  qualitative  view  of  solubilities.  But 
chemists  are  not  satisfied  with  the  statement  that 
a  substance  has  low  solubility.  We  must  know 
how  much  of  the  substance  dissolves.  We  must  be 
able  to  treat  solubility  in  a  quantitative  fashion. 
Fortunately  this  can  be  done  with  the  aid  of  the 


174 


SOLUBILITY    EQUILIBRIA    I    CHAP.     10 


principles  of  equilibrium,  developed  in  Chap- 
ter 9. 

As  mentioned  earlier,  the  quantitative  concen- 
tration relationship  that  exists  at  equilibrium  is 
shown  in  the  Equilibrium  Law  Relation: 


EXERCISE  10-6 


K  = 


\EV\F]> 


WW 


(2) 


Expression  (2)  applies  to  a  solubility  equilibrium, 
provided  we  write  the  chemical  reaction  to  show 
the  important  molecular  species  present.  In  Sec- 
tion 10-1  we  considered  the  solubility  of  iodine 
in  alcohol.  Since  iodine  dissolves  to  give  a  solu- 
tion containing  molecules  of  iodine,  the  con- 
centration of  iodine  itself  fixed  the  solubility. 
The  situation  is  quite  different  for  substances 
that  dissolve  to  form  ions.  When  silver  chloride 
dissolves  in  water,  no  molecules  of  silver  chlo- 
ride, AgCl,  seem  to  be  present.  Instead,  silver 
ions,  Ag+,  and  chloride  ions,  Cl~,  are  found  in 
the  solution.  The  concentrations  of  these  species, 
Ag+  and  CI-,  are  the  ones  which  fix  the  equilib- 
rium solubility.  The  counterpart  of  equation  (7) 
will  be 

AgC\(s)  q=b  Ag+(aq)  +  Or(aq)  (19) 

and  equilibrium  will  exist  when  the  concentra- 
tions are  in  agreement  with  the  expression 


K  =  a  constant  =  [Ag+]  X  [CI-] 


(20) 


Just  as  in  expression  (4),  the  "concentration"  of 
the  solid  (silver  chloride)  does  not  appear  in  the 
equilibrium  expression  (20);  it  does  not  vary. 
To  consider  a  more  complicated  example,  con- 
sider the  application  of  expression  (2)  to  the 
solubility  of  lead  chloride,  PbCl2: 


PbC\2(s) 


Pb+*(aq)  +  2Cl-(aq)        (21) 


At  equilibrium, 

K  =  a  constant  =  [Pb+2]  X  [CI"]2        (22) 

Solubility  equilibrium  constants,  such  as  (20) 
and  (22),  are  given  a  special  name — the  solubility 
product.  It  is  symbolized  Ksp.  A  low  value  of 
Ksp  means  the  concentrations  of  ions  are  low  at 
equilibrium.  Hence  the  solubility  must  be  low. 
Table  10-11  lists  solubility  products  for  some 
common  compounds. 


Write  the  equation  for  the  dissolving  of  calcium 
sulfate,  CaS04,  and  the  solubility  product  ex- 
pression. 


EXERCISE  10-7 


Write  the  equation  for  the  dissolving  of  silver 
chromate,  Ag2Cr04,  and  the  solubility  product 
expression.  Silver  chromate  dissolves  to  give  Ag+ 
and  Cr04~2  ions. 


10-2.4    Calculation  off  the  Solubility 
of  Cuprous  Chloride  in  Water 

The  solubility  product  is  learned  from  measure- 
ments of  the  solubility.  In  turn,  it  can  be  used  as 
a  basis  for  calculations  of  solubility.  Suppose  we 
wish  to  know  how  much  cuprous  chloride,  CuCl, 
will  dissolve  in  one  liter  of  water.  We  begin  by 
writing  the  balanced  equation  for  the  reaction: 

CuClfsj  +±:  Cu+(aq)  +  C\~(aq)         (23) 

From  this  equation,  we  can  write  the  equilibrium 
expression: 

K,p  =  [Cu+][C1-]  (24) 

Now  the  numerical  value  of  Ktp  is  found  in  Table 
10-11: 

K,p  =  3.2  X  10-7  -  [Cu+][C1-]  (25) 

Table  10-11 

SOME    SOLUBILITY    PRODUCTS    AT 
ROOM    TEMPERATURE 


TIC1 

1.9  X  10-4 

SrCr04 

3.6  X  10-' 

CuCl 

3.2  X  10-7 

BaCr04 

8.5  X  10-" 

AgCl 

1.7  X  10-'° 

PbCr04 

2     X  10-18 

TIBr 

3.6  X  10-6 

CaS04 

2.4  X  lO"4 

CuBr 

5.9  X  10-9 

SrSO« 

7.6  X  10"7 

AgBr 

5.0  X  10~18 

PbSO, 

1.3  X  10-» 

BaS04 

1.5  X  lO"9 

Til 

8.9  X  10-8 

RaSO« 

4     X  lO"11 

Cul 

1.1  X  io-« 

Agl 

8.5  X  10-17 

AgBr03 

5.4  X  10"8 

AglOj 

3.1  X  lO"8 

SEC.    10-2    I   AQUEOUS    SOLUTIONS 


175 


Expression  (25)  indicates  that  cuprous  chloride 
dissolves,  according  to  reaction  (23),  until  the 
molar  concentrations  of  cuprous  ion  and  chlo- 
ride ion  rise  enough  to  make  their  product  equal 
to  3.2  X  10"7. 

Now  is  the  time  to  dust  off  the  algebra  and  put 
it  to  work.  Suppose  we  designate  the  solubility 
of  cuprous  chloride  in  water  by  a  symbol,  s.  This 
symbol  s  equals  the  number  of  moles  of  solid 
cuprous  chloride  that  dissolve  in  one  liter  of 
water.  Remembering  equation  (23),  we  see  that  s 
moles  of  solid  cuprous  chloride  will  produce  s 
moles  of  cuprous  ion,  Cu+,  and  s  moles  of  chlo- 
ride ion,  CI-.  Hence  these  concentrations  must 
be  equal,  as  shown  below. 

[Cu+]  =  [CI-]  =  s  moles/liter 

=  moles  CuCl  dissolved  (26) 

Substituting  (26)  into  (25),  we  have 

K,p  =  3.2  X  10-7  =  (s)  X  (s)  =  a» 

s1  =  3.2  X  10"7  =  32  X  10-8 

s  =  V32  X  10-8  =  5.7  X  10-<  mole/liter 

=  0.00057  M  (27) 

EXERCISE  10-8 

Calculate  the  solubility,  in  moles  per  liter,  of 
calcium  sulfate  in  water,  using  the  solubility 
product  given  in  Table  10-11. 


10-2.5    Will  a  Precipitate  Form? 

When  two  solutions  are  mixed,  a  precipitate  may 
form.  For  example,  suppose  solutions  of  calcium 
chloride,  CaCl2,  and  sodium  sulfate,  Na2S04,  are 
mixed.  The  mixture  contains  both  calcium  ions, 
Ca+2,  and  sulfate  ions,  S04-2,  so  solid  calcium 
sulfate  may  form.  The  solubility  product  per- 
mits us  to  predict  with  confidence  whether  it  will 
or  not. 

Let  us  consider  two  cases  to  show  '  ow  the 
prediction  is  made: 

(1)  Equal  volumes  of  0.02  M  CaCl2  and  0.0004  M 
Na2S04  are  mixed. 

(2)  Equal  volumes  of  0.08  M  CaCl2  and  0.02  M 
Na2S04  are  mixed. 


Fig.  10-7.  Will  a  precipitate  form? 


Will  a  precipitate  form  in  either  case?  In  both? 

The  first  step  is  to  write  the  balanced  equation 
for  the  reaction  of  calcium  sulfate  dissolving  in 
water  and  then  use  the  Equilibrium  Law: 

CaSO/sj  +±l  Ca+i(aq)  +  SQZ*(aq)      (28) 

K.p  =  [Ca+2][S04-2]  (29) 

The  next  step  is  to  find  the  concentration  of  each 
ion  in  the  final  mixture.  After  we  mix  equal  vol- 
umes, each  ion  is  present  in  twice  as  much  solu- 
tion so  the  concentration  is  only  half  as  great  as 
before  mixing.  Therefore,  in  case  1 , 

[Ca+2]  =  0.02M  =  Q  Q1  Msslx  10-2  M 


[SCV1]  = 


0.0004  M 


=  0.0002  M  =  2  X  10-«  M 


A  trial  value  of  the  ion  product  must  be  com- 
pared to  K,v\ 

[Ca+J]  X  [StV2]  =  (1  X  10-')  X  (2  X  10"<) 
=  2  X  10"« 

The  trial  product,  2  X  10-6,  is  less  than  K,v  = 
2.4  X  10-4  so  precipitation  will  not  occur  in  the 
mixture  of  case  1.  In  case  2, 


176 


SOLUBILITY    EQUILIBRIA    |    CHAP.     10 


[Ca*»]  =  ™±M  =  0.04  M 


4  X  10~2  M 


[SCV2]  =  °-02M  =  0.01  M  =  \  X  10~2  M 

Again  we  calculate  a  trial  value  of  the  ion 
product, 

[Ca+2]  X  [SOf2]  =  (4  X  10-2)  X  (1  X  10~2) 
=  4  X  10-" 

This  time  the  trial  product,  4  X  10-4,  is  greater 
than  Ksp  =  2.4  X  10-5  so  a  precipitate  does  form. 
Solid  calcium  sulfate,  CaS04,  will  continue  to 
form,  lowering  the  concentrations  [Ca+2]  and 
[S04-2]  until  they  are  low  enough  that  the  ion 
product  equals  Ksp.  Then  equilibrium  exists  and 
no  more  precipitation  occurs. 

EXERCISE  10-9 

A  50  ml  volume  of  0.04  M  Ca(N03)2  solution  is 
added  to  150  ml  of  0.008  M  (NH4)2S04  solution. 
Show  that  a  trial  value  of  the  calcium  sulfate  ion 
product  is  6  X  10~5.  Will  a  precipitate  form? 


10-2.6    Precipitations  Used  for  Separations 

A  chemist  is  often  interested  in  separating  sub- 
stances in  a  solution  mixture.  Such  a  problem  is 


solved  by  applying  equilibrium  considerations. 
Suppose  we  have  a  solution  known  to  contain 
both  lead  nitrate,  Pb(N03)2,  and  magnesium 
nitrate,  Mg(N03)2.  The  lead  and  magnesium  can 
be  separated  by  removing  from  the  solution  all 
of  the  lead  ion,  Pb+2,  as  a  solid  lead  compound. 
We  must  avoid,  of  course,  precipitation  of  any 
magnesium  compound.  Consulting  Figure  10-5 
or  Table  10-1,  we  see  that  lead  ion  and  sulfate 
ion  form  a  compound  with  low  solubility.  If 
enough  sodium  sulfate,  Na2S04,  is  added,  lead 
sulfate,  PbS04,  will  precipitate.  Since  Figure 
10-5B  and  Table  10-1  indicate  that  magnesium 
sulfate,  MgS04,  is  soluble,  there  will  be  no  pre- 
cipitate of  MgS04.  The  solid  can  be  removed 
from  the  liquid  by  filtration  and  the  desired 
separation  has  been  obtained. 


EXERCISE  10-10 

Use  Figures  10-5  and  10-6  or  Table  10-1  to  decide 
which  of  the  following  soluble  salts  would  permit 
a  separation  of  magnesium  and  lead  through  a 
precipitation  reaction:  sodium  iodide,  Nal;  so- 
dium sulfide,  Na2S;  sodium  carbonate,  Na2C03. 


QUESTIONS  AND  PROBLEMS 


1 .  Sugar  is  added  to  a  cup  of  coffee  until  no  more 
sugar  will  dissolve.  Does  addition  of  another 
spoonful  of  sugar  increase  the  rate  at  which  the 
sugar  molecules  leave  the  crystal  phase  and  enter 
the  liquid  phase?  Will  the  sweetness  of  the  liquid 
be  increased  by  this  addition?  Explain. 

2.  In  view  of  the  discussion  of  the  factors  that  de- 
termine the  rate  of  dissolving  (Section  10-1.2), 
propose  two  methods  for  increasing  the  rate  at 
which  sugar  dissolves  in  water. 

3.  When  a  solid  evaporates  directly  (without  melt- 
ing), the  process  is  called  sublimation.  Evapo- 
ration of  "dry  ice"  (solid  C02)  is  a  familiar 
example.  Two  other  substances  that  sublime  are 
FCN  and  ICN: 


FCN(s)  +±  FCN(g)        AH  =  +5.7  kcal 
lCN(s)  +±  lCN(g)         AH  =  +14.2  kcal 

(a)  In  sublimation,  does  the  tendency  toward 
maximum  randomness  favor  solid  or  gas? 

(b)  In  sublimation,  does  the  tendency  toward 
minimum  energy  favor  solid  or  gas? 

(c)  The  vapor  pressure  of  solid  FCN  is  760  mm 
at  201°K.  In  view  of  part  b,  would  you  expect 
solid  ICN  to  have  a  lower  or  higher  vapor 
pressure  than  solid  FCN  at  this  same  tem- 
perature, 201  °K? 

4.  Liquid  chloroform,  CHC13,  and  liquid  acetone, 
CH3COCH3,  dissolve  in  each  other  in  all  pro- 
portions (they  are  said  to  be  miscible). 
(a)  When  pure  CHC13  is  mixed  with  pure  ace- 
tone, is  randomness  increased  or  decreased  ? 


QUESTIONS    AND    PROBLEMS 


177 


.(b)  Does  the  tendency  toward  maximum  ran- 
domness favor  reactants  or  product  in  the 
reaction : 


CHCh(l)  +  CH3COCH3(l) 


>-  1 : 1  solution 
AH  =  -495  cal 


(c)  Considering  the  sign  of  AH  shown  in  part  b, 
does  the  tendency  toward  minimum  energy 
favor  reactants  or  product? 

(d)  In  view  of  your  answers  to  parts  b  and  c, 
discuss  the  experimental  fact  that  these  two 
liquids  are  miscible. 

5.  Which  of  the  following  substances  can  be  ex- 
pected to  dissolve  in  the  indicated  solvent  to 
form,  primarily,  ions  ?  Which  would  form  mole- 
cules? 


(a)  sucrose  in  water. 

(b)  RbBr  in  water. 

(c)  CHC13  in  water. 

(d)  CsN03  in  water. 


(e)  HNO3  in  water. 

(f)  S8  in  carbon  disul- 
fide, CS2. 

(g)  IC1  in  ethyl  alcohol. 


6.  Which  of  the  substances  listed  in  Problem  5 
would  be  called  electrolytes? 

7.  Assume  the  following  compounds  dissolve  in 
water  to  form  separate,  mobile  ions  in  solution. 
Write  the  formulas  and  names  for  the  ions  that 
can  be  expected. 


(a)  HI 

(b)  CaCl2 

(c)  Na2C03 


(d)  Ba(OH)2 

(e)  KNO3 

(f)  NH4C1 


8.  Write  the  equation  for  the  reaction  that  occurs 
when  each  of  the  following  electrolytes  is  dis- 
solved in  water. 

(a)  lithium  hydroxide  (solid). 

(b)  nitric  acid  (liquid). 

(c)  potassium  sulfate  (solid). 

(d)  sodium  nitrate  (solid). 

(e)  ammonium  iodide  (solid). 

(f)  potassium  carbonate  (solid). 

Answer,  (a)  LiOHfsJ  — >■  U+(aq)  +  OH~  (aq). 

9.  What  would  you  expect  to  happen  if  equal  vol- 
umes of  0.1  M  MgS04  and  0.1  M  ZnCl2  were 
mixed  together? 

10.  Predict  what  would  happen  if  equal  volumes  of 
0.2  M  Na2S03  and  0.2  M  MgS04  were  mixed.  If 
a  reaction  takes  place,  write  the  net  ionic  equa- 
tion. 


11.  Using  Figures  10-4  to  10-6  (or  Table  10-1),  make 
a  statement  about  the  solubilities  of  the  com- 
pounds containing  the  following  ions. 


Anion 

(a)  carbonate,  C03~ 2 

(b)  carbonate,  C03" 2 

(c)  sulfide,  S"2 

(d)  hydroxide,  OH~ 

(e)  chloride,  Cl" 


Cations 
alkali    ions,    Li+,    Na+, 

K+,  Rb+,  Cs+ 
alkaline  earth  ions,  Be+2, 

Mg+2,  Ca+2,  Sr+2,  Ba+2 
alkaline  earth  ions,  Be+2, 

Mg+2,  Ca+2,  Sr+2,  Ba+2 
the  cations  of  the  fourth 

row  of  the  periodic 

table 
the  cations  of  the  fifth 

row   of  the   periodic 

table 


12. 


Answer,  (a)  All  alkali  carbonates  are  soluble, 
(b)  All    alkaline    earth    carbonates 
have  low  solubilities. 

Write  the  empirical  formulas  for  each  of  the 
following  compounds  and  indicate  which  have 
low  solubilities. 

(a)  silver  sulfide. 

(b)  potassium  sulfide. 

(c)  ammonium  sulfide. 

(d)  nickel  sulfide. 

(e)  ferrous  sulfide  (Fe+2). 

(f)  ferric  sulfide  (Fe+3). 


13.  Write  net  ionic  equations  for  any  reactions  that 
will  occur  upon  mixing  equal  volumes  of  0.2  M 
solutions  of  the  following  pairs  of  compounds. 

(a)  silver  nitrate  and  ammonium  bromide. 

(b)  SrBr2  and  NaN03. 

(c)  sodium  hydroxide  and  aluminum  chloride. 

(d)  Nal  and  Pb(N03)2. 

(e)  barium  chloride  and  sodium  sulfate. 

Answer,  (a)  Ag+(aq)  +  Br~(aq)  — >-  AgBrfsJ. 

14.  What  ions  could  be  present  in  a  solution  if 
samples  of  it  gave: 

(a)  A    precipitate    when    either    C\~(aq)    or 
S04"%qj  is  added? 

(b)  A  precipitate  when  C\~(aq)  is  added  but 
none  when  S04~  (aq)  is  added? 

(c)  A  precipitate  when  SO4  2(aq)  is  added  but 
none  when  Cl~ (aq)  is  added? 


178 


SOLUBILITY    EQUILIBRIA    I    CHAP.     10 


15.  What  cations  from  the  fourth  row  of  the  periodic 
table  could  be  present  in  a  solution  with  the 
following  behavior. 

(a)  No  precipitate  is  formed  with  hydroxide  ion. 

(b)  A  precipitate  forms  with  hydroxide  ion  and 
with  sulfate  ion. 

(c)  A  precipitate  forms  with  hydroxide  ion  and 
with  sulfide  ion. 

(d)  A  precipitate  forms  with  carbonate  ion,  none 
with  sulfide  ion. 

16.  The  solubility  of  silver  chloride  is  so  low  that  all 
but  a  negligible  amount  of  it  is  precipitated  when 
excess  sodium  chloride  solution  is  added  to  silver 
nitrate  solution.  What  would  be  the  weight  of 
the  precipitate  formed  when  100  ml  of  0.5  M 
NaCl  is  added  to  50.0  ml  of  0.100  M  AgNOs? 

Answer.  0.715  gram. 

17.  Write  the  solubility  product  expression  for  each 
of  the  following  reactions. 


(a)  BaS04fsJ  *=■ 

(b)  Zn(OH)/S/)  : 

(c)  Ca3(P04)2(S; 


Ba+S (aq)  +  SOr(aq) 
±  Zn+2(aq)  +  20H-(aq) 
=±  3Ca+2f aq)  +  2P04"3(aqJ 


18.  Write  the  solubility  product  expression  applica- 
ble to  the  solubility  of  each  of  the  following 
substances  in  water. 

(a)  calcium  carbonate. 

(b)  silver  sulfide. 

(c)  aluminum  hydroxide. 

19.  The  solubility  product  of  AgCl  is  1.4  X  10~4  at 
100°C.  Calculate  the  solubility  of  silver  chloride 
in  boiling  water. 

20.  Experiments  show  that  0.0059  gram  of  SrC03 
will  dissolve  in  1.0  liter  of  water  at  25°C.  What 
is  K,p  for  SrCOa? 

Answer.  1.6  X  10"9. 


21.  How  many  milligrams  of  silver  bromide  dissolve 
in  20  liters  of  water?  (Use  the  data  given  in 
Table  10-11.) 

22.  To  one  liter  of  0.001  M  H2S04  is  added  0.002 
mole  of  solid  Pb(N03)2.  As  the  lead  nitrate  dis- 
solves, will  lead  sulfate  precipitate? 

23.  Suppose  10  ml  of  1.0  M  AgN03  is  diluted  to  one 
liter  with  tap  water.  If  the  chloride  concentration 
in  the  tap  water  is  about  10~6  M,  will  a  precipi- 
tate form? 

24.  The  test  described  in  Problem  23  does  not  give 
a  precipitate  if  the  laboratory  distilled  water  is 
used.  What  is  the  maximum  chloride  concentra- 
tion that  could  be  present  ? 

25.  Will  a  precipitate  exist  at  equilibrium  if  \  liter 
of  a  2  X  10~3  M  A1C13  solution  and  \  liter  of  a 
4  X  10-2  M  solution  of  sodium  hydroxide  are 
mixed  and  diluted  to  103  liters  with  water  at 
room  temperature?  (Ksp  =  5  X  10~33.) 

26.  Use  Figures  10-5  and  10-6  or  Table  10-1  to  decide 
which  of  the  following  soluble  substances  would 
permit  a  separation  of  aqueous  magnesium  and 
barium  ions.  For  those  that  are  effective,  write 
the  equation  for  the  reaction  that  occurs. 

(a)  ammonium  carbonate. 

(b)  sodium  bromide. 

(c)  potassium  sulfate. 

(d)  sodium  hydroxide. 

27.  To  a  solution  containing  0.1  M  of  each  of  the 
ions  Ag+,  Cu+,  Fe+2,  and  Ca+2  is  added  2  M 
NaBr  solution,  giving  precipitate  A.  After  filtra- 
tion, a  sulfide  solution  is  added  to  the  solution 
and  a  black  precipitate  forms,  precipitate  B. 
This  precipitate  is  removed  by  filtration  and  2  M 
sodium  carbonate  solution  is  added,  giving  pre- 
cipitate C.  What  is  the  composition  of  each 
precipitate,  A,  B,  and  C? 


CHAPTER 


li 


Aqueous  Acids 
and  Bases 


The  acids,  •  •  •  are  compounded  of  two  substances  •  •  •  the  one 
constitutes  acidity,  and  is  common  to  all  acids,  •  •  •  the  other  is  pecul- 
iar to  each  acid,  and  distinguishes  it  from  the  rest   •    •    • 

A.    LAVOISIER,     1789 


In  Chapter  10  we  used  the  principles  of  equilib- 
rium to  help  us  understand  solubility  in  liquids. 
In  such  a  system  constituents  in  solution  reach 
the  dynamic  balance  of  equilibrium  with  another 
phase,  a  solid  or  a  gas.  Equilibrium  can  also 
exist  among  two  or  more  constituents  present  in 
the  same  solution.  One  of  the  examples  already 
encountered  (in  Chapter  9  and  in  Experiment 
15)  is 

Ft**(aq)  +  SCN-faqj  +±:  FeSCN^(aq)     (1) 

for  which  experiment  showed 


„  _     [FeSCN^I 
[Fe+3][SCN-] 


(2) 


In  reaction  (/),  all  of  the  molecular  species  in- 
volved in  the  equilibrium  are  in  the  solution  as 
dissolved  species.  Though  the  equilibrium  rela- 
tionship that  exists  among  the  concentrations 
is  a  little  more  complicated  than  in  the  solu- 
bility product  expressions,  the  guiding  princi- 
ples are  the  same. 

In  this  chapter  we  shall  explore  some  more 
equilibria  like  (/),  in  which  all  of  the  important 
species  are  dissolved.  We  will  consider  mainly 
those  equilibria  in  which  one  of  the  ions  is 
H+(aq).  This  type  of  equilibrium  furnishes  one 
of  the  most  important  classes  of  chemical  reac- 
tions of  all  those  that  occur  in  water. 


11-1    ELECTROLYTES— STRONG  OR  WEAK 


In  Section  10-2  we  considered  electrolytes,  sub- 
stances that  dissolve  in  water  to  give  solutions 
containing  ions.  Thus  far,  we  have  considered 
electrolytes  such  as  UC\(g)  and  NaOHfsJ: 

HClf*;  +  water  — >■  H+(aq)  +  C\-(aq)        (5) 


NaOHfsJ  +  water  — ■+-  Na+(aq)  +  OR-(aq)    (4) 

According  to  reaction  (3),  when  HC1  gas  dis- 
solves in  water,  all  of  the  HC1  molecules  break 
up,  or  dissociate,  into  ions,  H+(aq)  and  C\~(aq). 
There  is  no  experimental  evidence  for  the  pres- 

179 


180 


AQUEOUS    ACIDS    AND    BASES    I    CHAP.     1  1 


ence  of  molecules  of  HC1  in  aqueous  hydro- 
chloric acid  solutions.  In  a  similar  way,  there  is 
no  evidence  for  the  presence  of  molecules  of 
NaOH  in  aqueous  solution — apparently  the  so- 
dium hydroxide  crystal  breaks  up  completely 
into  sodium  ions,  Na+(agJ  and  hydroxide  ions, 
OH-(aq).  A  substance  that  dissolves  and  exclu- 
sively gives  ions  is  called  a  strong  electrolyte. 
Not  all  substances  that  form  conducting  solu- 
tions break  up,  or  dissociate,  so  completely.  For 
example,  vinegar  is  just  an  aqueous  solution  of 
acetic  acid.  Such  a  solution  conducts  electric 
current,  showing  that  ions  are  present: 

CH3COOH  +±l  H+(aq)  +  CHaCOO-faqj     (5) 

The  conductivity  of  a  0.1  M  acetic  acid  solu- 
tion is  much  lower,  however,  than  that  of  a 
0. 1  M  hydrogen  chloride  solution.  This  and  other 
experiments  show  that  only  a  small  fraction  of 
the  dissolved  acetic  acid,  CH3COOH,  has  formed 
ions.  Such  a  substance  that  dissolves  and  dis- 
sociates to  ions  only  to  a  limited  extent  is  called  a 
weak  electrolyte. 


Fig.  11-1.  A  strong  electrolyte  solution  conducts  bet- 
ter than  a  weak  electrolyte  solution. 


Careful  measurements  show  that  water  is, 
itself,  a  weak  electrolyte.  We  shall  consider  it 
first. 

11-1.1    Water:  A  Weak  Electrolyte 

Pure  water  does  not  conduct  electric  current 
readily.  Yet  an  extremely  sensitive  meter  shows 
that  even  the  purest  water  has  a  tiny  conduc- 
tivity. To  conduct  electric  current,  water  must 
dissociate  to  a  very  small  extent,  forming  ions. 
The  ions  prove  to  be  hydrogen  ion,  H+(aq),  and 
hydroxide  ion,  OH~( aq): 

HiOd)  +±  H+(aq)  +  OH-(aq)  (5) 

This  is  an  equilibrium  involving  three  species  in 
the  liquid  phase,  H20,  H+(aq),  and  OH~(aq). 
The  equilibrium  law  can  be  written 


[H+]fOH-] 
[H20] 


(7) 


However,  as  we  have  remarked  in  Section  9-2.2, 
the  concentration  of  H20  in  water  is  so  large, 
55.5  M,  and  so  few  ions  are  formed  that  its  con- 
centration is  virtually  constant.  Consequently, 
expression  (7)  is  usually  simplified  by  incorporat- 


HCl  :    a    strong    electrolyte 


CH3COOH  ■'    a    weak     elecrtrolylre 


SEC.     11-1    |    ELECTROLYTES  — STRONG    OR    WEAK 


181 


ing  the  factor  55.5  in  the  constant.  We  shall  do 
this,  and  label  the  constant  Kw  to  indicate  that 
it  includes  the  factor  [H2OJ: 

Km  =  [H+][OH-]  (8) 

Kw  =  [H,0]  X  K  =  55.5  X  K  (9) 

The  magnitude  of  Kw  is  given  in  Table  9-IV 
at  25°C, 

Kw=  1.00  X  10-"  (70) 

Having  this  value  of  Kw,  we  can  calculate  the  ion 
concentrations,  using  the  same  methods  we  ap- 
plied to  solubility. 

In  reaction  (6)  there  is  one  H+(aq)  ion  formed 
for  every  OH~(aq)  ion.  Hence,  in  pure  water 
where  the  only  source  of  ions  is  reaction  (6),  the 
concentrations  of  H+(aq)  and  OH~(aq)  must 
be  equal.  That  is, 

in  pure  water,  [OH"]  -  [H+]  (77) 

Substituting  (77)  into  the  equilibrium  expression, 
we  can  calculate  the  concentrations  of  the  two 
types  of  ions: 

K„  =  [H+]  X  [OH-]  =  [H+]  X  [H+]  =  [H+]2 


or, 


[H+]  =  Vkw  =  Vim  x  io-14 

[H+]  =  1.00  X  10~7M 


Because  of  (77),  we  can  also  write 

[OH-]  =  [H+]  =  1.00  X  10"7M 


(72) 


(7i) 


Now  we  can  explain  the  low  conductivity  of 
pure  water.  Though  water  dissociates  into  ions, 
H+(aq)  and  OH~(aq),  it  does  so  only  to  a  very 
slight  extent.  At  equilibrium,  the  ion  concentia- 
tions  are  only  10-7  M.  Water  is  a  weak  elec- 
trolyte. 

11-1.2    The  Change  off  K,  with  Temperature 

Experiments  show  that  reaction  (6)  absorbs  energy: 

HiO(l)  +  13.68  kcal  — )-  H+(aq)  +  OH(aq)    (14) 

Even  though  only  a  minute  fraction  of  the  water  present 
actually  is  dissociated  at  equilibrium,  if  we  measure  the 
energy  effect  and  divide  by  the  number  of  moles,  we  find 
that  it  takes  13.68  kcal  per  mole  of  water  broken  into 
ions. 

This  heat  effect  can  be  used  in  predicting  how  Kv 
changes  with  temperature.  Le  Chatelier's  Principle  indi- 


cates that  increased  temperature  will  shift  the  equilibrium 
condition  toward  larger  concentrations  of  the  ions  (so  as 
to  absorb  heat).  Hence,  Ku  is  expected  to  increase.  This 
agrees  with  the  experimental  values  for  Kw,  given  for 
various  temperatures  in  Table  11-1. 


Table  II -I 

VALUES    OF    K      AT    VARIOUS 
TEMPERATURES 

TEMPERATURE  (°C) 


0 

0.114  X  10-" 

10 

0.295  X  10-'< 

20 

0.676  X  10-" 

25 

1.00  x  io-» 

60 

9.55  X  10-" 

11-1.3    The  Special  Roles  off  H    aq    and 
OH  (aqj  in  Water 

The  equilibrium  reaction  (6)  gives  the  two  ions 
W+(aq)  and  OH~(aq)  special  roles  in  aqueous 
solutions: 


W(l) 


H+(aq)  +  OH-(aq) 


(6) 


In  pure  water,  where  the  only  source  of  ions 
is  reaction  (6),  the  concentrations  of  H+(aq)  and 
OH~(aq)  must  be  equal.  But  what  if  we  add 
some  HC1  to  the  solution?  We  have  already 
noted  that  HC1  is  a  strong  electrolyte,  dissolving 
to  give  the  ions  W+(aq)  and  C\~(aq).  Thus,  hy- 
drogen chloride  adds  H+(aq)  but  not  OH~(aq) 
to  the  solution.  The  concentrations  [H+]  and 
[OH-]  are  no  longer  equal.  However,  they  are 
still  found  to  be  "tied  together"  by  the  equilib- 
rium relationship 


or, 


Kw  =  [H+][OH-] 


[OH-]  = 


[H+] 


(5) 
(75) 


Expression  (75)  shows  that  as  the  concentration 
of  H+  rises  (for  example,  as  we  add  HC1  to 
water),  the  concentration  of  OH"  must  de- 
crease. 

Suppose,  on  the  other  hand,  that  we  add  so- 
dium hydroxide,  NaOH,  to  pure  water.  Sodium 
hydroxide  is  also  a  strong  electrolyte,  adding 


182 


AQUEOUS    ACIDS    AND    BASES    |    CHAP.    11 


OH-  ions  to  the  solution.  Now  we  can  rewrite 
(8)  in  the  form 


[H+]  = 


[OH-] 


(16) 


and  we  see  that  raising  the  concentration  [OH-] 
lowers  the  concentration  [H+]. 

In  this  way  the  two  ions  H+(aq)  and  OU~(aq) 
are  connected  in  the  chemistry  of  water.  When 
[H+]  is  high,  [OH-]  must  be  low.  When  [OH"] 
is  high,  [H+]  must  be  low. 

Let's  consider  an  example.  Suppose  we  dis- 
solve 0.10  mole  of  hydrogen  chloride  in  1.0  liter 
of  water.  Since  HC1  is  a  strong  electrolyte,  0.10 
mole  of  HC1  forms  0.10  mole  of  H+(aq)  ions 
and  0.10  mole  of  C\~(aq)  ions  in  the  one  liter 
volume.  The  concentrations  of  H+(aq)  and 
C\~(aq)  due  to  the  HC1  are  equal:* 

[H+]  =  [CI"]  -  0.10  M  (77) 

If  the  concentration  [H+]  is  0.10  M,  expres- 
sion (75)  allows  us  to  calculate  the  concentration 
[OH-]: 

roH-i  =  -^  =  im  x  10~"  =  10Q  x  10~14 
L       J      [H+]  o.io         "   l.o  x  10"1 

[OH"]  =  1.0  X  10-"  (18) 

We  see  that  adding  the  0.10  mole  of  HC1  to  1.0 
liter  of  water  lowered  the  OH"  concentration 
from  10-7  M  (its  value  in  pure  water)  to  10-13  M, 
a  change  by  a  factor  of  a  million! 

EXERCISE  11-1 

Show  that  the  addition  of  0.010  mole  of  solid 
NaOH  to  1 .0  liter  of  water  reduces  the  concen- 
tration of  H+(aq)  to  1.0  X  10"12  M. 

EXERCISE  11-2 

Suppose  that  3.65  grams  of  HC1  are  dissolved  in 
10.0  liters  of  water.  What  is  the  value  of  [H+]  ? 
Use  expression  (75)  to  show  that  [OH-]  = 
1.00  X  l0~nM. 


Thus  we  see  that  the  concentrations  of  the  two 


*  There  is  a  small  additional  concentration  of  H+(aq), 
owing  to  the  dissociation  of  water,  but  it  will  be  quite 
negligible  compared  with  the  0.10  M  concentration  pro- 
vided by  the  HC1. 


ions,  H+(aq)  and  OH~(aq),  always  remain  re- 
lated through  the  equilibrium  relation  (8).  In 
pure  water  they  are  equal,  [H+]  =  [OH-].  In  a 
solution  of  HC1,  in  which  [H+]  is  high,  [OH"] 
must  be  low.  In  a  solution  of  NaOH,  in  which 
[OH-]  is  high,  [H+]  must  be  low.  Furthermore, 
as  shown  in  the  calculation  (18)  and  Exercise 
11-1,  with  rather  low  concentrations  of  HC1 
and  NaOH,  the  concentration  of  H+(aq)  [or 
OH-(aq)]  can  be  varied  from  0.1  M  to  10"13  M, 
by  the  immense  factor  of  1012  (a  million  million). 

This  ease  with  which  we  can  control  and  vary 
the  concentrations  of  H+(aq)  and  OH~(aq) 
would  be  only  a  curiosity  but  for  one  fact.  The 
ions  H+(aq)  and  OH~(aq)  take  part  in  many  im- 
portant reactions  that  occur  in  aqueous  solution. 
Thus,  if  H+(aq)  is  a  reactant  or  a  product  in  a 
reaction,  the  variation  of  the  concentration  of 
hydrogen  ion  by  a  factor  of  1012  can  have  an 
enormous  effect.  At  equilibrium  such  a  change 
causes  reaction  to  occur,  altering  the  concentra- 
tions of  all  of  the  other  reactants  and  products 
until  the  equilibrium  law  relation  again  equals 
the  equilibrium  constant.  Furthermore,  there  are 
many  reactions  for  which  either  the  hydrogen 
ion  or  the  hydroxide  ion  is  a  catalyst.  An  ex- 
ample was  discussed  in  Chapter  8,  the  catalysis 
of  the  decomposition  of  formic  acid  by  sulfuric 
acid.  Formic  acid  is  reasonably  stable  until  the 
hydrogen  ion  concentration  is  raised,  then  the 
rate  of  the  decomposition  reaction  becomes  very 
rapid. 

In  these  ways,  by  affecting  equilibrium  condi- 
tions and  by  changing  reaction  rates,  the  con- 
centrations of  H+(aq)  and  OH~(aq)  give  us 
immense  leverage  in  controlling  the  chemistry 
of  aqueous  solutions. 

EXERCISE  11-3 

The  color  of  a  solution  of  potassium  chromate, 
K2Cr04,  changes  to  the  color  of  a  solution  of 
potassium  dichromate,  K2Cr207,  when  a  few 
drops  of  HC1  solution  are  added.  Write  the 
balanced  equation  for  the  reaction  between 
CrO;2(aq)  and  H+(aq)  to  produce  Cr207"2  and 
explain  the  color  change  on  the  basis  of  Le 
Chatelier's  Principle. 


SEC.    11-2    I    EXPERIMENTAL    INTRODUCTION    TO    ACIDS    AND    BASES 


183 


11-2    EXPERIMENTAL  INTRODUCTION  TO  ACIDS  AND  BASES 


The  significant  influence  that  H+(aq)  and 
OH~(aq)  exert  on  aqueous  solution  chemistry 
was  recognized  long  ago.  Consequently  chemists 
have  long  identified  by  the  class  names  "acids" 
and  "bases"  those  substances  that  change  the 
concentrations  of  these  two  ions.  We  shall  see, 
later  in  this  chapter,  that  the  meanings  of  the 
terms  "acid"  and  "base"  are  evolving  and  have 
become  more  general  in  modern  usage.  This  is  a 
natural  and  desirable  development  as  chemists 
attempt  to  relate  the  chemistry  of  aqueous  solu- 
tions to  the  chemistry  that  occurs  in  other 
solvents  and  in  other  phases.  For  the  moment, 
however,  we  shall  keep  our  attention  focused  on 
aqueous  solutions  and  consider  how  a  chemist 
recognizes  acids  and  bases. 


11-2.1    Properties  of  Aqueous  Solutions 
off  Acids 

Consider  the  following  compounds: 

HC1  :  hydrochloric   acid   (or,   hydrogen 

chloride), 
HN03  :  nitric  acid, 

CH3COOH  :  acetic  acid, 
H2SO4  :  sulfuric  acid, 

H3PO4  :  phosphoric  acid. 

Each  of  these  five  compounds  is  called  an 
acid.  They  all  share  this  name  because  they  have 
the  following  important  properties  in  common. 

Hydrogen  Containing.  Each  of  these  com- 
pounds contains  hydrogen. 


Fig.  11-2.  Some  familiar  acids. 


Electrical  Conductivity.  Each  of  these  com- 
pounds dissolves  in  water  to  form  solutions  that 
conduct  electricity.  Ions  are  present  in  these 
aqueous  solutions. 

Liberation  of  Hydrogen  Gas.  The  aqueous 
solutions  of  each  compound  produce  hydrogen 
gas,  H2,  if  zinc  metal  is  added. 

Color  of  Litmus.  Litmus,  a  dye,  is  red  in  color 
when  placed  in  these  aqueous  solutions. 

Taste.  The  dilute,  aqueous  solutions  of  each 
compound  are  sour  tasting.* 

Because  the  aqueous  solutions  of  these  com- 
pounds have  these  properties  in  common,  the 
compounds  are  conveniently  classed  together 
and  identified  as  acids.  In  fact,  these  properties 
constitute  the  simplest  definition  of  an  acid.  They 
provide  a  basis  for  deciding  whether  some  other 
compound  should  be  classified  as  an  acid. 

What  is  the  common  factor  that  makes  these 
different  substances  behave  in  the  same  ways? 
In  water  they  all  form  conducting  solutions;  we 
conclude  that  they  all  form  ions  in  water.  Each 
substance  contains  hydrogen  and  each  reacts 
with  zinc  metal  to  produce  hydrogen  gas.  Per- 
haps all  of  these  aqueous  solutions  contain  the 
same  ion  and  this  ion  accounts  for  the  formation 
of  H2(g).  It  is  reasonable  to  propose  that  the 
common  ion  is  H+(aq).  We  postulate:  a  sub- 
stance has  the  properties  of  an  acid  if  it  can  re- 
lease hydrogen  ions. 


11-2.2    Properties  off  Aqueous  Solutions 

of  Bases 

Consider  the  following  compounds: 

NaOH 

:  sodium  hydroxide, 

KOH 

:  potassium  hydroxide, 

Mg(OH)2 

:  magnesium  hydroxide, 

Na2C03 

:  sodium  carbonate, 

NH3 

:  ammonia. 

Citric 


Acetic 


Tartaric 


*  Many  chemicals,  including  some  acids  and  some 
bases,  are  poisons.  Therefore  chemists  rarely  use  taste  as 
a  voluntary  means  of  deciding  if  an  unidentified  solution 
is  acidic  or  basic. 


184 


AQUEOUS    ACIDS    AND    BASES   |   CHAP.    11 


Each  of  these  five  compounds  is  called  a  base. 
They  all  share  this  name  because  they  have  the 
following  important  properties  in  common. 

Electrical  Conductivity.  Like  acids,  these  com- 
pounds dissolve  in  water  to  form  conducting 
solutions.  Ions  are  present  in  an  aqueous  solu- 
tion of  a  base. 

Reaction  with  Acids.  When  one  of  these  com- 
pounds is  added  to  an  acid  solution,  it  destroys 
the  identifying  properties  of  the  acid  solution — 
all  but  electrical  conductivity. 

Color  of  Litmus.  The  dye,  litmus,  is  blue  in 
color  when  placed  in  an  aqueous  solution  of  any 
of  these  compounds. 

Taste.  The  dilute  aqueous  solutions  taste  bit- 
ter.* 

Feel.  The  aqueous  solutions  feel  "slippery." 

Again,  these  properties  constitute  the  simplest 
definition  of  a  base.  These  properties  provide  a 
basis  for  deciding  whether  some  other  compound 
should  be  classified  as  a  base. 

11-2.3    An  Explanation  of  the  Properties 
of  Bases 

Using  the  same  argument  we  used  for  acids,  we 
can  seek  a  common  factor  that  accounts  for  the 
similarities  of  bases.  Because  of  the  electrical 
conductivity,  we  might  seek  an  ion.  Because  of 
the  ability  to  counteract  the  properties  of  acids, 
we  ought  to  seek  an  ion  which  can  remove  the 
hydrogen  ion,  H+(aq),  since  hydrogen  ion  ac- 
counts for  the  properties  of  acids. 

Sodium  hydroxide,  NaOH,  when  dissolved  in 
water,  gives  a  solution  with  the  properties  of  a 
base.  The  hydroxides  of  many  elements — those 
from  the  left  side  of  the  periodic  table — behave 
in  the  same  way.  Perhaps  they  dissolve  to  form 
ions  of  the  sort 

NaOHfsj  q=t  Na+faq;  +  OH~(aq)        (19) 

KOH(s)  +±:  K+(aq)  +  OH-(aq)  (20) 

Mg(OH)2fsJ  =?=*:  Mg»(aq)  +  20Br(aq)    (21) 

Ca(OH)2fsj  *=±  C***(aq)  +  20H-(aq)     (22) 

The  hydroxide  ion,  OH-(aq),  could  react  with 

*  If  you  have  forgotten  the  danger  of  tasting  chemicals, 
reread  the  preceding  footnote. 


hydrogen  ion  to  account  for  the  second  property 
of  bases,  the  removal  of  acid  properties: 

OH'(aq)  +  H+(aq)  +±:  H20  (23) 

The  similarities  among  the  hydroxides  are  ob- 
vious. Let's  compare  sodium  carbonate  and  am- 
monia. Sodium  carbonate,  Na2C03,  dissolves  in 
water  to  give  a  solution  with  the  properties  that 
identify  a  base.  Quantitative  studies  of  the  solu- 
bilities of  carbonates  show  that  carbonate  ion, 
C03"2,  can  react  with  water.  The  reactions  are 

Na2COj(sJ  z^=±  2Na+( aq)  +  CO;2(aq)    (24) 

CO^(aq)  +  H20  3=fc 

HC03-  (aq)  +  OH-(aq)     (25) 

Reaction  (25)  indicates  that  the  presence  of  car- 
bonate ion  in  water  increases  the  hydroxide  ion, 
OH~,  concentration.  This  is  a  constituent  that 
is  present  in  the  solutions  of  NaOH,  KOH, 
MgCOH),,  and  Ca(OH)2. 

Reaction  (25)  also  provides  a  basis  for  under- 
standing the  removal  of  acid  properties.  If  C03~2 
readily  forms  bicarbonate  ion,t  HCOf,  then  re- 
action (26)  is  likely: 

C03-2f aq)  +  H+(aq)  z<=±:  HCO^(aq)     (26) 

We  see  that  the  existence  of  the  stable  bicar- 
bonate ion,  HCO3"  (aq)  produces  the  chemical 
species,  OH~(aq)  in  common  with  solutions  of 
the  hydroxides.  We  can  postulate  that  OH~(aq) 
accounts  for  the  slippery  "feel"  and  bitter  taste 
of  the  basic  solutions.  The  stability  of  bicarbo- 
nate ion  also  explains  the  removal  of  acid  prop- 
erties through  reaction  (26). 

The  fifth  substance  listed  as  a  base  is  ammonia, 
NH3.  Ammonia  readily  forms  ammonium  ion, 
NH^.  Ammonia  can  react  with  water, 

NHsf aq)  +  H20  +±  NHt(aq)  +  OH~(  aq)  (27) 
and  with  hydrogen  ion, 

NH,(aqJ  +  H+(aq) 


NHt(aq)        (28) 


Once  again,  the  formation  of  NH^  accounts  for 
ammonia  having  the  properties  of  a  base.  Reac- 
tion (27)  produces  hydroxide  ion,  which,  by  our 
postulate,  accounts  for  the  taste  and  "feel"  prop- 
erties of  solutions  of  bases.  Reaction  (28)  shows 

tBicarbonate  ion,  HCOf,  is  also  called  "hydrogen 
carbonate"  or  "monohydrogen  carbonate." 


SBC.    11-2    I   EXPERIMENTAL   INTRODUCTION   TO    ACIDS    AND    BASES 


185 


how  ammonia  can  act  to  destroy  the  acid  proper- 
ties of  a  solution  containing  hydrogen  ions, 
H+(aq). 

Investigation  of  the  reactions  of  other  com- 
pounds that  have  the  properties  of  a  base  shows 
that  each  compound  can  produce  hydroxide  ions 
in  water.  The  OH~(aq)  ions  may  be  produced 
directly  (as  when  solid  NaOH  dissolves  in  water) 
or  through  reaction  with  water  (as  when  Na2C03 
and  NH3  dissolve  in  water): 


NaOHfsJ 
COfYagj  +  H20 

NlUfaq)  +  H20 


zt  Na+(aq)  +  OH-(aq)    (19) 

~HCOa(aq)  +  OH-(aq)     (25) 

±NHtJaq)  +  OH-(aq)  (27) 


Furthermore,  any  substance  that  can  produce  hy- 
droxide ions  in  water  also  can  combine  with 
hydrogen  ions: 

OH-(aq)  +  H+(aq)  z<=h  H20  (23) 


CO^(aq)  +  H+(aq) 
mitfaq)  +  H+(aq) 


HCO;  (aq) 

NOt(aq) 


(26) 
(28) 


Since  production  of  OH-(aq)  and  reaction  with 
H+(aq)  go  hand  in  hand  when  we  are  dealing 
with  aqueous  solutions,  a  base  can  be  described 
either  as  a  substance  that  produces  OH~(aq)  or 
as  a  substance  that  can  react  with  H+(aq).  In 
solvents  other  than  water,  the  latter  description 
is  more  generally  useful.  Therefore,  we  postulate: 
a  substance  has  the  properties  of  a  base  if  it  can 
combine  with  hydrogen  ions. 

U-2.4    Acids  and  Bases:  Summary 

Let  us  repeat  our  two  definitions  and  explana- 
tions. 

Fig.  11-3.  Some  familiar  bases. 


DEFINITIONS 

An  acid  is  a  hydrogen-containing  substance  that 
has  the  following  properties  when  dissolved  in 
water: 

it  is  an  electrical  conductor; 
it  reacts  with  Zn  to  give  H2(gJ; 
it  makes  litmus  red; 
it  tastes  sour. 

A  base  is  a  substance  that  has  the  following 
properties  when  dissolved  in  water: 

it  is  an  electrical  conductor; 

it  reacts  with  an  acid,  removing  the  acidic 

properties; 
it  makes  litmus  blue; 
it  tastes  bitter; 
it  feels  slippery. 

EXPLANATIONS 

A  substance  is  an  acid  if  it  can  release  hydrogen 
ions,  H+(aq). 

A  substance  is  a  base  if  it  can  react  with  hy- 
drogen ions,  H+(aq). 


11-2.5    The  Nature  of  H+(aq) 

Both  of  our  explanations  of  the  properties  of  acids  and 
bases  involve  the  hydrogen  ion,  H+(aq).  This  species  has 
great  importance  in  the  chemistry  of  aqueous  solutions, 
so  we  shall  consider  what  is  known  about  it. 

Before  considering  what  a  chemist  means  by  the  sym- 
bols H+(aq),  we  must  discuss  more  generally  the  inter- 
action of  ions  with  water.  Lithium  chloride  provides  a 
good  example.  Lithium  chloride  dissolves  in  water  spon- 
taneously at  25°C,  forming  a  conducting  solution.  At 
equilibrium,  it  has  a  high  solubility: 


LiC\(s) 


Li+(aq)  +  a-(aq)        t  =  25°C    (29) 


Lithium  chloride  melts  spontaneously  above  613°C  and 
forms  a  liquid  that  conducts  electricity: 


LiClfsj 


Li+(7J  +  Cl-(l)        t  =  613°C     (30) 


Sodium  carbona-te 


Ammonia 


Third 


Equilibrium  in  any  reaction  is  determined  by  a  compro- 
mise between  tendency  toward  minimum  energy  ("golf 
balls  roll  downhill")  and  tendency  toward  maximum 
randomness.  Reaction  (29)  and  reaction  (30)  both  involve 
increase  in  randomness  since  the  regular  solid  lattice  dis- 
solves or  melts  to  become  part  of  a  disordered  liquid 
state.  Both  reactions  produce  ions.  But  reaction  (29) 
proceeds  readily  at  25°C,  whereas  reaction  (30)  does  not 


186 


AQUEOUS    ACIDS    AND    BASES   |   CHAP.    1  1 


Li  +  Caq) 
Li+-4If20 


Fig.  11-4.  Possible     tetrahedral  *     arrangements     of 

water  molecules  around  Li*  and  H*  ions. 


occur  until  the  solid  is  quite  hot,  613°C.  The  difference 
must  be  in  the  special  stabilities  of  Li+  and  Cl~  ions  in 
water.  The  high  melting  point  of  lithium  chloride  shows 
that  the  crystal  is  very  stable.  The  high  solubility  of 
lithium  chloride  in  water  can  be  explained  only  by  saying 
that  Li+ (aq)  and  Or(aq)  must  also  be  very  stable.  This 
means  that  water  must  interact  strongly  with  these  ions. 
A  similar  situation  exists  for  hydrochloric  acid,  HC1. 
This  gaseous  compound  dissolves  readily  in  water  at 
25°C: 


ions  as  lA+(aq),  H+(aq),  and  C\~(aq).  The  notation  (aq) 
reminds  us  that  the  ions  interact  strongly  with  the  solvent. 
The  symbol  is  purposely  vague  because,  in  most  cases, 
chemists  are  not  certain  of  the  arrangement  of  the  water 
molecules  around  a  given  ion.  For  Li+(aq)  and  H+(aq), 
there  may  be  simple  packing  of  four  water  molecules 
around  each  ion,  as  shown  in  Figure  11-4.  This  figure 
suggests  similarity  between  Li+(aq)  and  H+(aq). 

However,  most  chemists  feel  there  is  quite  a  difference 
between  these  two  ions.  After  all,  H+  is  unique — the 
proton  has  no  electrons.  Many  chemists  suggest  that  the 
proton  attaches  itself  strongly  to  one  molecule  of  water, 
forming  a  new  molecular  species,  HjO+ (aq).  Figure  11-5 


HC\(g)  *=fc  H+(aq)  +  C\-(aq)        t  =  25°C    (57)      ^    /;_5    A  mode,  of  hydronium  ion>  h3Q  . 


The  HC1  molecule  is  a  stable  one— it  must  be  heated  to  a 
few  thousand  degrees  before  the  atoms  will  separate. 
Even  then,  neutral  atoms  are  obtained  and  still  higher 
temperatures  are  needed  before  gaseous  ions  are  ob- 
tained. 


HCl(g) 


H(g)  +  Cl(g)        t  very  high     (32) 


The  high  temperature  required  to  separate  the  two  atoms 
of  a  molecule  of  HC1  shows  that  HC1  is  very  stable. 
Again,  we  can  explain  the  solubility  of  HC1  in  water  by 
saying  H+(aq)  and  C\~(aq)  must  also  be  very  stable. 
Water  must  interact  strongly  with  these  ions. 
This  is  why  we  have  been  symbolizing  these  aqueous 


*  The  word  tetrahedral  means  four-sided.  If  the  oxygen 
atoms  of  the  four  water  molecules  are  connected  by  lines, 
the  lines  form  a  four-sided  figure. 


SEC.    11-2    I    EXPERIMENTAL    INTRODUCTION    TO    ACIDS    AND    BASES 


187 


shows  a  model  of  this  proposed  ion — called  hydronium 
ion — surrounded  by  solvent.  Notice  that  the  three  hy- 
drogen atoms  are  pictured  as  equivalent.  There  is  a 
pleasing  similarity  to  the  formation  of  the  well-established 
ammonium  ion,  NH<+,  with  its  four  equivalent  hydrogen 
atoms: 

H,0  +  H+  +±  HjOVaqj  (33) 

NH3(aqj  +  H+  +=±  NH<+(aq)  (34) 

Still  other  chemists  feel  there  are  probably  several  ar- 
rangements of  water  molecules  around  the  proton.  In 
addition  to  HjO+,  there  may  be  molecules  such  as  HsO,j+, 
H7CV,  H9CV,  etc. 

Unfortunately,  the  experimental  data  do  not  provide 
a  definite  answer  to  the  nature  of  H+(aq).  The  hydronium 
ion,  shown  in  Figure  11-5,  does  exist  in  certain  crystal 
structures.*  Spectroscopic  studies f  indicate  that  several 
species  are  present  in  water.  Thermal  and  electrical  con- 


*  Hydronium  ion,  HsO+,  is  a  structural  unit  in  solid 
perchloric  acid  hydrate,  HC104  •  H2Of  as  shown  by  nuclear 
magnetic  resonance  studies. 

t  Spectroscopy  refers  to  the  study  of  the  absorption 
of  light — in  this  case,  by  aqueous  solutions  of  acids  such 
as  HC1. 


ductivities  of  aqueous  acid  solutions  have  been  inter- 
preted to  indicate  the  presence  of  a  molecular  unit,  H90«+. 

Figure  11-4  shows  the  hydrogen  ion  surrounded  by  four 
oxygen  atoms,  each  part  of  a  water  molecule.  We  could 
write  the  formula  for  this  arrangement  as  H+-4H20  or 
H9(X+.  But  Figure  11-6  shows  how  an  H30+  ion  can  serve 
as  the  basis  for  another  structure  with  the  formula 
HjO+-3HjO,  again  giving  H9CV.  Still,  there  seems  to  be 
no  compelling  experimental  basis  for  preference. 

So  we  are  faced  with  at  least  three  plausible  structures 
for  the  species  H+(aq),  as  shown  in  Figures  11-4,  11-5, 
and  11-6.  Under  these  circumstances,  convenience  in  dis- 
cussing the  experimental  properties  of  H+(aq)  governs 
our  choice. 

So  far  in  this  chapter,  we  have  referred  to  the  aqueous 
hydrogen  ion  as  H+(aq).  Later  in  the  chapter  we  will 
consider  a  more  general  theory  of  acids  and  bases.  Then 
it  will  be  more  convenient  to  designate  the  aqueous  hy- 
drogen ion  as  H30+(aq).  The  reason  will  be  clear— such 
usage  aids  us  in  seeing  regularity  in  the  behavior  of  a 


Fig.  11-6.  A  possible  model  for  H»(V  based  upon 

HaO*. 


188 


AQUEOUS    ACIDS    AND    BASES    I    CHAP.     1  1 


larger  class  of  acids  and  bases.  That  is  sufficient  basis  for 
the  use  of  any  theory. 


11-2.6    Acid-Base  Titrations 

We  have  noted  that  the  two  concentrations  [H+] 
and  [OH"]  are  "tied  together."  If  0.100  mole  of 
HC1  is  dissolved  in  0.100  liter  of  water  to  raise 
[H+]  to  LOOM,  then  [OH"]  goes  down  to 
1.00  X  10-14  M  and  the  product  [H+]  X  [OH"] 
remains  equal  to  Kw  =  1.00  X  10-14.  Perhaps 
you  are  wondering  what  happens  to  the  hydrox- 
ide ions  to  reduce  their  concentration  from 
1.00  X  10-7  M  (their  concentration  in  pure  wa- 
ter) to  the  new  value,  1.00  X  10~14  M.  The  an- 
swer is  that  they  are  consumed  through  reaction 
with  the  added  H+(aq): 


OR-(aq)  do  not  satisfy  the  equilibrium  expres- 
sion. Their  product  far  exceeds  1.00  X  10~14: 


OH(aq)  +  H+(aq) 


H20 


(35) 


This  is  in  accord  with  Le  Chatelier's  Principle. 
Addition  of  HC1  to  water  raises  [H+].  By  Le 
Chatelier's  Principle,  processes  take  place  that 
tend  to  counteract  partially  the  imposed  change. 
Reaction  with  OH~(aq)  does  tend  to  counteract 
the  raised  concentration  of  H+(aq). 

In  the  case  we  have  considered,  the  actual 
amount  by  wh'ch  [H+]  is  reduced  is  extremely 
small.  To  reduce  [OH"]  from  1.00  X  10~7  M  to 
1.00  X  10~14M  (by  a  factor  of  107),  reaction 
(35)  must  consume  about  10-7  mole  of  hydrox- 
ide ion  for  every  liter  of  solution.  Since  one  mole 
of  OH~(aq)  reacts  with  one  mole  of  H+(aq),  the 
amount  of  H+(aq)  required  is  also  10-7  mole 
for  every  liter  of  solution.  Subtracting  10-7 
mole/liter  from  a  concentration  near  1  mole/liter 
causes  such  a  small  change  in  [H+]  that  it  need 
not  be  considered  in  calculations  (such  as  in 
Exercises  11-1  and  11-2). 

HC1    AND   NaOH    IN   THE   SAME 

solution:  excess  HC1 

Suppose  that  to  the  0.100  liter  of  LOOM  HC1 
solution  we  add  0.090  mole  of  solid  sodium  hy- 
droxide. Now  we  have  added  both  H+(aq)  and 
OH~(aq)  in  high  concentration  to  the  same  solu- 
tion. What  will  happen? 

Immediately  after  the  sodium  hydroxide  dis- 
solves,   the    concentrations    of    H+(aq)    and 


initial  [H+]  =  LOOM 

0.090  mole 


initial  [OH"]  = 


0.100  liter 


=  0.90  M 


initial  product,  [H+]  X  [OH"]  =  9.0  X  10"1 
(far  exceeding  1.00  X  10~14) 

Again  Le  Chatelier's  Principle  tells  us  quali- 
tatively what  will  occur  and  the  equilibrium 
expression  tells  us  quantitatively.  If  we  add 
OH-(aq),  a  change  will  take  place  that  tends  to 
counteract  partially  the  resulting  increase  in 
[OH-].  This  occurs  through  the  reaction  be- 
tween OH-(aq)  and  H+(aqJ,  consuming  both 
ions  and  reducing  the  value  of  [H+]  X  [OH-]. 
Reaction  continues  until  this  product  reaches  the 
equilibrium  value,  Kw  =  1.00  X  10-14. 

Since  Kw  is  so  small,  the  reaction  consumes 
almost  all  of  one  of  the  constituents  (H+  or 
OH~)  if  the  other  is  present  in  excess.  In  our 
example,  [H+]  initially  exceeds  [OH-]  by  0.10 
mole/liter: 

initial  [H+]  -  initial  [OH"]  =  excess  [H+] 
LOOM     -       0.90  M       =      0.10  M 

In  Section  11-1.3  we  calculated  that  if  the 
concentration  [H+]  =  0.100  M,  then  at  equilib- 
rium, [OH"]  =  LOO  X  10"13  M.  Thus  the  rather 
small  excess  of  hydrogen  ion,  0.10  M,  is  sufficient 
to  guarantee  that  the  reaction  between  H+(aq) 
and  OH~(aq)  consumes  most  of  the  0.90  M 
OH",  reducing  [OH"]  to  1.0  X  10"13  M. 


EXERCISE  11-4 

Suppose  that  0.099  mole  of  solid  NaOH  is 
added  to  0.100  liter  of  1.00  M  HC1. 

(a)  How  many  more  moles  of  HC1  are  present  in 
the  solution  than  moles  of  NaOH? 

(b)  From  the  excess  number  of  moles  and  the 
volume,  calculate  the  concentration  of  excess 
H+(aqj. 

(c)  Calculate  the  excess  concentration  of  H+(aq) 
from  the  difference  between  the  initial  con- 
centrations of  HC1  and  NaOH. 


SEC.     11-2    I    EXPERIMENTAL    INTRODUCTION    TO    ACIDS    AND    BASBS 


189 


(d)  Calculate  the  concentration  of  OH~(aq)  at 
equilibrium  (see  your  calculations  for  Exer- 
cise 11-2). 


HC1    AND    NaOH    IN   THE   SAME 

solution:  excess  NaOH 

Returning  to  our  original  0.100  liter  of  1.00  M 
HC1,  let  us  now  consider  the  addition  of  0.101 
mole  of  solid  NaOH.  Again  we  have  added  both 
H+(aq)  and  OH~(aq)  to  the  same  solution,  and 
the  concentrations  immediately  after  mixing  do 
not  satisfy  the  equilibrium  expression: 

initial  [H+]  -  LOOM 

•»•  i  rAu-i       0.101  mole       ,  A1  „,, 
initial   OH      =  ...... —  =  1.01  M 

L        J       0.100  liter 

initial  product,  [H+]  X  [OH"]  =  1.01 
(far  exceeding  1.00  X  10~14) 

This  solution  contains  excess  hydroxide  ion. 
Since  OH~(aq)  is  in  excess,  almost  all  of  the 
H+(aq)  will  be  consumed,  forming  water: 

initial  [OH"]  -  initial  [H+]  =  excess  [OH"] 
1.01  M       -      LOOM      =       0.01  M 

In  Exercise  11-1  we  calculated  the  equilibrium 
concentration  [H+]  in  a  solution  containing 
[OH-]  =  0.01  M.  The  result  is  [H+]  =  1.0  X 
10"12  M. 

HC1  and  NaOH  in  the  samb 
solution:  no  excess  of  either 

In  each  example  used  in  this  section,  a  number 
of  moles  of  NaOH  was  added  to  0.100  liter  of 
1 .00  M  HC1  with  one  or  the  other  constituent  in 
excess.  Reaction  between  H+(aq)  and  OH~(aq) 
consumes  almost  all  of  the  constituent  not  in 
excess.  Let  us  now  consider  the  case  in  which 
there  is  excess  of  neither  HC1  nor  NaOH. 

Suppose  we  add  0. 100  mole  of  NaOH  to  0. 100 
liter  of  LOOM  HC1.  The  initial  values  of  [H+] 
and  [OH-]  are  equal  and  their  product  far  ex- 
ceeds 1.00  X  10-14 

initial  [H+]  =  LOO  M 

initial  [OH-]  =  ai,00niole  =  LOOM 


0.100  liter 

initial  product,  [H+]  X  [OH~] 
(far  exceeding  1.00  X  10-14) 


1.00 


Reaction  between  H+(aq)  and  OH~(aq)  must 
occur,  forming  water: 

OH-(aq)  +  H+(aq)  3=fc  H20 

Since  one  mole  of  OH~(aq)  consumes  one  mole 
of  H+(aq),  the  concentrations  [H+]  and  [OH-] 
remain  equal  as  reaction  (35)  proceeds.  When 
equilibrium  is  reached,  they  will  still  be  equal. 
This  is  exactly  the  situation  in  pure  water.  As  we 
saw  in  Section  11-1.2, 

[H+]  =  [OH-]  -  VY„  =  LOO  X  10"7  M 

A  solution  containing  exactly  equivalent 
amounts  of  acid  and  base  is  neither  acidic  nor 
basic.  Such  a  solution  is  called  a  neutral  solu- 
tion.* 

progressive  addition  of  NaOH 
to  HC1:  a  titration 

Now  we  have  considered  the  progressive  addi- 
tion of  more  and  more  NaOH  to  a  fixed  amount 
of  HC1  solution.  The  results  are  compiled  in 
Table  11-11. 

Table  11-11  shows  that  the  concentration  of 
hydrogen  ions  changes  drastically  as  the  amount 
of  NaOH  nears  the  equivalent  amount  of  HC1. 
There  is  a  change  of  [H+]  by  a  factor  of  1010  as 
the  initial  concentration  of  OH~(aq)  is  changed 
from  0.99  to  1.01.  Thus  the  concentration 
changes  by  a  huge  factor  near  the  point  at  which 
the  amounts  of  acid  and  base  are  equivalent. 
Because  of  this,  the  progressive  addition  of  a  base 
to  an  acid,  a  titration,  furnishes  a  sensitive  means 
of  comparing  the  concentrations  of  an  acid  and 
a  base  solution. 

An  acid-base  titration  is  carried  out  by  adding 
carefully  measured  amounts  of  a  base  solution 
to  a  known  volume  of  the  acid  solution.  The  acid 
solution  contains  some  substance  that  provides 
visual  evidence  of  the  magnitude  of  [H+],  The 
dye  litmus  is  such  a  substance.  As  mentioned  in 
Sections  11-2.1  and  11-2.2,  litmus  is  red  in  solu- 
tions containing  excess  [H+].  Litmus  is  blue  in 

*  This  use  of  the  word  "neutral"  for  a  solution  with 
equal  amounts  of  H+  and  OH~  has  its  disadvantages 
because  the  same  word  is  used  in  reference  to  electrica' 
neutrality.  Aqueous  solutions  are  always  electrically  neu- 
tral whether  there  is  an  excess  of  either  H+  or  OH"  or 
an  excess  of  neither. 


190 


AQUEOUS    ACIDS    AND    BASES    I    CHAP.    11 


Table  11 -II.    the  concentrations  [h  ]  and  coh -]  in  solutions  containing 

BOTH    HCI    AND    NaOH 


INITIAL 
CONC. 

H+ 


INITIAL 
CONC. 

OH- 


EXCESS 
CONC. 

H+  or  OH- 


CALC. 

[H+] 


CALC. 
[OH"! 


1.00  M 

none 

1.00MH+ 

1.00  M 

1.0  X  10-»  M 

1.00 

0.90  M 

0.10MH+ 

1.0  X  io-1 

1.0  X  10"u 

1.00 

0.99 

0.01  MH+ 

1.0  x  io-j 

1.0  x  io-» 

1.00 

1.00 

none 

1.0  X  io-7 

1.0  X  io-7 

1.00 

1.01 

0.01  MOH- 

1.0  x  io-» 

1.00  x  io-» 

solutions  in  which  [H+]  is  less  than  [OH-].  A 
small  amount  of  litmus  in  the  solution  will  tint 
the  solution  red  until  the  point  in  the  titration 
at  which  the  number  of  moles  of  OH~(aq)  be- 
comes equal  to  the  number  of  moles  of  H+(aq). 
Then,  the  slightest  addition  of  more  OH~(aq) 
causes  a  drastic  reduction  in  [H+]  and  the  solu- 
tion color  turns  to  blue.  A  dye  whose  color  is 
sensitive  to  the  change  of  [H+]  (such  as  litmus) 
is  called  an  acid-base  indicator. 


11-2.7     pH 

For  compact  expression  of  H+(aq)  concentrations,  chem- 
ists use  a  quantity,  pH,  defined  by  the  equation 

pH  =  -log.o  [H+] 

Since  [H+]  =  10~7  M  in  a  neutral  solution  at  25°C,  it 
follows  that  for  such  a  solution 

pH  =  -log.o  [IO"7]  =  -(-7)  =  +7 


This  result  helps  us  to  understand  the  symbol  pH,  first 
defined  by  a  Danish  chemist,  Sorenson.  He  used  the  p  to 
stand  for  the  Danish  word  poienz  (power)  and  H  to  stand 
for  hydrogen.  After  a  change  of  sign,  pH  is  the  power  of 
ten  needed  to  express  the  hydrogen  ion  concentration  in 
moles  per  liter.  In  acidic  solutions,  p\\  is  less  than  7 
(pH  <  7);  and  in  basic  solutions,  pH  is  greater  than  7 
(jjH  >  7).  Table  1 1  -III  expresses  the  results  of  Table 
11-11  in  terms  of  pH. 


Table  11 -III 

CONCENTRATIONS    OF    H     aq      AND 
OH   (aq)    EXPRESSED    IN    TERMS    OF    pH 

ACIDITY  OR 


BASICITY 

[H+] 

pH 

acidic 

1.00 

0 

acidic 

io-» 

1 

acidic 

io-1 

2 

neutral 

io-7 

7 

basic 

io-»s 

13 

11-3    STRENGTHS  OF  ACIDS 


Earlier  in  this  chapter,  strong  and  weak  electro- 
lytes were  distinguished  in  terms  of  the  degree 
to  which  the  dissolved  material  forms  ions.  As 
a  particular  case,  such  distinctions  can  be  made 
in  terms  of  acids,  furnishing  a  quantitative  basis 
for  defining  the  strength  of  an  acid. 

11-3.1    Weak  Acids 

We  have  contrasted  the  electrical  conductivities 
of  0.1  M  aqueous  solutions  of  hydrochloric  acid 


and  acetic  acid  (see  Figure  11-1).  In  water,  hy- 
drochloric acid  dissociates  completely  to  ions; 
HCI  is  a  strong  electrolyte.  Because  one  of  the 
ions  released  is  H+(aq),  HCI  is  also  called  a 
strong  acid.  Acetic  acid,  on  the  other  hand,  dis- 
sociates to  ions  only  to  a  slight  extent;  acetic 
acid  is  a  weak  electrolyte.  Because  one  of  the  ions 
released  is  H+(aq),  acetic  acid  is  called  a  weak 
acid. 

These  qualitative  ideas  can  be  expressed  more 
usefully  in  terms  of  the  principles  of  equilibrium. 


SEC.     11-3    I    STRENGTHS    OF    ACIDS 


191 


For  example,  let  us  contrast  the  behavior  of  two 
weak  acids,  acetic  acid  and  hydrofluoric  acid: 

CH3COOHfaqJ  +± 

H+(aq)  +  CH3COO(aq)     (36) 

HF(aq)  3=fc  H+(aq)  +  ¥~(aq)  (37) 

Measurements  of  the  electrical  conductivities  of 
0.10  M  solutions  of  these  two  acids  show  that 
there  are  more  ions  present  in  the  HF  solution 
than  in  the  acetic  acid  solution.  We  can  conclude 
that  acetic  acid  is  a  weaker  acid  than  HF.  This 
information  is  conveyed  quantitatively  in  terms 
of  the  equilibrium  constants  for  reactions  (36) 
and  (37): 


y  [H+][CH3COO-]       ,  8  v  in_6 

*CH>C00H  =     [CHaCOOH]      =  L8  X  10 


KuF  -  [H+][F1 

KHF  -      [HF] 


6.7  X  10-< 


(38) 


(39) 


Since  A"Hf  is  a  larger  number  than  KCHiC00n, 
hydrofluoric  acid  dissociates  in  water  to  a  larger 
extent  than  does  acetic  acid.  Though  HF  is  a 
weak  acid  (only  partially  dissociated),  it  is  a 
stronger  acid  than  is  acetic  acid. 

We  can  express  these  ideas  in  terms  of  a 
general  acid,  HB.  The  acidic  nature  of  HZ?  is  con- 
nected to  its  ability  to  release  hydrogen  ions, 

HB(aq)  +±:  H+(aq)  +  B(aq)  (40) 


The  equilibrium  constant  for  reaction  (40)  meas- 
ures quantitatively  the  ease  with  which  Hi?  re- 
leases H+(aq)  ions, 


Ka  =  [H+]ffl-] 
A  [HB] 


(41) 


Tables  of  KA  furnish  a  quantitative  measure  of 
acid  strengths  with  which  we  can  compare  dif- 
ferent acids  and  predict  their  properties.  Several 
values  of  KA  are  given  in  Table  1 1-IV. 

We  see  in  Table  1 1-IV  that  the  equilibrium 
view  of  acid  strengths  suggests  that  we  regard 
water  itself  as  a  weak  acid.  It  can  release  hydro- 
gen ions  and  the  extent  to  which  it  does  so  is 
indicated  in  its  equilibrium  constant,  just  as  for 
the  other  acids.  We  shall  see  that  this  type  of 
comparison,  stimulated  by  our  equilibrium  con- 
siderations, leads  us  to  a  valuable  generalization 
of  the  acid-base  concept. 

EXERCISE  11-5 

Which  of  the  following  acids  is  the  strongest 
acid  and  which  the  weakest? 

nitrous  acid,  HN02;  Kym0i  =  5.1  X  10-4 
sulfurous  acid,  H2S03;  Kns0l  =  1.7  X  10~2 
phosphoric  acid,  H3P04;  /vH,po,  =  7.1  X  10~3 


Table  11-IV.    relative  strengths  of  acids 

AT    ROOM    TEMPERATURE 


~        [H±MJ 
AA  [HB] 


ACID 


STRENGTH 


N    AQUEOUS    SOLUTION 


REACTION 


KA 


HC1 

HNO3 

H2S04 

HSOr 

HF 

CH3COOH 

H2CO3  (C02  +  H20) 

H2S 

NH<+ 

HCO,- 

H20 


very  strong 

t 

t 

very  strong 

.  \.  t  ..  . 

strong 

We  a 


k 


:M 


I 

weak 

I 

very  weak 


HC\(g) 

HNCVsJ 

H2S04 

HSOc(aq) 

HF (aq) 

CH3COOH(aqJ 

H2COj(aq) 

H£(aq) 

KHS(aq) 

HC03~(aq) 

H2Q(aq) 


H+(aq) 
H+(aq) 
H+(aq) 

H+(aq) 
H+(aq) 
H+(aq) 
H+(aq) 
H+(aq) 
H+(aq) 
H+(aq) 
H+(aq) 


+  C\~(aq) 
+  N03~(aq) 
+  HSO4- (aq) 

+  SOrYaqJ 

+  r-(aq) 

+  CH3COO(aq) 
+  HCOr(aq) 
+  HS-(aq) 
+  NH3(aq) 
+  COr'(aq) 
+  OH-(aq) 


very  large 
very  large 
large 

1.3  X  10-* 

6.7  X  10-< 

1.8  X  10-' 

4.4  X  10-7 

1.0  x  10-7 

5.7  X  10-10 

4.7  X  10-" 

1.8  X  10-'«  • 


A',, 


1.00  X  10"u 


*The  equilibrium  constant,  KA,  for  water  equals  7777b;  =  — — 777 See  Section  11-1.1. 

[H2UJ  jj.j 


192 


AQUEOUS    ACIDS    AND    BASES    I    CHAP.    11 


11-3.2    Equilibrium  Calculations  of  Acidity 

The  acidity  of  a  solution  has  pronounced  effects 
on  many  chemical  reactions.  It  is  therefore  im- 
portant to  be  able  to  learn  and  control  the  hy- 
drogen ion  concentration.  This  control  is  ob- 
tained through  application  of  the  Equilibrium 
Law.  Common  types  of  calculation,  based  on 
this  law,  are  those  needed  to  determine  KA  from 
experimental  data  and  those  using  KA  to  find 
[H+].  We  will  illustrate  both  of  these  types,  using 
benzoic  acid,  C6H6COOH,  as  an  example. 

DETERMINATION    OF    KA 

To  apply  the  Equilibrium  Law  to  acid  solutions, 
a  chemist  must  know  the  numerical  value  of  the 
equilibrium  constant,  KA.  Experiments  which 
provide  this  information  require  the  measure- 
ment of  hydrogen  ion  concentration.  Acid-sensi- 
tive dyes,  such  as  litmus,  offer  the  easiest  esti- 
mate of  [H+]. 

A  typical  example  is  as  follows.  Benzoic  acid, 
C6H5COOH,  is  a  solid  substance  with  only  mod- 
erate solubility  in  water.  The  aqueous  solutions 
conduct  electric  current  and  have  the  other  prop- 
erties of  an  acid  listed  in  Section  1 1-2.1.  We  can 
describe  this  behavior  with  reaction  (42)  leading 
to  the  equilibrium  relation  (43): 


C6H5COOH(aqJ 


H+(aq)  +  C6UbCOO-(aq) 
(42) 


=  [H+][C6H5COO-] 
[QH5COOH] 


(43) 


The  following  experiment  was  performed  to 
determine  the  equilibrium  constant  in  (43).  A 
1.22  gram  sample  of  benzoic  acid  was  dissolved 
in  1.00  liter  of  water  at  25°C.  With  dyes  whose 
color  is  sensitive  to  acidity  (indicators)  the  con- 
centration of  H+(aq)  was  estimated  to  be 
8  X  10-"  M. 

To  make  use  of  these  data,  we  must  first  ex- 
press all  quantities  in  terms  of  moles.  The  mo- 
lecular weight  of  benzoic  acid,  C6H5COOH,  is 
122.1  grams/mole.  Hence, 

1.22  g 


1.22  g  QH5COOH  = 


122.1  g  mole 
=  0.0100  mole  C6H5COOH 


zoic  acid  if  we  assume  very  little  of  it  has  reacted 
to  form  H+(aq)  according  to  reaction  (42): 

moles        0.0100  mole 


[C6H5COOH] 
[QH6COOH] 


volume         1.00  liter 
0.0100  M  =  1.00  X  10~2M 

and,  by  measurement, 

[H+]  =  8  X  10-"M 

(Notice     that     [H+]     is    less    than 


(44) 


(45) 

is  less  man  10%  of 
[C6H5COOH],  verifying  our  assumption  that 
very  little  of  the  acid  reacted.)  Now  we  know  two 
of  the  concentrations  in  expression  (43)  and,  to 
complete  the  calculation,  we  must  know  the  con- 
centration of  benzoate  ion,  [C6H5COO_].  Since 
the  benzoic  acid  was  dissolved  in  pure  water,  the 
only  source  of  C6H5COO~  is  reaction  (42).  This 
is  also  the  source  of  hydrogen  ion,  H+(aq).  Since 
these  two  ions  are  both  produced  only  by  reac- 
tion (42),  their  concentrations  must  be  equal  in 
this  solution.  That  is, 

[H+]  =  [QH5COO-] 

Thus,  if  [H+]  =  8  X  10~4,  then,  also 

[QH5COO-]  =  8  X  10-"  (46) 

Now  we  can  complete  the  calculation  by  sub- 
stituting (44),  (45),  and  (46)  into  (43): 

=  [H+][C6H5COO]  =  f8  X  10~41  X  f8  X  10~4] 
A  [C6H5COOH]  [1.0  X  10-2] 

=  64  X  10'8 

1.0  x  10-2 

KA  =  64  X  10-«  =  6.4  X  10-B 

CALCULATION    OF    [H+] 

Having  established  experimentally  the  numerical  value  of 
KA,  we  can  use  it  in  calculations  of  equilibrium  concen- 
trations. 

As  an  example,  suppose  a  chemist  needs  to  know  the 
hydrogen  ion  concentration  in  a  solution  containing  both 
0.010  M  benzoic  acid,  C6H6COOH,  and  0.030  M  sodium 
benzoate,  C6H6COONa.  Of  course,  he  could  go  to  the 
laboratory  and  proceed  to  investigate  the  colors  of  in- 
dicator dyes  placed  in  the  solution.  However,  it  is  easier 
to  calculate  the  value  of  [H+],  using  the  accurate  value 
of  Ka  listed  in  Appendix  2. 

Sodium  benzoate  is  a  strong  electrolyte;  its  aqueous 
solutions  contain  sodium  ions,  Nd+(aq),  and  benzoate 
ions,  QHsCOO-faqJ.  Hence  the  equilibrium  involved  is 
the  same  as  before: 


Now  we  can  calculate  the  concentration  of  ben-       c6VUCOOH(aq)  -<-*"  H+faqJ  +  C6HbCOQ-(aq)    (42) 


SEC.    11-3    I    STRENGTHS    OF    ACIDS 


193 


At  equilibrium,  the  concentrations  must  be  in  accord  with 
the  equilibrium  expression.  That  is, 


KA  = 


rH+irc6Hscoo-] 

[C6HsCOOH] 


=  6.6  X  10-* 


(43) 


First,  let  us  assume  that  very  little  [H+]  is  formed  through 
dissociation  of  benzoic  acid.  This  assumption  implies  that 
the  concentrations  of  benzoate  ion  and  benzoic  acid  are 
very  little  affected  by  reaction  (42).  Assuming  this,  we  see 
that  two  of  the  concentrations  in  (39)  are  already  speci- 
fied: 

[C6H6COOH]  =  0.010  M 
[QH6COO-]  -  0.030  M 


_  rH+irC^COO-]  =  rH+1(0.030) 
A  [C6H5COOH]  (0.010) 

Rearranging  (47),  we  obtain 
[H+]  =  2.2  X  10-* 


(47) 


0.030 


The  calculation  cannot  be  considered  complete  until 
we  check  the  assumption.  Was  it  reasonable  to  assume 
that  the  concentrations  of  benzoate  ion  and  benzoic  acid 
were  not  changed  by  reaction  (42)1  To  decide,  we  com- 
pare the  magnitude  of  [H+],  2.2  X  10~5  M,  to  the  ben- 
zoate ion  and  benzoic  acid  concentrations.  We  find  that 
[QH6COOH]  =  0.010  M  is  about  500  times  larger  than 
the  concentration  change  necessary  to  form  2.2  X  10-5  M 
H+.  The  same  argument  applies  to  [QH6COO-].  The 
assumption  is  valid. 


11-3.3    Competition  for  H    Among  Weak  Acids 

We  have  explained  the  properties  of  acids  in 
terms  of  their  abilities  to  release  hydrogen  ions, 
H+(aq).  Thus  acetic  acid  is  a  weak  acid  because 
of  the  slight  extent  to  which  reaction  (48)  re- 
leases U+(aq): 


CH3COOH(aq) 


H+(aq)  +  CH3COO-(aq)    (48) 


We  have  explained  the  properties  of  bases  in 
terms  of  their  abilities  to  react  with  hydrogen  ion. 
Thus  ammonia  is  a  base  because  it  can  react  as 
in  (49): 

NH3(aq)  +  Y\+(aq)  Zf±:  NH  +  (aq)        (49) 

Now  consider  the  result  of  mixing  aqueous  solu- 
tions of  acetic  acid  and  ammonia.  The  reaction 
that  occurs  can  be  compared  to  a  sequence  of 
reactions, 


CH3COOH(aq) 
NH3(aq)  +  H+(aq) 


H+(aq)  +  CH3COO-(aq) 

NH4+  (aq) 


Net  reaction 

CH3COOH(aq)  +  NH3f  aq)  +± 

CH3COO  ~(aq)  +  WHt(aq)     (50) 

Practically,  the  result  of  reactions  (48)  and  (49) 
is  reaction  (50).  In  reaction  (50),  we  see  that 
acetic  acid  acts  as  an  acid  in  the  same  sense  that 
it  does  in  (48).  In  either  case,  it  releases  hydrogen 
ions.  In  (48)  acetic  acid  releases  hydrogen  ions 
and  forms  H+(aq)  and  in  (50)  it  releases  hydro- 
gen ions  to  NH3  and  forms  NH^ .  In  the  same 
way,  ammonia  acts  as  a  base  in  (50)  by  reacting 
with  the  hydrogen  ion  released  by  acetic  acid. 
So  reaction  (50)  is  an  acid-base  reaction,  though 
the  net  reaction  does  not  show  H+(aq)  explicitly. 
Now  by  taking  one  more  step  we  can  view 
acid-base  reaction  in  a  broader  sense.  Suppose 
we  mix  aqueous  solutions  of  ammonium  chlo- 
ride, NH4CI,  and  sodium  acetate,  CH3COONa. 
A  sniff  indicates  ammonia  has  been  formed.  Re- 
action occurs, 


NH4+  (aq)  +  CH3COO~(  aq)  =e=t 

CHaCOOHfaqj  +  NH3(aq) 


(51) 


Reaction  (57)  is  just  the  reverse  of  reaction 
(50).  Inspection  of  this  reaction  reveals  that  re- 
action (51),  too,  is  an  acid-base  reaction!  Once 
again  there  is  an  acid  that  releases  H+ — it  is  NH^ 
— and  a  base  that  accepts  H+ — the  base  is 
CH3COO-.  Once  again  the  net  effect  of  the  re- 
action is  transfer  of  a  hydrogen  ion  from  one 
species  to  another.  We  see  that  the  acid-base 
reaction  between  acetic  acid  and  ammonia  gave 
two  products,  one  an  acid,  NH^,  and  one  a 
base,  CH3COO-.  A  little  thought  will  convince 
you  that  every  acid-base  reaction  does  so.  The 
transfer  of  a  hydrogen  ion  from  an  acid  to  a 
base  necessarily  implies  that  it  might  be  handed 
back.  The  reaction  of  handing  it  back,  the  re- 
verse reaction,  is  just  as  much  a  hydrogen  ion 
transfer,  hence  an  acid-base  reaction,  as  is  the 
original  transfer. 

Notice  that  we  are  now  referring  to  reactions 
in  which  a  hydrogen  ion  is  transferred  from  an 
acid  to  a  base  without  specifically  involving  the 
aqueous  species  H+(aq).  A  hydrogen  ion,  H+,  is 


194 


AQUEOUS    ACIDS    AND    BASES    I    CHAP.     11 


nothing  more  than  a  proton.  Consequently  we 
can  frame  a  more  general  view  of  acid-base 
reactions  in  terms  of  proton  transfer.  The  main 
value  of  this  view  is  that  it  is  applicable  to  a 
wider  range  of  chemical  systems,  including  non- 
aqueous systems. 

We  generalize  our  view  of  the  acid-base  type 
of  reaction  as  follows.  In  our  example,  reaction 
(50), 


CH3COOH  +  NH3 

an  acid  a  base 


NH4+  +  CH3COO-     (50) 

an  acid  a  base 


The  acetic  acid  reacts  as  an  acid,  giving  up  its 
proton,  to  form  acetate,  CH3COO~,  a  substance 
that  can  act  as  a  base.  We  can  write  (50)  in  a 
general  form: 


HB!  +    B2 
Acidi  -f  Base2 


HB2  +     Bi 
Acid2  +  Basex 


(52) 
(53) 


We  see  that  an  acid  and  a  base  react,  through 
proton  transfer,  to  form  another  acid  and  another 
base* 

We  can  use  this  more  general  view  to  discuss 
the  strengths  of  acids.  In  our  generalized  acid- 
base  reaction  (52),  the  proton  transfer  implies 
the  chemical  bond  in  HBi  must  be  broken  and 
the  chemical  bond  in  HB2  must  be  formed.  If  the 
HB!  bond  is  easily  broken,  then  HBi  will  be  a 
strong  acid.  Then  equilibrium  will  tend  to  favor 
a  proton  transfer  from  HBi  to  some  other  base, 
B2.  If,  on  the  other  hand,  the  HB!  bond  is  ex- 
tremely stable,  then  this  substance  will  be  a  weak 
acid.  Equilibrium  will  tend  to  favor  a  proton 
transfer  from  some  other  acid,  HB2,  to  base  Bi, 
forming  the  stable  HBi  bond. 

11-3.4    Hydronium  Ion  in  the  Proton  Transfer 
Theory  off  Acids 

In  the  proton  transfer  view  of  acid-base  reac- 
tions, an  acid  and  a  base  react  to  form  another 
acid  and  another  base.  Let  us  see  how  this  theory 
encompasses  the  elementary  reaction  between 
H+(aq)  and  OH~(aq)  and  the  reaction  of  disso- 

*  This  more  general  view  of  acids  and  bases  is  named 
the  Bronsted-Lowry  theory  after  the  two  scientists  who 
proposed  it,  J.  N.  Brpnsted  and  T.  M.  Lowry. 


ciation  of  acetic  acid,  reactions  (54)  and  (55): 

H+(aq)  +  OH(aq)  q=*z  H20  (54) 

CH3COOHfaqj  z^r 

H+(aq)  +  CH3COO-f  aqj     (55) 

It  does  so  by  making  a  specific  assumption 
about  the  nature  of  the  species  H+(aq).  It  is 
considered  to  have  the  molecular  formula 
H30+(aq).  Thus,  when  HC1  dissolves  in  water, 
the  reaction  is  written 


HCl(g)  +  H20 
instead  of 
HCl(g) 


H3Q+(aq)  +  Cl-(aq)     (56) 


H+(aq)      +C\-(aq)     (57) 


Whenever  H+(aq)  might  appear  in  an  equation 
for  a  reaction,  it  is  replaced  by  the  hydronium 
ion,  H30+,  and  a  molecule  of  water  is  added  to 
the  other  side  of  the  equation.  We  write  (55)  in 
the  form 


CH3COOH(aqj  +  H20  :p± 

H3Q+(aq)  +  CH3COO-(aq) 


(58) 


Now  the  dissociation  of  acetic  acid  can  be  re- 
garded as  an  acid-base  reaction.  The  acid 
CH3COOH  transfers  a  proton  to  the  base  H20 
forming  the  acid  H30+  and  the  base  CH3COO_. 
The  reaction  (54)  now  takes  the  form 


H3Q+(aq)  +  OH-(aq) 


H20  +  H20    (59) 


In  (59)  the  acid  H30+  transfers  a  proton  to  the 
base  OH-,  forming  an  acid,  H20,  and  a  base, 
H20.  We  see  that  within  the  proton  transfer 
theory,  the  molecule  H20  must  be  assigned  the 
properties  of  an  acid  and,  as  well,  those  of  a  base. 
This  designation  of  the  species  H+(aq)  in 
terms  of  hydronium  ion,  H30+,  is  not  necessi- 
tated by  experimental  evidence  that  proves  the 
unique  existence  of  this  molecule,  H30+,  in  dilute 
aqueous  solutions  (see  Section  1 1-2.5).  Neverthe- 
less, the  convenience  of  this  assumption,  as  an 
aid  in  correlating  acid-base  behavior,  amply 
justifies  its  use. 

11-3.5    Contrast  off  Acid-Base  Definitions 

In  this  chapter  we  first  identified  acids  and  bases 
in  aqueous  solution  by  investigating  the  proper- 


SEC.    11-3    I   STRENGTHS    OF    ACIDS 


195 


ties  possessed  by  acid  solutions  and  base  solu- 
tions. By  so  doing,  we  defined  an  acid  in  terms 
of  the  properties  of  solutions  of  acids.  Now  we 
are  explaining  the  behavior  of  an  acid  in  terms 
of  the  process  of  proton  transfer.  Because  this 
explanation  fits  a  large  number  of  experimental 
facts  well  and  conveniently,  it  has  come  into 
common  use.  If  a  chemist  is  asked  what  sub- 
stances are  acids,  he  is  liable  to  refer  to  the  ex- 
planation rather  than  to  the  identifying  proper- 
ties. At  this  point,  he  has  shifted  to  a  new 
definition.  Let  us  compare  these  two  definitions. 
Definition  1 .  An  acid  is  a  substance  that  has 
the  properties  listed  below  when  dissolved  in 
water: 

it  is  an  electrical  conductor; 
it  reacts  with  Zn  to  give  H2(g); 
it  makes  litmus  red; 
it  tastes  sour. 

Definition  2.  An  acid  is  a  substance  that  can 
release  protons. 

The  first  definition  is  of  the  type  called  an 
"operational  definition."  To  understand  this 
term,  consider  the  intended  meaning  of  the  word 
"definition."  According  to  the  dictionary,  "defi- 
nition" means  "a  statement  of  what  a  thing  is." 
With  a  definition  to  help,  it  is  possible  to  sort 
the  universe  into  two  piles — one  pile  containing 
those  objects  that  fit  the  definition,  and  another 
pile  containing  those  that  do  not.  The  "state- 
ment" gives  criteria  by  which  this  sorting  can  be 
carried  out.  An  "operational  definition"  is,  then, 
a  definition  that  lists,  as  criteria,  measurements 
or  observations  (that  is,  "operations")  by  which 
you  could  decide  whether  a  given  object  is  "in" 
or  "out." 

The  second  definition  is  a  "conceptual  defini- 
tion." It  defines  the  class  in  terms  of  an  expla- 
nation of  why  the  class  has  its  properties.  To  see 
the  difference — and  the  relative  merit — of  the 
two  kinds  of  definition,  let  us  consider  an  anal- 
ogy. 

Contrast  the  following  pair  of  definitions. 

(1)  A  "star"  is  an  athlete  who  regularly  scores 


an  unusually  large  number  of  points  or  who, 
in  clever  defensive  acts,  repetitiously  and  ad- 
vantageously contributes  to  the  welfare  of 
his  team. 
(2)  A  "star"  is  an  athlete  with  unusual  muscular 
coordination. 

The  first  definition  is  "operational."  It  ex- 
plicitly states  criteria  for  deciding  whether  a 
player  is  a  "star."  He  is  one  who  "regularly 
scores  an  unusually  large  number  of  points"  or 
who  "in  clever  defensive  acts,  repetitiously 
and.  •  •  -"To  decide  if  a  player  is  a  "star,"  you 
count  his  scoring  or,  alternatively,  number  the 
outstanding  defensive  plays  he  makes. 

The  second  definition  might  be  called  a  "con- 
ceptual definition."  It  offers  an  explanation  of 
why  the  athlete  enjoys  special  success  in  his 
sport.  It  is  one  step  removed  from  things  that 
show  on  the  scoreboard. 

The  first  definition  gives  no  clue  why  one 
player  is  a  "star"  and  another  player  is  not.  Its 
value  is  that  it  is  practical,  useful,  and  surely 
correct.  It  does  not  matter  how  awkward  a 
basketball  player  is;  if  he  scores  30  points  per 
game  he  is  assured  of  a  considerable  amount  of 
public  acclaim. 

The  value  of  the  second  (conceptual)  definition 
is  that  it  contains  more  information  about  "star- 
dom." If  accurate,  it  has  the  deeper  significance. 
It  might  help  the  basketball  coach  more  in  de- 
veloping the  optimum  characteristics  of  his 
squad.  It  permits  him  to  predict  athletic  skill  in 
advance  of  the  first  game. 

Returning  to  our  two  definitions  of  an  acid, 
the  first,  the  operational  definition,  gives  clearcut 
instructions  on  how  to  decide  whether  a  given 
substance  is  an  acid.  Dissolve  it  in  water  and  see 
if  it  has  certain  properties.  The  second  (concep- 
tual) definition,  however,  has  the  deeper  sig- 
nificance since  it  includes  our  knowledge  of  why 
an  acid  has  these  particular  properties.  It  pro- 
vides a  basis  for  finding  hidden  likenesses  be- 
tween acid-base  reactions  in  water  and  other 
reactions  in  other  solvents.  Each  type  of  defini- 
tion has  its  merit ;  neither  is  the  definition. 


196 


AQUEOUS    ACIDS    AND    BASES    I    CHAP.    11 


QUESTIONS  AND  PROBLEMS 


1 .  What  is  the  concentration  of  H+(aq)  in  an  aque- 
ous solution  in  which  [OH~]  =  1.0  X  10~3  Ml 

2.  100  ml  of  the  HC1  solution  described  in  Exercise 
11-2  (p.  182)  is  diluted  with  water  to  1.00  liter. 
What  is  the  concentration  of  H+(aq)l  What  is 
[OH-]  in  this  solution? 

3.  Vinegar,  lemon  juice,  and  curdled  milk,  all  taste 
sour.  What  other  properties  would  you  expect 
them  to  have  in  common? 

4.  Give  the  name  and  formula  of  three  hydrogen 
containing  compounds  that  are  not  classified  as 
acids.  State  for  each  compound  one  or  more 
properties  common  to  acids  that  it  does  not 
possess. 

5.  As  a  solution  of  barium  hydroxide  is  mixed  with 
a  solution  of  sulfuric  acid,  a  white  precipitate 
forms  and  the  electrical  conductivity  decreases 
markedly.  Write  equations  for  the  reactions  that 
occur  and  account  for  the  conductivity  change. 

6.  An  eyedropper  is  calibrated  by  counting  the 
number  of  drops  required  to  deliver  1.0  ml. 
Twenty  drops  are  required. 

(a)  What  is  the  volume  of  one  drop? 

(b)  Suppose  one  such  drop  of  0.20  M  HC1  is 
added  to  100  ml  of  water.  What  is  [H+]  ? 

(c)  By  what  factor  did  [H+]  change  when  the 
one  drop  was  added? 

Answer,  (c)  X  1000. 

7.  Suppose  drops  (from  the  same  eyedropper)  of 
0.10  M  NaOH  are  added,  one  at  a  time,  to  the 
100  ml  of  HC1  in  Problem  6b. 

(a)  What  will  be  [H+]  after  one  drop  is  added? 

(b)  What  will  be   [H+]   after  two  drops  are 
added? 

(c)  What  will  be  [H+]  after  three  drops  are 
added? 

8.  Calculate  [H+]  and  [OH-]  in  a  solution  made 
by  mixing  50.0  ml  0.200  M  HC1  and  49.0  ml 
0.200  M  NaOH. 

Answer.  [OH"]  =  5  X  10"12  M. 

9.  Calculate  [H+]  and  [OH-]  in  a  solution  made  by 
mixing  50.0  ml  0.200  M  HC1  and  49.9  ml 
0.200  M  NaOH. 


10.  How  much  more  0.200  M  NaOH  solution  need 
be  added  to  the  solution  in  Problem  9  to  change 
[H+]  to  10~7  Ml 

11.  An  acid  is  a  substance  HB  that  can  form  H+(aq) 
in  the  equilibrium : 

HB(aq)  *=±:  H+(aq)  +  B~{aq) 

(a)  Does  equilibrium  favor  reactants  or  products 
for  a  strong  acid? 

(b)  Does  equilibrium  favor  reactants  or  prod- 
ucts for  a  very  weak  acid? 

(c)  If  acid  HBi  is  a  stronger  acid  than  acid  HB2, 
is  Kx  a  larger  or  smaller  number  than  K*l 

m  [H+J[Z?r]        K  =  [H+][B2~] 

12.  (a)  Which  of  the  following  acids  is  the  strongest 

and  which  is  the  weakest  ? 

ammonium  ion,  NH/  (in  an  NH4C1 

solution); 
bisulfate  ion,  HS04"  (in  a  KHS04 

solution); 
hydrogen  sulfide,  H2S. 
(b)  If  0.1  M  solutions  are  made  of  NH»C1, 
KHSO4,  and  H£,  in  which  will  [H+]  be 
highest  and  in  which  will  it  be  lowest? 

13.  (a)  Nitric  acid  is  a  very  strong  acid.  What  is 

[H+]  in  a  0.050  M  HNO3  solution? 
(b)  Hydrogen  peroxide,  H202,  is  a  very  weak 
acid.  What  is  [H202]  in  a  0.050  M  H202 
solution? 

14.  From  a  study  of  Appendix  2,  what  generaliza- 
tion can  you  make  concerning  acids  which  con- 
tain more  than  one  atom  of  hydrogen  in  their 
molecules  or  ions? 

15.  A  0.25  M  solution  of  benzoic  acid  (symbolize  it 
HB)  is  found  to  have  a  hydrogen  ion  concentra- 
tion [H+]  =  4  X  10-»  M. 

(a)  Assuming  the  simple  reaction  HB(aq)  +± 
H+(aq)  +  B~(aq),  calculate  KA  for  benzoic 
acid. 

(b)  Compare  the  values  of  [HB],  [H+],  [B~], 
and  KA  used  in  this  problem  to  the  corre- 
sponding quantities  in  the  benzoic  acid  cal- 
culation presented  in  the  text,  Section  1 1-3.2. 


QUESTIONS    AND    PROBLEMS 


197 


16.  If  23  grams  of  formic  acid,  HCOOH,  are  dis- 
solved in  10.0  liters  of  water  at  20°C,  the  [Hi 
is  found  to  be  3.0  X  10  3  M.  Calculate  KA. 

17.  A  chemist  dissolved  25  grams  of  CH3COOH  in 
enough  water  to  make  one  liter  of  solution.  What 
is  the  concentration  of  this  acetic  acid  solution? 
What  is  the  concentration  of  Hi  aq>?  Assume 
a  negligible  change  in  [CH3COOH]  because  of 
dissociation  to  H(aq). 

18.  When  sodium  acetate,  CH3COONa,  is  added  to 
an  aqueous  solution  of  hydrogen  fluoride,  HF, 
a  reaction  occurs  in  which  the  weak  acid  HF 
loses  H+. 

(a)  Write  the  equation  for  the  reaction. 

(b)  What  weak  acid  is  competing  with  HF  for 
H+? 

Answer,  (b)  CH3COOH,  acetic  acid. 

19.  (a)  Write  the  equation  for  the  reaction  that 

shows  the  acid-base  reaction  between  hy- 
drogen sulfide,  H2S,  and  carbonate  ion, 
C03  2. 

(b)  What  are  the  two  acids  competing  for  H+? 

(c)  From  the  values  of  KA  for  these  two  acids 
(see  Table  11  -IV),  predict  whether  the  equi- 
librium favors  reactants  or  products. 

Answer,  (c)  Products. 


20.  Write  the  equations  for  the  reaction  between 
each  of  the  following  acid-base  pairs.  For  each 
reaction,  predict  whether  reactants  or  products 
are  favored  (using  the  values  of  KA  given  in 
Appendix  2). 

(a)  HNO/aqj  +  NH3f«f,J  +±: 

(b)  NH4+(ac/J  +  F(aq)  zp± 

(c)  C6H6COOH(«rjJ  +  CHzCOO(aq)  +± 

Answer,  (a)  HNChfaqJ  +  NH3faqJ 

q=±  NOt-(aq)  +  NHt+(aq). 
Products,  N02 ~(aq)  and 
NH4+((iqJ  favored. 

21.  Write  the  equations  for  the  reactions  between 
each  of  the  following  acid  base  pairs.  For  each 
reaction,  predict  whether  reactants  or  products 
are  favored. 

(a)  H:SO,(aq)  +  HC03(aq)  +=± 

(b)  HCOztaq)  +  S03"laq)  +± 
(C)   HS03(aq)  +  SO3  2(aq)  Zf±: 

22.  If  the  p\\  of  a  solution  is  5,  what  is  [H+]  ?  Is  the 
solution  acidic  or  basic? 

23.  What  is  [H+]  in  a  solution  of  p\\  =  8?  Is  the 
solution  acidic  or  basic?  What  is  [OH-]  in  the 
same  solution? 

24.  Devise  an  operational  and  also  a  conceptual 
definition  of  a  gas. 


SVANTE  AUGUST  ARRHENIUS  1859-192) 


During  a  full  life  this  great  Swedish  chemist  met  practically 
all  the  important  men  of  science  of  his  day  and  won  their 
affection  as  well  as  their  highest  regard.  He  is  said  to  have 
had  a  genius  for  friendship.  Nevertheless,  his  early  career 
was  filled  with  a  battle  for  acceptance. 

At  22  Arrhenius  had  performed  many  experiments  con- 
cerned with  the  passage  of  electricity  through  aqueous 
solutions  and  he  decided  to  continue  this  work  in  prepara- 
tion for  his  doctorate.  For  two  years  he  collected  voluminous 
data  on  hundreds  of  solutions  and  concentrations  while 
working  in  the  laboratory  of  the  University  of  Upsala.  He 
then  formulated  a  carefully  considered  hypothesis  that 
aqueous  solutions  contain  charged  species,  ions.  This  was  a 
revolutionary  suggestion  and  his  professors  found  it  so 
different  from  their  own  ideas  that  they  only  grudgingly 
granted  his  degree. 

Undiscouraged,  Arrhenius  sent  copies  of  his  thesis  to 
other  scientists.  Although  few  took  his  radical  idea  seri- 
ously, the  great  German  scientist,  Ostwald,  became  so 
excited  that  he  traveled  to  Sweden  to  meet  Arrhenius. 
Encouraged  by  this  support,  Arrhenius  traveled  and  studied 
in  Germany  and  Holland.  Finally,  in  1889,  his  paper  "On 
the  Dissociation  of  Substances  in  Aqueous  Solutions"  was 
published. 

He  was  invited  to  come  to  Leipzig  as  a  professor  at  the 


University  but  chose  to  return  to  Sweden  as  a  lecturer  and 
teacher  at  a  high  school  in  Stockholm.  His  theory  was  still 
not  generally  accepted,  and  those  who  did  not  agree  with  it 
dubbed  its  proponents  "the  wild  horde  of  the  Ionians." 
Even  Arrhenius'  assignment  as  professor  at  Stockholm  in 
1893  was  questioned  until  a  storm  of  protest  came  from 
scientists  in  Germany.  Within  two  years  of  this  appoint- 
ment he  was  elected  President  of  the  University  and  was 
named  a  Nobel  laureate,  the  award  being  only  the  third 
such  in  chemistry.  Arrhenius  finally  received  the  acclaim  he 
had  so  long  deserved.  He  was  offered  the  coveted  position 
of  professor  of  chemistry  at  Berlin  but  the  King  of  Sweden 
founded  the  Nobel  Institute  for  Physical  Chemistry  and,  in 
1905,  Arrhenius  became  its  director.  He  continued  as  a 
tireless  experimenter  and  an  extremely  versatile  scientist 
until  his  death  in  1927. 

Arrhenius'  success  in  science  must  be  credited  not  only 
to  his  brilliance  as  a  scientist  but  also  to  his  conviction  in 
his  views.  His  understanding  of  the  electrical  properties  of 
aqueous  solutions  was  so  far  ahead  of  contemporary  thought 
that  it  would  have  been  ignored  but  for  his  confidence  in 
the  usefulness  of  his  theory  and  his  refusal  to  abandon  it. 
It  is  fitting  tribute  that  the  ionic  model  of  aqueous  solutions 
has  changed  permanently  the  face  of  inorganic  chemistry. 


CHAPTER 


12 


Oxidation- Reduction 
Reactions 


Chemical  thermodynamics  enables  one  to  state  what  may  happen  when 
two  substances  react. 

WENDELL     M.     LATIMER,     1953 


We  have  now  made  use  of  the  principles  of  equi- 
librium in  two  general  types  of  reactions.  In 
Chapter  10  we  considered  reactions  involving  a 
solid  and  a  solution:  dissolving  and  precipita- 
tion. In  Chapter  1 1  we  turned  to  reactions  occur- 


ring entirely  in  solution  and  involving  proton 
transfer.  Now  we  shall  take  a  more  general  view 
of  equilibrium  in  aqueous  solutions,  a  view  pro- 
vided by  an  investigation  of  the  chemistry  of  an 
electrochemical  cell. 


12-1    ELECTROCHEMICAL  CELLS 

Electrochemical  cells  are  familiar — a  flashlight 
operates  on  current  drawn  from  electrochemical 
cells  called  dry  cells,  and  automobiles  are  started 
with  the  aid  of  a  battery,  a  set  of  electrochemical 
cells  in  tandem.  The  last  time  you  changed  the 
dry  cells  in  a  flashlight  because  the  old  ones  were 
"dead,"  did  you  wonder  what  had  happened 
inside  those  cells?  Why  does  electric  current  flow 
from  a  new  dry  cell  but  not  from  one  that  has 
been  used  many  hours?  We  shall  see  that  this  is 
an  important  question  in  chemistry.  By  studying 
the  chemical  reactions  that  occur  in  an  electro- 
chemical cell  we  discover  a  basis  for  predicting 
whether  equilibrium  in  a  chemical  reaction  fa- 


vors reactants  or  products.  The  reactions  are  of 
the  type  called  oxidation-reduction  reactions, 
which  is  the  subject  of  this  chapter. 

12-1.1    The  Chemistry  of  an  Electrochemical 
Cell 

Let's  begin  our  investigation  of  an  electrochemi- 
cal cell  by  assembling  one.  Fill  a  beaker  with  a 
dilute  solution  of  silver  nitrate  (about  0.1  M  will 
do)  and  another  beaker  with  dilute  copper  sul- 
fate. Put  a  silver  rod  in  the  AgN03  solution  and 
a  copper  rod  in  the  CuS04  solution.  With  a  wire, 
connect  the  silver  rod  to  one  terminal  of  an 

199 


200 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.     12 


im:  Cu.so4 


.  I  JvC    CuS04 


Fig.  12-1.  An  electrochemical  cell. 


SEC.     12-1    I    ELECTROCHEMICAL    CELLS 


201 


ammeter  to  measure  the  electric  current.  Con- 
nect the  other  terminal  of  the  ammeter  through 
a  wire  resistance,  R,  to  the  copper  rod.  Finally, 
connect  the  two  solutions  to  complete  the  elec- 
tric circuit.  Figure  12-1  shows  suitable  equip- 
ment. These  two  drawings  show  how  a  connec- 
tion can  be  made  between  the  two  solutions  to 
complete  the  electric  circuit.  A  glass  tube  con- 
taining a  sodium  nitrate  solution  furnishes  an 
electrical  path.  It  is  called  a  salt  bridge. 

As  soon  as  the  last  connection  is  made,  things 
start  to  happen.  The  ammeter  needle  deflects — 
electric  current  is  moving  through  the  meter  and 
the  wire  resistance,  R.  The  direction  of  current 
flow  is  that  of  electrons  moving  from  the  copper 
rod  to  the  silver  rod.  The  resistance  becomes 
warm — the  cell  is  doing  work  as  it  forces  elec- 
trons through  R.  In  the  beakers,  the  copper  rod 
dissolves  and  the  silver  rod  grows.  As  time  goes 
by,  the  ammeter  shows  less  and  less  current  flow 
until,  finally,  there  is  none. 

Now  let's  be  more  quantitative.  Let's  repeat 
the  experiment,  weighing  the  metal  rods  before 
and  after  the  test.  The  weighing  shows  that  dur- 
ing the  test  the  copper  rod  has  become  0.635 
gram  lighter  and  the  silver  rod  has  become  2.16 
grams  heavier.  Chemical  reaction  has  occurred 
and,  as  any  good  chemist  will  do,  we  imme- 
diately ask,  "How  many  moles  of  copper  and 
silver  are  involved?" 


Moles  Cu  dissolved  = 


wt  Cu  dissolved 


atomic  wt  Cu 

=      0.635  g 
63.5  g/mole 

=  0.0100  mole 

Moles  Ag  deposited  =  wt  Ag  deposited 
atomic  wt  Ag 

2.16  g 
108  g/mole 

=  0.0200  mole 

We  see  that  there  is  a  simple  relationship  between 
the  weight  of  copper  dissolved  and  the  weight  of 
silver  deposited.  One  mole  of  copper  dissolves  in 
the  right  beaker  for  every  two  moles  of  silver 
deposited  in  the  left  beaker.  Copper  ions, 
Cu+2(aq),  are  formed  in  the  right  beaker  from 


the  neutral  copper  metal  atoms.  This  means 
atoms  of  copper  release  electrons  into  the  copper 
rod.  These  electrons  move  into  the  wire,  through 
the  resistance,  and  through  the  ammeter.  They 
arrive  at  the  silver  rod  in  the  left  beaker,  where 
silver  metal  is  formed  from  silver  ions,  Ag+(aq). 
Here,  the  positive  silver  ions  draw  electrons  from 
the  silver  rod  to  become  neutral  silver  metal 
atoms.  Summarizing  these  processes,  we  have: 

In  the  right  beaker,  Cu(s)  — >- 

Oi^(aq)  +  2e~    (7) 

In  the  left  beaker,  2Ag+(aqj  +  2e~  — >-  2Ag(s)    (2) 

Overall  reaction,  Cu(s)  +  2Ag+(aq)   — ►■ 

2Ag(s)  +  Oi-"(aq)    (5) 

The  overall  reaction  describes  what  goes  on  in 
the  entire  electrochemical  cell.  In  half  of  the  cell, 
the  right  beaker,  reaction  (7)  occurs.  In  the  other 
half  of  the  cell,  the  left  beaker,  reaction  (2) 
occurs.  Hence,  reactions  (7)  and  (2)  are  called 
half-cell  reactions  or  half-reactions. 

There  are  several  interesting  features  about 
these  half-reactions: 

(1)  The  two  half-reactions  are  written  separately. 
In  our  electrochemical  cell  the  half-reactions 
occur  in  separate  beakers.  As  the  name  im- 
plies, there  must  be  two  such  reactions. 

(2)  Electrons  are  shown  as  part  of  the  reaction. 
Our  ammeter  shows  that  electrons  are  in- 
volved. They  flow  when  the  reaction  starts, 
and  do  not  flow  when  the  reaction  stops.  The 
meter  also  indicates  that  the  electrons  leave 
the  copper  rod,  pass  through  the  wire,  and 
enter  the  silver  rod. 

(3)  New  chemical  species  are  produced  in  each 
half  of  the  cell.  The  copper  rod  is  converted 
to  copper  ions  (the  rod  loses  weight)  and  the 
silver  ions  are  changed  to  metal  (the  silver 
rod  gains  weight).  The  new  species  can  be 
explained  in  terms  of  gain  of  electrons  (by 
silver)  and  loss  of  electrons  (by  copper). 

(4)  The  half-reactions,  when  combined,  express 
the  overall,  or  net,  reaction. 

The  net  reaction  (3)  is  obtained  by  combining 
(7)  and  (2)  so  as  to  cause  the  exact  balancing  of 
electrons  lost  by  copper  atoms,  in  (7),  and  elec- 
trons gained  by  silver  ion,  in  (2).  This  cancella- 


202 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.    12 


tion  is  necessary  because  electrical  measurements 
show  that  the  electrochemical  cell  operates  with- 
out accumulation  or  consumption  of  electric 
charge.  The  reaction  mixture  always  remains 
electrically  neutral.  The  number  of  electrons  lost 
equals  the  number  of  electrons  gained. 

We  see  that  the  overall  chemical  reaction  that 
occurs  in  an  electrochemical  cell  is  conveniently 
described  in  terms  of  two  types  of  half-reactions. 
In  one,  electrons  are  lost;  in  the  other,  they  are 
gained.  To  distinguish  these  half-reactions  we 
need  two  identifying  names. 

The  half-reaction  in  which  electrons  are  lost  is 
called  oxidation. 

Oxidation       Cu(s)  — >-  Cu^(aq)  +  2e~    (/),  (4)* 

The  half-reaction  in  which  electrons  are  gained 
is  called  reduction. 


Reduction       2Ag+(aq)  +  2e~ 


2Ag(s)    (2),  (5) 


The  overall  reaction  is  called  an  oxidation- 
reduction  reaction. 

Oxidation-reduction  reaction 


Cu(s)  +  2Ag+(aq) 


Cu+*(aq)  +  2Ag(s)    (J),  (6) 


It  is  often  convenient  and  usually  informative 
to  treat  oxidation-reduction  in  terms  of  half- 
reactions.  When  it  is  convenient,  oxidation  is  in- 
volved in  the  half-reaction  showing  loss  of  elec- 
trons, and  reduction  is  involved  in  the  half- 
reaction  showing  gain  of  electrons. 

12-1.2    Oxidation-Reduction  Reactions 
in  a  Beaker 

These  ideas,  developed  for  an  electrochemical 
cell,  have  great  importance  in  chemistry  because 
they  are  also  applicable  to  chemical  reactions 
that  occur  in  a  single  beaker.  Without  an  elec- 
tric circuit  or  an  opportunity  for  electric  cur- 
rent to  flow,  the  chemical  changes  that  occur  in 
a  cell  can  be  duplicated  in  a  single  solution.  It  is 
reasonable  to  apply  the  same  explanation. 

*  In  this  and  later  chapters,  each  equation  is  assigned 
a  consecutive  number  (given  on  the  right).  If  the  equation 
has  occurred  earlier  in  the  chapter,  the  earlier  number  is 
given  as  well  (on  the  left). 


COPPER    OXIDIZED    BY    Ag+(aq) 
IN    A    BEAKER 

We  can  easily  demonstrate  that  reaction  (J)  can 
occur  even  when  the  half-cells  are  not  separated. 
You  did  this  in  Experiment  7.  A  copper  wire 
immersed  in  AgN03  solution  caused  copper  to 
dissolve  [the  blue  color  of  Cu+2(aq)  appeared] 
and  metallic  silver  was  precipitated.  The  ratio  of 
(Cu  dissolved)  to  (Ag  formed)  was  the  same  as 
that  in  our  cell,  hence  the  net  reaction  was  the 
same. 


EXERCISE  12-1 

Compare  the  mole  ratio  Ag/Cu  derived  from 
your  own  data  for  Experiment  7  to  the  electro- 
chemical data  given  in  Section  12-1.1. 


The  moles  of  silver  deposited  per  mole  of  cop- 
per dissolved  are  the  same  whether  reaction  (3) 
is  carried  out  in  an  electrochemical  cell  or  in  a 
single  beaker,  as  in  Experiment  7.  If,  in  the  cell, 
electrons  are  transferred  from  copper  metal 
(forming  Cu+2)  to  silver  ion  (forming  metallic 
silver),  then  electrons  must  have  been  transferred 
from  copper  metal  to  silver  ion  in  Experiment  7. 


Fig.  12-2.  Oxidation-reduction    reactions    can    occur 
in  a  beaker. 


Cu(s)-+  Cufaf)  +  2e 
oxt'dairion 

2Af(aqh2e--+  2A9  (s) 
reduction. 


SEC.    12-1    I    ELECTROCHEMICAL    CELLS 


203 


Thus,  Experiment  7  involved  the  same  oxidation- 
reduction  reaction  but  the  electron  transfer  must 
have  occurred  locally  between  individual  copper 
atoms  (in  the  metal)  and  individual  silver  ions 
(in  the  solution  near  the  metal  surface).  This 
local  transfer  replaces  the  wire  "middleman"  in 
the  cell,  which  carries  electrons  from  one  beaker 
(where  they  are  released  by  copper)  to  the  other 
(where  they  are  accepted  by  silver  ions). 

ZINC    OXIDIZED    BY    U+(aq)    IN    A    BEAKER 

Many  oxidation-reduction  reactions  (nicknamed 
"redox"  reactions)  take  place  in  aqueous  solu- 
tion. One  of  these  was  mentioned  in  Section 
11-2.1  when  we  characterized  acids: 

Zn(s)  +  2H+(aq)  — ►-  Zn^-(aq)  +  HJg)    (7) 

Each  zinc  atom  loses  two  electrons  in  changing 
to  a  zinc  ion,  therefore  zinc  is  oxidized.  Each 
hydrogen  ion  gains  an  electron,  changing  to  a 
hydrogen  atom,  therefore  hydrogen  is  reduced. 
(After  reduction,  two  hydrogen  atoms  combine 
to  form  molecular  H2.)  As  before,  reaction  (7) 
can  be  separated  into  two  half-reactions: 

Zn(s)  — ►-  Zn**(aq)  +  2e~  (8) 

2H+(aq)  +  2e~  — >  HJg)*  (9) 

Net  reaction 


Zn(s)  +  2H+ 


Zn+2  +  H2(g)        (7),  (10) 


Thus,  the  reaction  by  which  a  metal  dissolves  in 
an  acid  is  conveniently  discussed  in  terms  of  oxi- 
dation and  reduction  involving  electron  transfer. 
The  reaction  can  be  divided  into  half-reactions 
to  show  the  electron  gain  (by  H+  ions)  and  the 
electron  loss  (by  metal  atoms). 

Not  all  metals  react  with  aqueous  acids. 
Among  the  common  metals,  magnesium,  alumi- 
num, iron,  and  nickel  liberate  H2  as  zinc  does. 
Other  metals,  including  copper,  mercury,  silver, 
and  gold,  do  not  produce  measurable  amounts 
of  hydrogen  even  though  we  make  sure  that  the 
equilibrium  state  has  been  attained.  With  these 
metals,  hydrogen  is  not  produced  and  it  is  surely 
not  just  because  of  slow  reactions.  Apparently 

*  From  this  point  on  in  this  chapter  we  will  consider 
only  aqueous  solutions,  hence  we  will  not  specify  (aq) 
for  each  ion. 


Copper 
me"tal 


1M  HCl 


Fig.  12-3.  Some  metals  release  electrons  to  H*  and 
others  do  not. 


some  metals  release  electrons  to  H+  [as  zinc  does 
in  reaction  (70)]  and  others  do  not. 

ZINC    OXIDIZED    BY    C\l+2(aq) 
IN    A    BEAKER 

As  a  third  oxidation-reduction  example,  suppose 
a  strip  of  metallic  zinc  is  placed  in  a  solution  of 
copper  nitrate,  Cu(N03)2.  The  strip  becomes 
coated  with  reddish  metallic  copper  and  the 
bluish  color  of  the  solution  disappears.  The 
presence  of  zinc  ion,  Zn+2,  among  the  products 
can  be  shown  when  the  Cu+2  color  is  gone.  Then 
if  hydrogen  sulfide  gas  is  passed  into  the  mixture, 
white  zinc  sulfide,  ZnS,  can  be  seen.  The  reaction 
between  metallic  zinc  and  the  aqueous  copper 
nitrate  is 


Zn(s)  +  Cu+2 


Zn+2  +  Cu(s) 


VD 


Zinc  has  lost  electrons  in  reaction  (77)  to  form 
Zn+2: 

Zn(s)  — *~  Zn+2  +  2e-  (8),  (12) 

Zinc  is  oxidized.  If  zinc  is  oxidized,  releasing 
electrons,  something  must  be  reduced,  accepting 
these  electrons.  Copper  ion  is  reduced: 


Cu+2  +  2e- 


Cu(s) 


U3) 


This  time,  copper  ion  gains  electrons  from  the 
zinc,  in  contrast  to  the  behavior  in  Experiment  7, 
where  copper  metal  lost  electrons  to  silver. 


204 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.     12 


j  .1M  Cu.S04 


Cu,  (s)  — —  CuT   +  2i 


.llvf  Zn  (wo3)2 


.lKf  CilSO+ 


Cu      +  2e-^-~Cu.(s) 


Fig.  12-4.  Two  electrochemical  cells  involving  copper:  with  silver,  copper  is  oxidized;  with  zinc,  Cu*2  is  reduced. 


SEC.     12-1    I    ELECTROCHEMICAL    CELLS 


205 


What  about  the  state  of  equilibrium  for  the 
reaction  represented  by  equation  (77)?  Let  us 
place  a  strip  of  metallic  copper  in  a  zinc  sulfate 
solution.  No  visible  reaction  occurs  and  attempts 
to  detect  the  presence  of  cupric  ion  by  adding 
H2S  to  produce  the  black  color  of  cupric  sulfide, 
CuS,  fail.  Cupric  sulfide  has  such  low  solubility 
that  this  is  an  extremely  sensitive  test,  yet  the 
amount  of  Cu+2  formed  cannot  be  detected.  Ap- 
parently the  state  of  equilibrium  for  the  reaction 
(77)  greatly  favors  the  products  over  the  reac- 
tants. 

12-1.3    Competition  for  Electrons 

These  reactions  can  be  viewed  as  a  competition 
between  two  kinds  of  atoms  (or  molecules)  for 
electrons.  Equilibrium  is  attained  when  this  com- 
petition reaches  a  balance  between  opposing  re- 
actions. In  the  case  of  reaction  (3),  copper  metal 
reacting  with  silver  nitrate  solution,  the  Cu(s) 
releases  electrons  and  Ag+  accepts  them  so 
readily  that  equilibrium  greatly  favors  the  prod- 
ucts, Cu+2  and  Ag(s).  Since  randomness  tends  to 
favor  neither  reactants  nor  products,  the  equi- 
librium must  favor  products  because  the  energy 
is  lowered  as  the  electrons  are  transferred.  If  we 
regard  reaction  (3)  as  a  competition  between 
silver  and  copper  for  electrons,  stability  favors 
silver  over  copper. 

The  same  sort  of  competition  for  electrons 
is  involved  in  reaction  (77),  in  which  Zn(s)  re- 
leases electrons  and  Cu+2  accepts  them.  This  time 
the  competition  for  electrons  is  such  that  equi- 
librium favors  Zn+2  and  Cu(s).  By  way  of  con- 
trast, compare  the  reaction  of  metallic  cobalt 
placed  in  a  nickel  sulfate  solution.  A  reaction 
occurs, 

Co(s)  +  Ni+2  +±:  Co+2  +  Ni(s)         (14) 

At  equilibrium,  chemical  tests  show  that  both 
Ni+2  and  Co+2  are  present  at  moderate  concen- 
trations. In  this  case,  neither  reactants  (Co  and 
Ni+2)  nor  products  (Co+2  and  Ni)  are  greatly 
favored. 

This  competition  for  electrons  is  reminiscent 
of  the  competition  for  protons  among  acids  and 
bases.  The  similarity  suggests  that  we  might  de- 
velop a  table  in  which  metals  and  their  ions  are 


listed  by  tendency  to  release  electrons  just  as  we 
did  in  Table  11-IV  (p.  191)  in  which  the  acid 
strength  indicates  tendency  for  an  acid  to  re- 
lease H+. 

We  can  already  make  some  comparisons.  We 
might  begin  by  listing  some  of  the  half-reactions 
we  have  encountered  in  this  chapter.  We  shall 
write  them  to  show  the  release  of  electrons  and 
then  arrange  them  in  order  of  their  tendency  to 
do  so.  First  we  considered  reaction  (5)  and  dis- 
covered that  copper  releases  electrons  to  silver 
ion.  Therefore,  we  shall  write  our  first  two  half- 
reactions  in  the  order 

Cu(s)  — >-  Cu+2  +  2e~  (1),  (15) 

Ag(s)  — >-  Ag+  +  e~  (16) 

Listing  the  Cu-Cu+2  half-reaction  first  indicates 
that  it  releases  electrons  more  readily  than  does 
the  Ag-Ag+  half-reaction. 

Now  consider  reaction  (77).  Since  zinc  releases 
electrons  to  copper  ion,  we  know  that  we  must 
add  it  to  our  list  at  the  top: 

Zn(s)  — *-  Zn+2  +  2e~  (8),  (17) 

Cu(s)  — >-  Cu+2  +  2e~  (1),  (18) 

Ag(s)  —*■  Ag+  +  e~  (16),  (19) 

Listing  the  Zn-Zn+2  half-reaction  first  tells  us 
that  it  releases  electrons  more  readily  than  does 
the  Cu-Cu+2  half-reaction.  But  if  this  is  true, 
then  the  Zn-Zn+2  half-reaction  must  also  release 
electrons  more  readily  than  does  the  Ag-Ag+ 
half-reaction.  Our  list  leads  us  to  expect  that  zinc 
metal  will  release  electrons  to  silver  ion,  reacting 
to  produce  zinc  ion  and  silver  metal. 

We  should  test  this  proposal!  We  dip  a  piece 
of  zinc  metal  in  a  solution  of  silver  nitrate.  The 
result  confirms  our  expectation;  zinc  metal  dis- 
solves and  bright  crystals  of  metallic  silver  ap- 
pear. 

Our  data  allow  us  to  make  one  more  addition 
to  the  list.  By  reaction  (7),  zinc  reacts  with  H+ 
to  give  Zn+2  and  U2(g).  The  half-reaction 
H2-2H+  must  be  placed  below  the  Zn-Zn+2  half- 
reaction.  How  far  below?  To  answer  that,  re- 
member that  copper  does  not  react  with  H+  to 
produce  H2.  This  indicates  that  the  half-reaction 


206 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.    12 


H2-2H+  releases  electrons  more  readily  than 
does  the  half-reaction  Cu-Cu+2.  Now  we  can 
expand  our  list  to  that  given  in  Table  12-1. 

Table  12-1 

SOME  HALF-REACTIONS  LISTED  IN 
ORDER  OF  TENDENCY  TO  RELEASE 
ELECTRONS 


Zn(s) 

Cu(s) 
Agfa,) 


Zn+2  +  2e~ 
2H+  +  2e~ 
Cu+2  +  2e- 
Ag+  +  e~ 


(8),  (20) 

(21) 

(/),  (22) 

(16),  (23) 


EXERCISE  12-2 

From  the  statement  in  the  text  that  nickel  metal 
reacts  with  H+  to  give  H2(g)  and  the  additional 
information  that  zinc  metal  reacts  readily  with 
nickel  sulfate  solution,  decide  where  to  add  the 
half-reaction  Ni-Ni+2  in  our  list. 


The  value  of  this  list  is  obvious.  Any  half- 
reaction  can  be  combined  with  the  reverse  of 
another  half-reaction  (in  the  proportion  for 
which  electrons  gained  is  equal  to  electrons  lost) 
to  give  a  possible  chemical  reaction.  Our  list 
permits  us  to  predict  whether  equilibrium  favors 
reactants  or  products.  We  would  like  to  expand 
our  list  and  to  make  it  more  quantitative.  Elec- 
trochemical cells  help  us  do  this. 

12-1.4    Operation  off  an  Electrochemical  Cell 

Now  let's  take  a  more  detailed  look  into  the 
electrochemical  cell.  Figure  12-5  shows  a  cross- 
section  of  a  cell  that  uses  the  same  chemical  re- 
action as  that  depicted  in  Figure  12-1.  The  only 
difference  is  that  the  two  solutions  are  connected 
differently.  In  Figure  12-1  a  tube  containing  a 
solution  of  an  electrolyte  (such  as  KN03)  pro- 
vides a  conducting  path.  In  Figure  12-5  the  silver 
nitrate  is  placed  in  a  porous  porcelain  cup.  Since 
the  silver  nitrate  and  copper  sulfate  solutions  can 
seep  through  the  porous  cup,  they  provide  their 
own  connection  to  each  other. 


-Porous  cup 

Fig.  12-5.  The  operation  of  an  electrochemical  cell. 


Before  examining  the  processes  in  a  cell,  we 
should  name  the  parts  of  a  cell  and  clear  away 
some  language  matters.  The  electrons  enter  and 
leave  the  cell  through  electrical  conductors — the 
copper  rod  and  the  silver  rod  in  Figure  12-5 — 
called  electrodes.  At  one  electrode,  the  copper 
electrode,  electrons  are  released  and  oxidation 
occurs.  The  electrode  where  oxidation  occurs  is 
called  the  anode.  At  the  other  electrode,  the 
silver  electrode,  electrons  are  gained  and  reduc- 
tion occurs.  The  electrode  where  reduction  oc- 
curs is  called  the  cathode. 

As  electrons  leave  the  cell  from  the  anode 
(electrons  are  released  where  oxidation  occurs), 
positively  charged  Cu+2  ions  are  produced.  Nega- 
tive charge  is  leaving  (by  means  of  the  electron 
movement)  and  positive  charge  is  produced  (the 
Cu+2  ions)  in  this  half  of  the  cell.  How  is  elec- 
trical neutrality  maintained?  It  must  be  main- 


SEC.     12-2    I    ELECTRON    TRANSFER    AND    PREDICTING    REACTIONS 


207 


tained  by  the  movement  of  ions  through  the 
solution.  Negative  ions  drift  toward  the  anode 
and  positive  ions  move  away.  It  is  because  nega- 
tive ions  in  a  cell  always  drift  toward  the  anode 
that  negative  ions  are  called  anions  (pronounced 
an'ions).  Since  positive  ions  drift  away  from  the 
anode  and  toward  the  cathode,  positive  ions  are 
called  cations  (pronounced  cations). 
Here  is  our  electrochemical  glossary: 

Electrodes:  The  conductors  at  which  reactions 
occur  in  an  electrochemical  cell. 

Anode:  The  electrode  at  which  oxidation 

occurs. 

Cathode:  The  electrode  at  which  reduction 
occurs. 

Anion:  Negatively  charged  ion. 

Cation:  Positively  charged  ion. 

With  the  verbal  matters  out  of  the  way,  let's 
take  an  electrical  tour  through  the  cell  shown  in 
Figure  12-5.  We'll  start  at  the  surface  of  the 
copper  rod  and  follow  the  process  around  the 
entire  circuit  and  back  to  the  copper  rod.  Let  us 
begin  with  a  particular  copper  atom  that  loses 
two  electrons: 

Cu(s)  — ►-  Cu+2  +  2e~  (/),  (24) 


The  Cu+2  ion  drifts  away  into  the  solution  but 
the  electrons  remain  in  the  copper  rod.  They 
move  up  through  the  copper  anode,  through  the 
wire,  and  enter  the  silver  cathode.  At  the  surface 
of  this  rod,  the  electrons  encounter  Ag+  ions  in 
the  solution.  The  electrons  react  with  Ag+  ions 
to  give  neutral  silver  atoms  which  remain  on  the 
rod  as  silver  metal: 


2Ag4  +  2e- 


2kg(s)  (2),  (25) 


Now  there  is  an  excess  of  positive  charge  in  the 
solution  near  the  copper  anode  and  a  deficiency 
of  positive  charge  in  the  solution  near  the  silver 
cathode.  Two  negative  charges  have  been  moved 
from  the  anode  half-cell  through  the  wire  to  the 
cathode  half-cell.  This  charge  movement  causes 
all  of  the  negative  ions  (anions)  in  the  solution 
(S04~2  and  N03")  to  start  drifting  toward  the 
anode.  All  of  the  positive  ions  (cations)  start 
drifting  toward  the  cathode.  When  the  movement 
of  all  of  these  ions  amounts  to  the  charge  move- 
ment of  two  negative  charges  from  the  cathode 
porous  cup  to  the  anode  beaker,  our  tour  of  the 
cell  is  completed.  Electrical  neutrality  has  been 
restored  and  the  net  reaction  is 


Cu(s)  +  2Ag+ 


Cu+2  +  2Ag(sj    (3),  (26) 


12-2    ELECTRON  TRANSFER  AND  PREDICTING  REACTIONS 


The  usefulness  of  Table  12-1  is  clear.  Qualitative 
predictions  of  reactions  can  be  made  with  the 
aid  of  the  ordered  list  of  half-reactions.  Think 
how  the  value  of  the  list  would  be  magnified  if 
we  had  a  quantitative  measure  of  electron  losing 
tendencies.  The  voltages  of  electrochemical  cells 
furnish  such  a  quantitative  measure. 

12-2.1    Electron  Losing  Tendency 

The  circuit  shown  in  Figure  12-1  includes  a  wire 
resistance  coil,  R.  As  the  current  flows  through 
R,  heat  is  generated.  The  cell  is  doing  electrical 
work  in  forcing  the  electron  current  through  R. 
Again  we  apply  the  Law  of  Conservation  of 
Energy.  As  the  electrons  leave  R,  they  must  have 
lower  potential  energy  than  they  had  when  they 


entered.  As  they  fall  to  the  lower  potential  en- 
ergy, the  energy  change  appears  as  heat.  This 
potential  energy  change  is  measured  by  voltage. 
Just  as  lowering  a  mass  from  a  higher  altitude 
decreases  its  potential  energy,  moving  an  electric 
charge  to  a  lower  voltage  lowers  its  potential 
energy. 

So  the  voltage  of  an  electrochemical  cell  meas- 
ures its  capacity  for  doing  electrical  work.  Dif- 
ferent cells  show  different  voltages.  To  see  the 
importance  of  this  voltage,  consider  the  experi- 
ment shown  in  Figure  12-6.  In  Figure  12-6A  we 
have  a  cell  based  upon  reaction  (27): 

Zn(s)  +  Ni+2  — ►-  Zn+2  +  NifaJ          (27) 

If  the  concentrations  in  the  cell  are  1  M,  the 
voltage  in  the  cell  is  0.5  volt.  We  will  call  this 


208 


OXIDATION-REDUCTION    REACTIONS    |    CHAP.     12 


Zn 


Zn 


Fie.   12-6.  Two  cells  In  opposition.  In  which  direction  will  current  flow? 


SEC.     12-2    I    ELECTRON    TRANSFER    AND    PREDICTING    REACTIONS 


209 


voltage  E°  (Zn-Ni+2).  As  the  cell  is  shown,  elec- 
trons move  clockwise  through  the  meter.  On  the 
right  in  Figure  12-6B  is  a  cell  based  upon  reac- 
tion (28).  The  voltage  of  this  cell,  E°  (Zn-Ag+), 
is  1.5  volts  for  1  M  concentrations. 


Zn(s)  +  2Ag+ 


Zn+2  +  2Ag(s)        (28) 


The  electrical  hookup  in  this  cell  causes  electrons 
to  move  counterclockwise  through  the  meter. 

Then,  in  Figure  12-6C,  the  two  cells  are  recon- 
nected in  opposition  to  each  other.  The  zinc- 
nickel  cell  pushes  the  electrons  clockwise  and 
the  silver-zinc  cell  pushes  the  electrons  counter- 
clockwise. In  each  cell,  the  zinc  electrode  has  a 
tendency  to  dissolve,  releasing  electrons  into  the 
external  circuit.  But  electrons  cannot  flow  in 
both  directions.  Which  cell  will  win  out?  Ex- 
periment shows  that  electron  current  flows  coun- 
terclockwise from  the  zinc  electrode  of  the 
Zn-Ag+  cell  into  the  zinc  electrode  of  the 
Zn-Ni+2  cell.  We  can  explain  this  intuitively  by 
saying  the  stronger  (1.5  volts)  cell  has  over- 
powered the  weaker  (0.5  volt)  cell.  The  Zn-Ag+ 
cell  proceeds  to  react  in  the  sense  it  would  if  it 
were  alone  in  the  circuit.  The  other  cell  is  forced 
to  react  in  the  direction  opposite  to  that  it  would 
take  spontaneously.  The  reaction  that  generates 
the  higher  voltage  prevails.  The  reaction  which 
generates  the  higher  voltage  has  the  greater  tend- 
ency to  proceed  and  the  observed  voltage  meas- 
ures the  tendency. 

We  can  learn  one  more  valuable  lesson  from 
these  two  cells  working  against  each  other.  The 
measured  voltage  of  our  double  cell  (Figure  12-6) 
is  1.0  volt.  Since  the  voltage  for  each  cell  is  a 
measure  of  the  tendency  to  send  out  electrons, 
we  could  calculate  the  net  voltage  (1.0  volt)  by 
combining  the  individual  cell  voltages.  In  this 
case  we  must  subtract  the  two  because  they  are 
hooked  up  to  oppose  each  other.  So  we  find 

£2et  =  £°(Zn-Ag+)  -  £°(Zn-Ni+2) 
=  1.5  -  0.5 
=  1.0  volt 

It  is  interesting  to  write  an  overall  reaction  for 
these  two  cells  when  they  are  operating  as  in 
Figure  12-6: 

Zn- Ag+  cell : 

Zn(s)  — ►-  Zn«  +  2e~         (8),  (29) 


2Ag+  +  2e~  -+  2Ag(s)  (2),  (30) 

Zn-Ni+2cell: 

Zn+2  +  2e~  — ►-  Zn(s)  (8),  (31) 

NifsJ  — ►-  Ni+2  +  2e~  (32) 

Net  reaction  in  both  cells: 

(29)  +  (30)  +  (31)  +  (32) 

NifsJ  +  2Ag+  — >-  Nr"  +  2AgfsJ  (33) 

But  the  net  reaction  is  just  the  reaction  that  oc- 
curs in  a  Ni-Ag+  cell!  What  is  the  voltage  of 
such  a  cell?  Experiment  shows  that  such  a  cell 
has  a  voltage  of  1 .0  volt.  We  find  our  double  cell 
(Figure  12-6)  has  a  voltage  identical  to  that  of  a 
Ni-Ag+  cell  alone.  The  tendency  of  zinc  to  re- 
lease electrons  through  reaction  (29)  must  have 
influenced  the  voltage  of  the  Zn-Ag+  cell.  The 
same  tendency  of  zinc  to  release  electrons  must 
also  have  influenced  the  voltage  of  the  Zn-Ni+2 
cell.  When  these  cells  are  put  in  opposition,  the 
tendency  of  zinc  to  release  electrons  exactly 
cancels. 

This  shows  that  the  voltage  of  a  given  cell  may 
be  thought  of  as  being  made  up  of  two  parts,  one 
part  characteristic  of  one  of  the  half-reactions 
and  one  part  characteristic  of  the  other  half- 
reaction.  Chemists  call  these  two  parts  "half-cell 
potentials,"  a  term  that  emphasizes  the  relation 
between  voltage  and  potential  energy.  The  half- 
cell  potentials  are  symbolized  E°. 

Thus  we  write  for  the  Zn-Ag+  cell 

E\  =  £°(Zn-Zn+2)  -  £°(Ag-Ag+)  =  1.5  volts    (34) 

and  for  the  Zn-Ni+2  cell 

El  =  £°(Zn-Zn+2)  -  £°(Ni-Ni+2)  =  0.5  volt      (35) 

If  we  place  these  batteries  in  opposition,  the  net 
voltage  will  be 

E°3  =  E\  -  E°2  =  [£°(Zn-Zn+2)  -  £°(Ag-Ag+)] 

-  [£°(Zn-Zn+2)  -  £°(Ni-Ni+2)] 
El  =  ^(Ni-Ni*2)  -  £°(Ag-Ag+)  =  1.0  volt      (36) 

and  also 

El  =  E\  -  El  =  (1.5)  -  (0.5)  =  1.0  volt 

We  see  that: 

(a)  the  value  £°(Zn-Zn+2)  cancels  in  (36); 

(b)  (36)  is  just  the  expected  sum  of  half-reactions 
for  a  Ni-Ag+  cell; 


210 


OXIDATION-REDUCTION    REACTIONS   I   CHAP.    12 


(c)  the  calculated  difference  between  7f°  and  El 
is  just  1.0  volt,  the  same  as  measured  for  a 
Ni-Ag+  cell. 

MEASURING    HALF-CELL    POTENTIALS 

We  would  like  to  measure  the  contribution  each 
half-reaction  makes  to  the  voltage  of  a  cell.  Yet 
every  cell  involves  two  half-reactions  and  every 
cell  voltage  measures  a  difference  between  their 
half-cell  potentials.  We  can  never  isolate  one 
half-reaction  to  measure  its  E°.  An  easy  escape 
is  to  assign  an  arbitrary  value  to  the  potential  of 
some  selected  half-reaction.  Then  we  can  com- 
bine all  other  half-reactions  in  turn  with  this 
reference  half-reaction  and  find  values  for  them 
relative  to  our  reference.  The  handiest  arbitrary 
value  to  assign  is  zero  and  chemists  have  decided 
to  give  it  to  the  half-reaction 


Wg) 


2H+  +  2e~ 


E°  =  0.00  •  •  •  volt    (27),  (37) 


We  must  control  concentration  during  these 
measurements  of  E°,  since  the  voltage  of  a  cell 
changes  as  concentrations  change.  For  example, 
in  the  laboratory  we  studied  a  cell  based  on  re- 
action (38): 

Zn(s)  +  Cu+2  — >-  Zn+2  +  Cu(s)    (77),  (38) 

We  found  that  the  voltmeter  readings  were  dif- 
ferent for  different  concentrations.  Were  these 
readings  in  agreement  with  predictions  we  would 
make  on  the  basis  of  equilibrium  principles? 
Increasing  the  concentration  of  Cu+2  ion  in  re- 
action (38)  should  increase  the  tendency  to  form 
the  products  (Le  Chatelier's  Principle).  Experi- 
mentally, we  find  an  increase  in  voltage  when 
more  solid  copper  sulfate  is  dissolved  in  the  solu- 
tion around  the  copper  electrode.  Conversely, 
decreasing  the  concentration  of  the  copper  ion 
should  decrease  the  tendency  to  form  the  prod- 
ucts. Again  in  agreement,  voltage  readings  de- 
crease when  the  concentration  of  Cu+2  ion  is 
reduced  by  precipitation  of  CuS.  The  voltage 
shows  the  tendency  for  reaction  to  occur. 

What  happens  to  any  cell  or  battery  as  it 
operates?  The  voltage  decreases  until,  finally,  it 
reaches  zero.  Then  we  say  the  cell  is  dead.  Equi- 
librium has  been  attained  and  the  reaction  that 


has  been  producing  the  energy  has  the  same 
tendency  to  proceed  as  does  the  reverse  reaction. 
Again  the  voltage  measures  the  net  tendency  for 
reaction  to  occur.  At  equilibrium  there  is  a  bal- 
ance between  forward  and  reverse  reactions, 
hence  there  is  no  net  tendency  for  further  reac- 
tion either  way.  Therefore,  the  voltage  of  a  cell 
at  equilibrium  is  zero. 

Since  concentration  variations  have  measura- 
ble effects  on  the  cell  voltage,  a  measured  voltage 
cannot  be  interpreted  unless  the  cell  concentra- 
tions are  specified.  Because  of  this,  chemists 
introduce  the  idea  of  "standard-state."  The 
standard  state  for  gases  is  taken  as  a  pressure  of 
one  atmosphere  at  25°C;  the  standard  state  for 
ions  is  taken  as  a  concentration  of  1  M ;  and  the 
standard  state  of  pure  substances  is  taken  as  the 
pure  substances  themselves  as  they  exist  at  25°C. 
The  half-cell  potential  associated  with  a  half- 
reaction  taking  place  between  substances  in  their 
standard  states  is  called  E°  (the  superscript  zero 
means  standard  state).  We  can  rewrite  equation 
(37)  to  include  the  specifications  of  the  standard 
states: 

H2(g,  1  atm)  — >-  2H+(aq,  1  M)  +  2er 

E°  =  0.00  •  •  •  volt  (assigned)    (39) 

Now  if  we  combine  a  Zn-Zn+2  half-cell  in  its 
standard  state  with  a  H2-2H+  half-cell  in  its 
standard  state,  the  voltage  (potential)  we  meas- 
ure (0.76  volt)  is  the  value  assigned  to  the  half- 
reaction: 

Zn(s)  — >-  Zn+2(aq,  1  M)  +  2e~ 

E°  =  +0.76  volt    (40) 

Similarly,  if  we  combine  a  Cu-Cu+2  half-cell  in 
its  standard  state  with  a  standard  H2-2H+  half- 
cell,  the  voltage  (potential)  we  measure  (0.34 
volt)  is  the  value  assigned  to  the  half-reaction: 

Cu(s)  — »-  Cu+^aq,  1  M)  +  2e~ 

E°  =  -0.34  volt    (41) 

Chemists  have  determined  a  large  number  of 
these  half-cell  potentials.  The  magnitude  of  the 
voltage  is  a  quantitative  measure  of  the  tendency 
of  that  half-reaction  to  release  electrons  in  com- 
parison to  the  H2-2H+  half-reaction.  If  the  sign 
is  positive,  the  half-reaction  has  greater  tendency 
to  release  electrons  than  does  the  H2-2H+  half- 


\ 


METALS 


\      NON  METALS 


\ 


\ 


\ 


\ 


* 


i 


-* 


• 


* 


f 


»« 


J^ 


R 


, 


* 


O 


I 


\ 


--  ■ 

1 

- 

- 

« 

- 

• 

• 

- 

— * 

/ 

% 

• 

# 

• 

• 

• 

& 

% 

* 

* 

°" 

■^ 

• 

• 

# 

» 

f 

* 

-* 

' ■  .J 

# 

t 

% 

» 

• 

• 

% 

• 

• 

• 

^J 

. 

• 

# 

• 

» 

• 

S7 

• 

• 

* 

• 

• 

# 

* 

TO 

x 

71 

• 

• 

* 

• 

OXIDES 

/Vate  /.  (Above)  The  elements  in  the  periodic  table.  (Below)  Compounds  in  the  periodic  table.     CHLORIDES 

MISCELLANEOUS 


g 


SEC.    12-2    I    ELECTRON    TRANSFER    AND    PREDICTING    REACTIONS 


211 


reaction.  If  the  sign  is  negative,  the  half-reaction     Hi(g)  — *-  2H+  +  2e 
has  less  tendency. 

Cursj  — *-  Cu+2  +  2e- 

AgfsJ  — >-  Ag+  +  e~ 


E°  =  0.000  volt       (21),  (43) 


£?  =  -0.34  volt      (/),  (44) 


E°  =  -0.80  volt    (16),  (45) 


TABLE   OF   HALF-CELL   POTENTIALS 

Table  12-11  gives  the  values  of  the  standard  oxi- 
dation potentials  for  a  number  of  half-reactions. 

A  more  complete  table  is  given  in  Appendix  3.  The  half-reactions,  listed  in  order  of  decreasing 

We  have  not  added  the  information  "1  M"  for  half-cell  potentials,  are  in  the  same  order  as  in 

each  ion  since  this  is  implied  by  the  symbol  E°.  Table  12-1,  which  was  dictated  by  laboratory  ex- 

For  the  same  reason,  25°C  and  1  atmosphere  penence. 

pressure  of  gases  are  understood.  


EXERCISE  12-3 


Table  12-11 

SELECTED    STANDARD    OXIDATION 
POTENTIALS    FOR    HALF-REACTIONS* 

HALF-REACTION  E° 

Reduced  state  Oxidized  state  (volts) 


Zn(a) 
Co(a) 

Nir«; 

H*(g) 
Cu(s) 
21- 

Agf«; 

2Br~ 

Mn+l  +  2H,0 

2C1- 

Mn+l  +  4H,0 


2e~  +  Zn+' 
2e~  +  Co+* 
2e~  +  Ni4* 
2e~  +  2H+ 
2e~  +  CU+* 
2e~  +  h(s) 
e~  +  Ag+ 
2e~  +  Br2(l) 
2e~  +  MnO,  +  4H+ 
2e~  +  Ch(g) 
5e~  +  MnO*-  +  8H+ 


+0.76 
+0.28 
+0.25 

0.000 
-0.34 
-0.53 
-0.80 
-1.06 
-1.28 
-1.36 
-1.52 


*  A  more  complete  table  is  given  in  Appendix  3. 

Let's  examine  this  table  to  see  if  it  agrees  with 
our  laboratory  experience.  Table  12-1  summa- 
rized some  of  these  results  in  a  qualitative  way. 
Extracting  these  four  half-reactions  from  Table 
12-11,  we  find 


Zn(s) 


Zn«  +  2e~ 


E°  =  +0.76  volt      (8),  (42) 


In  Exercise  12-2  you  placed  the  Ni-Ni+2  half- 
reaction  into  Table  12-1.  Check  your  placement 
by  examining  the  half-cell  potential  of  this  half- 
reaction  in  Table  12-11. 


Another  way  to  verify  the  usefulness  of  Table 
12-11  is  to  compare  its  voltage  predictions  to  one 
we  have  measured.  For  example,  we  found  a 
value  of  approximately  0.5  volt  for  a  cell  based 
on  reaction  (46): 

Zn(s)  +  Ni+2  — >-  Zn+2  +  N\(s)    (27),  (46) 

This    cell    involves    the    following    two    half- 
reactions: 


Zn(s) 


Ni(s) 


Zn+2  +  2e- 


Ni+2  +  2e~ 


E°  =  +0.76  volt      (8),  (47) 


E°  =  +0.25  volt    (32),  (48) 


By  the  values  of  E°,  we  conclude  that  the 
Zn-Zn+2  half-reaction  has  the  greater  tendency 
to  release  electrons.  Hence  it  will  tend  to  transfer 
electrons  to  nickel,  forcing  half-reaction  (48)  in 
the  reverse  direction.  Our  net  reaction  will  be 
obtained  by  subtracting  half-reaction  (48)  from 
(47): 


minus 


ZnfsJ 
Kx(s) 


Zn+2  +  2e" 
Ni«  +  2e~ 


Net  reaction    Zn(s)  +  Ni"«  — >-  Zn^  +  Nifsj 


£?  =  +0.76  volt  (47),  (49) 

E\  =  +0.25  volt  (48),  (50) 

E°  =  E\  -  El 

E°  =  (+0.76)  -  (+0.25) 

£°  =  +0.51  volt  (51) 


212 


OXIDATION-REDUCTION    REACTIONS    j    CHAP.     12 


Table  12-11  predicts  the  cell  will  operate  so  as  to 
dissolve  metallic  zinc  and  deposit  metallic  nickel, 
and  its  voltage  will  be  +0.51  volt  This  is  exactly 
what  occurs  in  such  a  cell.  Predicting  is  fun- 
let's  try  it  again!  Another  cell  we  studied  is  based 
on  reaction  (52): 

Zn(s)  +  2Ag+  — >-  Zn+2  +  2AgfsJ    (28),  (52) 

The  two  half-reactions  involved  are 


does  not  depend  upon  how  many  moles  we  con 
sider.  Thus: 


A%(s)  — >-  Ag+  +  e~ 


El  =  -0.80  volt    (16),  (58) 


Zn(s) 


Ag(s) 


Zn+2  +  2e~ 


Ag+  +  e~ 


E°  =  +0.76  volt      (8),  (53) 
E°  =  -0.80  volt    (16),  (54) 


2Ag(s)  — *-  2Ag+  +  2e~ 

E°2  =  -0.80  volt    (16),  (59) 

You  might  wonder  what  we  would  have 
learned  if  we  had  assumed  that  either  of  these 
two  cells  operates  with  the  reverse  reaction.  Sup- 
pose we  had  proposed  a  cell  based  on  oxidation 
of  nickel  and  reduction  of  zinc: 


minus 


Net  reaction    Nifsj  +  Zn+2 


NifsJ  — >-  Ni+2  +  2e~  E\  =  +0.25  volt  (32),  (60) 

Zn(s)  — ►■  Zn+2  +  2e~  E°2  =  +0.76  volt  (8),  (61) 

E°  =  E°  -  E°2 

=  (+0.25)  -  (+0.76) 
Ni+2  +  Zn(s)        E°  =  -0.51  volt  (62) 


By  the  half-cell  potentials,  we  conclude  the 
Zn-Zn+2  half-reaction  has  the  greater  tendency 
to  release  electrons.  It  will  tend  to  transfer  an 
electron  to  silver  ion,  forcing  (54)  in  the  reverse 
direction.  Hence  we  obtain  the  net  reaction  by 
subtracting  (54)  from  (53).  But  remember  that 
this  subtraction  must  be  in  the  proportion  that 
causes  no  net  gain  or  loss  of  electrons.  If  two 
electrons  are  lost  per  atom  of  zinc  oxidized  in 
(53),  then  we  must  double  half-reaction  (54)  so 
that  two  electrons  will  be  consumed. 


Our  assumption  concerning  the  chemistry  leads 
us  to  the  reverse  reaction  of  that  we  obtained 
earlier — reaction  (57) — and  to  an  equal  voltage 
but  with  opposite  sign.  The  significance  of  the 
negative  voltage  (  —  0.51  volt)  is  that  equilibrium 
in  the  reaction  favors  reactants,  not  products.  We 
obtain  the  same  prediction  we  did  before — since 
the  voltage  is  negative,  the  reaction  will  tend  to 
operate  a  cell  in  the  reverse  direction — to  dissolve 
zinc  metal  and  to  precipitate  nickel  metal.  The 
reaction  will  occur  in  the  direction  written  (con- 


minus 


Zn(s) 
2Agfs; 


Zn+2  +  2e~ 
2Ag+  +  2e- 


El  =  +0.76  volt 
El  =  -0.80  volt 


Net  reaction    Zn(s)  +  2AgH 


E°  =  El-  El 

=  (+0.76)  -  (-0.80) 
Zn+2  +  2Ag(sJ        £°  =  +1.56  volts 


(53),  (55) 
(54),  (56) 


(28),  (57) 


Our  conclusions  are  again  in  agreement  with 
experiment.  The  cell  operates  so  as  to  dissolve 
zinc  metal  and  precipitate  silver  metal.  The  volt- 
age is  indeed  about  1.5  volts.  Finally,  experiment 
shows  that  one  mole  of  zinc  does  react  with  two 
moles  of  silver  ion,  as  required  by  the  balance  of 
electrons. 

Notice  that  we  did  not  double  El  for  (54)  in 
obtaining  (57).  The  voltage  of  a  half-reaction 


suming  nickel  and  precipitating  zinc)  only  if  the 
cell  is  "overpowered"  by  an  opposing  cell  of 
higher  voltage  than  0.51  volt  (as  was  done  in  the 
experiment  pictured  in  Figure  12-6). 

12-2.2    Predicting  Reactions  from  Table  12-11 

The  ideas  we  have  developed  for  reactions  occur- 
ring in  electrochemical  cells  are  also  applicable 


SEC.     12-2    I    ELECTRON    TRANSFER    AND    PREDICTING    REACTIONS 


213 


to  reactions  that  occur  in  a  beaker.  Therefore, 
chemists  use  half-cell  potentials  to  predict  what 
chemical  reactions  can  occur  spontaneously. 

If  a  chemist  wishes  to  know  whether  zinc  can 
be  oxidized  if  it  is  placed  in  contact  with  a  solu- 
tion of  nickel  sulfate,  the  values  of  E°  help  him 
decide.  The  half-cell  potential  for  Zn-Zn+2  is 
+0.76  volt,  which  is  greater  than  that  for 
Ni-Ni+2  (which  is  +0.25  volt).  The  difference, 
+0.51  volt,  is  positive,  indicating  that  zinc  has 
a  greater  tendency  to  lose  electrons  than  does 
nickel.  Therefore,  zinc  can  transfer  electrons  to 
Ni+2.  The  chemist  predicts:  zinc  will  react  with 
Ni+2,  zinc  being  oxidized  and  nickel  being  re- 
duced. 


12-11  (something  which  can  be  oxidized),  and 
it  must  involve  a  substance  from  the  right 
column  (something  which  can  be  reduced). 
2.  A  substance  in  the  left  column  of  Table  12-11 
tends  to  react  spontaneously  with  any  sub- 
stance in  the  right  column  which  is  lower  in 
the  Table. 

Applying  these  rules,  we  would  predict:  cop- 
per metal  could  be  oxidized  to  Cu+2  by  Br2(l)  or 
Mn02(s)  but  not  by  Ni+2  or  Zn+2.  Of  course, 
copper  metal  cannot  be  oxidized  by  either  zinc 
metal  or  nickel  metal  because  neither  zinc 
metal  nor  nickel  metal  can  accept  electrons  to 
be  reduced  (as  far  as  we  know  from  Table  12-11). 


an  example:  copper  and  silver 

Suppose  the  question  is  whether  silver  will  be 
oxidized  if  it  is  immersed  in  copper  sulfate.  The 
half-cell  potential  for  Ag-Ag+  is  —0.80  volt  and 
that  for  Cu-Cu+2  is  -0.34  volt.  The  first  value, 
—  0.80  volt,  is  more  negative  than  the  second, 
—0.34  volt.  The  difference,  then,  is  still  negative: 
-0.80  -  (-0.34)  =  -0.46  volt.  The  negative 
answer  shows  that  Ag-Ag+  has  less  tendency  to 
lose  electrons  than  does  Cu-Cu+2.  The  reaction 
will  not  tend  to  proceed  spontaneously.  Silver 
will  not  be  oxidized  to  an  appreciable  extent  in 
copper  sulfate. 


EXERCISE  12-4 

Use  the  values  of  E°  to  predict  whether  cobalt 
metal  will  tend  to  dissolve  in  a  1  M  solution  of 
acid,  H+.  Now  predict  whether  cobalt  metal  will 
tend  to  dissolve  in  a  1  M  solution  of  zinc  sulfate 
(reacting  with  Zn+2). 


GENERALIZING    ON    PREDICTIONS    WITH    E° 

We  can  generalize  now  on  the  use  of  Table  12-11. 
A  substance  on  the  left  in  Table  12-11  reacts  by 
losing  electrons.  A  substance  on  the  right  reacts 
by  gaining  electrons.  We  may  draw  the  following 
conclusions: 

1.  An  oxidation-reduction  reaction  must  involve 
a  substance  from  the  left  column  of  Table 


EXERCISE  12-5 


Use  Table  12-11  to  decide  which  substances  in 
the  following  list  tend  to  oxidize  bromide  ion, 
Br~:  Ch(g),  H+,  Ni+2,  Mn(V. 


EXERCISE  12-6 


Use  Table  12-11  to  decide  which  substances  in 
the  following  list  tend  to  reduce  Br2(l):  C\~, 
H2(g),  Ni(s),  Mn+2. 


PREDICTIONS    AND    THE    EFFECT 
OF    CONCENTRATIONS 

All  of  our  predictions  have  been  based  upon  the 
values  of  E°  that  apply  to  standard  conditions. 
Yet  we  often  wish  to  carry  out  a  reaction  at 
other  than  standard  conditions.  Our  prediction 
then  must  be  adjusted  in  accord  with  Le  Cha- 
telier's  Principle  as  we  change  conditions  from 
standard  conditions  to  others  of  interest  to  us. 
For  example,  by  comparing  £°'s,  we  predicted 
that  zinc  would  dissolve  in  nickel  sulfate.  These 
£°'s  show  that  zinc  metal  will  dissolve  if  zinc  ion 
and  nickel  ion  are  both  present  at  1  M  concen- 
tration: 


Zn(s)  +  NiH 


Zn+2  +  Ni(s)    (27),  (63) 


In  our  case,  however,  no  zinc  ion  is  present  at 
all.  How  does  this  affect  our  prediction?  By 


214 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.     12 


Le  Chatelier's  Principle,  the  removal  of  Zn+2 
tends  to  shift  equilibrium  toward  the  products. 
Therefore,  removing  Zn+2  increases  the  tendency 
for  reaction  (63)  to  occur.  Our  prediction  of  re- 
action is  still  valid. 

This  will  not  always  be  the  case,  however. 
Consider  the  question:  "Will  silver  metal  dis- 
solve in  1  M  H+?"  According  to  Table  12-11, 


12-2.3    Reliability  of  Predictions 

There  is  one  more  limitation  on  the  reliability  of 
predictions  based  upon  £°'s.  To  see  it,  we  shall 
consider  the  three  reactions 
Cu(s)  +  2H+ 


Fe(s)  +  2H+ 


-*-  Cu+2  +  H,(g) 
E°  =  -0.34  volt 
■*-  Fe+2  +  U2(g) 
E°  =  +0.44  volt 


(65) 


(66) 


2Ag(s) 
2H+(1  M)  +  2e~ 


2Ag+(l  M)  +  2e~ 


E°  =  -0.80  volt 
E°  =      0.00  volt 


Net  reaction    2Ag(s)  +  2H+(1  M) 


2Ag+(l  M)  +  H,(g)        E°  =  -0.80  volt 


(64) 


The  negative  voltage  shows  that  the  state  of 
equilibrium  favors  the  reactants  more  than  the 
products  for  the  reaction  as  written.  For  stand- 
ard conditions,  the  reaction  will  not  tend  to 
occur  spontaneously.  However,  if  we  place  Ag(s) 
in  1  M  H+,  the  Ag+  concentration  is  not  1  M — 
it  is  zero.  By  Le  Chatelier's  Principle,  this  in- 
creases the  tendency  to  form  products,  in  opposi- 
tion to  our  prediction  of  no  reaction.  Some  silver 
will  dissolve,  though  only  a  minute  amount  be- 
cause silver  metal  releases  electrons  so  reluctantly 
compared  with  H2.  It  is  such  a  small  amount, 
in  fact,  that  no  silver  chloride  precipitate  forms, 
even  though  silver  chloride  has  a  very  low  solu- 
bility. 

That  some  silver  does  dissolve  to  form  Ag+ 
can  be  verified  experimentally  by  adding  a  little 
KI  to  the  solution.  Silver  iodide  has  an  even 
lower  solubility  than  does  silver  chloride.  The 
experiment  shows  that  the  amount  of  silver  that 
dissolves  is  sufficient  to  cause  a  visible  precipi- 
tate of  Agl  but  not  of  AgCl.  This  places  the  Ag+ 
ion  concentration  below  10-10  M  but  above 
10~17  M.  Either  of  these  concentrations  is  so 
small  that  we  can  consider  our  prediction  for  the 
standard  state  to  be  applicable  here  too — silver 
metal  does  not  dissolve  appreciably  in  1  M  HC1. 
In  general,  the  question  of  whether  a  prediction 
based  upon  the  standard  state  will  apply  to  other 
conditions  depends  upon  how  large  is  the  mag- 
nitude of  E°.  If  E°  for  the  overall  reaction  is  only 
one-  or  two-tenths  volt  (either  positive  or  nega- 
tive), then  deviations  from  standard  conditions 
may  invalidate  predictions  that  do  not  take  into 
account  these  deviations. 


3Fe(s)  +  2N03-  +  8H+ 
+  2NO(g)  +  4H..O 


-+-  3Fe+2 
E°  =  +1.40  volts 


(67) 


The  three  values  of  E°  are  easily  calculated  from 
half-cell  potentials.  Then,  we  can  predict  with 
confidence  that  reaction  (65)  will  not  occur  to  an 
appreciable  extent  if  solid  copper  is  immersed  in 
dilute  acid.  The  negative  value  of  E°  (—0.34  volt) 
indicates  that  equilibrium  in  (65)  strongly  favors 
the  reactants,  not  the  products. 

Furthermore,  we  can  predict  that  reactions 

(66)  and  (67)  might  occur.  The  positive  values  of 
E°  (+0.44  and  +1.40  volts)  show  that  equilib- 
rium strongly  favors  products  in  these  reactions. 
Again,  experiments  are  warranted.  A  piece  of 
iron  is  immersed  in  dilute  acid — bubbles  of  hy- 
drogen appear.  Reaction  (66)  does  occur.  A  piece 
of  iron  is  immersed  in  a  one  molar  nitric  acid 
solution — though  bubbles  of  hydrogen  may  ap- 
pear, no  nitric  oxide,  NO,  gas  appears.  Reaction 

(67)  between  iron  and  nitrate  ion  does  not  im- 
mediately occur:  the  reaction  rate  is  extremely 
slow.  This  slow  rate  could  not  be  predicted  from 
the  £°'s. 

So  the  equilibrium  predictions  based  on  E0,s 
do  not  make  all  experiments  unnecessary.  They 
provide  no  basis  whatsoever  for  anticipating 
whether  a  reaction  will  be  very  slow  or  very  fast. 
Experiments  must  be  performed  to  learn  the 
reaction  rate.  The  £°'s  do,  however,  provide 
definite  and  reliable  guidance  concerning  the 
equilibrium  state,  thus  making  many  experi- 
ments unnecessary;  the  multitude  of  reactions 
that  are  foredoomed  to  failure  by  equilibrium 
considerations  need  not  be  performed. 


SEC.     12-2    I    ELECTRON    TRANSFER    AND    PREDICTING    REACTIONS 


215 


12-2.4    E   and  the  Factors  That 
Determine  Equilibrium 

We  see  that  E°  furnishes  a  basis  for  predicting  the  equi- 
librium state.  In  Chapter  9  and  in  subsequent  chapters, 
equilibrium  was  treated  in  terms  of  two  opposing  tenden- 
cies— toward  minimum  energy  and  toward  maximum 
randomness.  What  is  the  connection  between  E°  and 
these  two  tendencies? 

Consider  two  reactions  for  which  E°  shows  that  prod- 
ucts are  favored,  one  an  exothermic  reaction,  and  the 
other  an  endothermic  reaction.  For  the  exothermic  reac- 
tion, when  the  reactants  are  mixed  they  are  driven  toward 
equilibrium  in  accord  with  the  tendency  toward  minimum 
energy.  Now  contrast  the  endothermic  reaction  for  which 
E°  shows  that  equilibrium  favors  products.  When  these 
reactants  are  mixed,  they  approach  equilibrium  against 
the  tendency  toward  minimum  energy  (since  heat  is  ab- 
sorbed). This  reaction  is  driven  by  the  tendency  toward 
maximum  randomness. 

Summarizing,  E°  measures  quantitatively  the  difference 
between  the  tendency  to  minimum  energy  and  tendency  to 
maximum  randomness  under  the  standard  state  conditions. 


12-2.5    Oxidation  Numbers — An  Electron 
Bookkeeping  Device 

The  reaction  between  ferric  ion,  Fe+3,  and  cu- 
prous ion,  Cu+,  to  produce  ferrous  ion,  Fe+2, 
and  cupric  ion,  Cu+2,  is  plainly  an  oxidation- 
reduction  reaction: 


Fe+3  +  Cu+ 


Fe+2  +  CuH 


(68) 


It  is  readily  separated  into  two  half-reactions 
showing  electron  transfer: 


Oxidation  (loss  of  electrons) 
Cu+ 

Reduction  (gain  of  electrons) 

Fe+3  +  e- 


Cu+2  +  e-    (69) 


Fe+2    (70) 


Because  of  the  presence  of  Cu+  ion,  ferric  ion  is 
reduced.  Chemists  say  that  Cu+  ion  acts  as  a 
reducing  agent  in  this  reaction — Cu+  ion  is  the 
"agent"  that  caused  the  reduction  of  ferric  ion. 
At  the  same  time,  Cu+  is  oxidized  because  of  the 
presence  of  ferric  ion.  Hence,  Fe+3  is  called  an 
oxidizing  agent  in  this  reaction. 
Another  reaction  by  which  ferric  ion  can  be 


reduced  involves  bisulfite*  ion,  HS03  .  The  bal- 
anced equation  is 

H20  +  HS03-  +  2Fe+3  — ►■ 

2Fe+2  +  HS04"  +  2H+    (71) 

Again  half-reaction  (70)  describes  what  happens 
to  ferric  ion: 


2Fe+3  +  2e- 


2FeH 


(72) 


Since  two  electrons  are  gained  by  the  two  ferric 
ions  in  half-reaction  (72),  two  electrons  must  be 
released  by  the  remaining  constituents  in  (71). 
The  other  half-reaction  can  be  found  by  sub- 
tracting (72)  from  (71)  to  give, 


H20  +  HSO3-  -  2e~  — ►-  HSOf  +  2H+ 


or 


H20  +  HSOi 


HS04-  +  2H+  +  2e~    (73) 


The  combination  of  H20  and  HS03~  acts  as  a 
reducing  agent  toward  Fe+3.  Since  water  solu- 
tions of  Fe+3  are  quite  stable,  HS03~  is  con- 
sidered to  be  the  actual  reducing  agent. 

Half-reaction  (73)  differs  from  the  others  we 
have  looked  at.  This  is  not  the  same  simple  situa- 
tion in  which  a  single  atom  is  oxidized  by  re- 
leasing electrons  or  is  reduced  by  accepting  them. 
Here  a  complicated  ion,  HS03~,  furnishes  elec- 
trons to  Fe+3  in  an  intricate  process  in  which  an 
oxygen  atom  is  transferred  from  H20  to  HS03" , 
forming  HSCXf  and  two  hydrogen  ions.  We  can- 
not assign  the  electron  loss  to  a  particular  atom. 
In  this  situation,  it  is  convenient  to  have  a  book- 
keeping device  that  at  least  keeps  track  of  the 
number  of  electrons  so  that  none  is  forgotten. 
The  bookkeeping  device  is  called  assigning  oxi- 
dation numbers. 

Since  we  don't  know  the  locations  of  the  elec- 
trons held  by  a  molecule  such  as  HS03~,  we 
assume  that  the  hydrogen  atom  has  a  -f  1  charge, 
that  each  oxygen  atom  has  a  —2  charge,  and 
that  the  sulfur  atom  has  all  the  rest  of  the  elec- 
trons in  the  molecule.  Of  course,  if  the  charges 
on  all  the  atoms  in  HS03~  are  added  together, 
they  must  sum  to  —1,  the  molecular  charge. 
Since  there  are  three  oxygen  atoms  in  HS03~ ,  the 
algebra  looks  like  this: 

*  This  ion,  HSOf ,  is  also  called  hydrogen  sulfite. 


216 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.     12 


f  charge  on  \  /charge  on\       /charge  on\       ,^.      ,    x 


\     atom     / 


atom 


atom 


(charge  on 
sulfur 
atom 


charge  on 
sulfur 
atom 


(-D 
(+4) 


This  fictitious  charge  is  called  the  oxidation  num- 
ber of  sulfur: 

Oxidation  number  sulfur  =  +4  in  HS03~     (74) 

The  same  process  can  be  applied  to  HS04~. 
Again  assuming  hydrogen  has  a  charge  of  +1 
and  each  of  the  four  oxygen  atoms  has  a  charge 
of  —2,  we  calculate  a  fictitious  charge  on  the 
sulfur  atom  of  +6: 

Oxidation  number  sulfur  =  +6  in  HS04~     (75) 

According  to  the  oxidation  number  bookkeep- 
ing, the  two  electrons  released  in  the  HS03_- 
HS04~  half-reaction  (73)  are  associated  with  the 
change  in  oxidation  number  of  sulfur  from  +4 
to  +6. 

This  arbitrary  scheme  of  assigning  oxidation 
numbers  turns  out  to  be  quite  useful,  provided 
we  don't  forget  that  the  oxidation  number  is  a 
fictitious  charge.  We  can  see  the  usefulness  by 
considering  a  reaction  related  to  the  oxidation 
of  HSO3-. 

Sulfur  forms  two  oxides,  S02  (a  gas  at  normal 
conditions)  and  S03  (a  liquid  that  boils  at 
44.8°C).  Under  suitable  conditions,  S02  reacts 
with  oxygen  to  form  S03: 


2SO,fgj  +  Q2(g) 


2SOi(g) 


(76) 


Is  this  an  oxidation-reduction  reaction?  His- 
torically, it  surely  is,  for  the  term  "oxida- 
tion" originally  referred  specifically  to  reactions 
with  oxygen.  Yet  our  electron-transfer  view  of 
oxidation-reduction  reactions  provides  no  help 
in  deciding  so.  Where  in  reaction  (76)  is  there 
any  evidence  of  electrons  being  gained  or  lost? 
In  such  a  doubtful  case,  our  oxidation  number 
scheme  provides  an  answer.  Applying  the  same 
assumptions  used  in  treating  the  HS03~~-HS04~ 


half-reaction,  we  can  calculate  the  oxidation 
numbers  of  sulfur  in  S02  and  S03. 


In  S02: 

(charge  on  \       /charge  on 
oxygen    J  +  I      sulfur 
atom     /      \     atom 


/  molecular  \ 
\    charge    / 


(-2) 
assumed 


(oxidation  \ 
number    1  = 
sulfur    / 


(0) 


Oxidation  number  sulfur  =  +4  in  S02       (77) 


In  S03: 

charge  on 

oxygen 

atom 


(charge  on 
sulfur 
atom 

(oxidation\ 
number    1 
sulfur    / 

Oxidation  number  sulfur 


/  molecular  \ 
\    charge    / 


(0) 
+6  in  S03       (75) 


Thus,  in  reaction  (76)  the  sulfur  atom  changes 
oxidation  number  from  +4  to  +6,  just  as  it  did 
in  the  HS03~-HS04-  half-reaction.  The  oxida- 
tion number  gives  a  basis  for  connecting  the 
S02-S03  change  to  the  oxidation  of  HS03"  to 
HS04~.  Both  changes  are  considered  to  be  ex- 
amples of  oxidation. 

Oxidation-reduction  reactions  occurring  in 
aqueous  solutions  are  conveniently  treated  in 
terms  of  half-reactions  showing  transfer  of  elec- 
trons. Under  more  general  conditions  (gaseous 
state,  other  solvents,  etc.),  it  is  more  convenient 
to  treat  oxidation-reduction  reactions  in  terms 
of  oxidation  numbers,  based  upon  the  arbitrary 
scheme  of  assigning  charge  +1  to  a  hydrogen 
atom  bound  to  an  unlike  atom  and  charge  —2 
to  an  oxygen  atom  when  it  is  bound  to  un- 
like atoms.  Generally,  then,  an  oxidation' 
reduction  reaction  is  one  in  which  oxidation 
numbers  change. 


SEC.     12-3        BALANCING    OXIDATION-REDUCTION    REACTIONS 


217 


EXERCISE  12-7 

The  reactions  by  which  S02  and  S03  dissolve  in 
water  are  not  considered  to  be  oxidation- 
reduction  reactions: 

SO/*  J  +  H,0  — +■  HSCV  (aq)  +  H+(aq)     (79) 
SOJg)  +  H,0  — >-  HS04" (aq)  +  U+(aq)     (80) 

Convince  yourself  that  none  of  the  atoms  in 
either  (79)  or  (80)  changes  oxidation  number. 


EXERCISE  12-8 

In  reaction  (76)  the  oxidation  number  of  sulfur 
changes  from  +4  to  '+6.  According  to  this,  two 
electrons  are  released  by  each  sulfur  atom  oxi- 
dized. Show  that  these  electrons  are  gained  by 
oxygen  if  we  assume  oxygen  has  oxidation  num- 
ber equal  to  zero  in  02. 


12-3     BALANCING  OXIDATION-REDUCTION  REACTIONS 


When  a  reaction  is  properly  written,  it  expresses 
certain  conservation  laws.  A  chemical  reaction 
does  not  destroy  or  produce  atoms.  Therefore 
there  must  be  the  same  number  and  types  of 
atoms  among  the  reactants  as  among  the  prod- 
ucts. A  chemical  reaction  also  does  not  destroy 
or  produce  electric  charge.  Therefore  the  sum  of 
the  charges  appearing  on  the  reactant  species 
must  be  the  same  as  the  sum  of  the  charges 
appearing  on  the  products.  The  process  by  which 
we  write  a  correct  chemical  reaction,  making 
sure  these  two  conservation  laws  are  followed, 
is  called  balancing  the  chemical  reaction.* 

Oxidation-reduction  reactions  must  be  bal- 
anced if  correct  predictions  are  to  be  made.  Just 
as  in  selecting  a  route  for  a  trip  from  San  Fran- 
cisco to  New  York,  there  are  several  ways  to 
reach  the  desired  goal.  Which  route  is  best  de- 
pends to  some  extent  upon  the  likes  and  dislikes 
of  the  traveler.  We  will  discuss  two  ways  to 
balance  oxidation-reduction  reactions — first,  us- 
ing half-reactions  and,  next,  using  the  oxidation 
numbers  we  have  just  introduced. 


12-3.1    Use  of  Half-Reactions  for  Balancing 
Oxidation-Reduction  Reactions 

Suppose  we  want  to  describe  by  an  equation 
what  happens  when  pure  lithium  metal  is  added 
to  a  1  M  HC1  solution.  Our  first  step  must  be  to 
decide  what  products  will  be  obtained.  This  can 
be  determined  only  by  experiment.  Often  you 

*  Some  chemists  prefer  to  call  this  process  balancing 
the  equation  for  the  reaction. 


will  know  already  what  the  experiment  would 
give.  Sometimes  what  the  products  will  be  is  not 
obvious  but  it  is  easily  learned  from  a  reference 
book.  Sometimes  you  must  do  the  experiment 
yourself  or  proceed  on  the  basis  of  an  assump- 
tion. 

For  lithium  metal  in  1  M  HC1,  the  observed 
facts  are  that  the  metal  dissolves  spontaneously 
and  a  gas  bubbles  out  of  the  solution.  From 
Appendix  3  we  select  the  two  half-reactions 
(notice  that  the  half-reactions  are  already  "bal- 
anced" in  both  charge  and  number  of  atoms): 

U(s)  — ►-  Li+  +  e~  (81) 

H2(g)  -+  2H+  +  2e~         (21),  (82) 

Both  of  these  half-reactions  show  production  of 
electrons.  But  we  know  there  must  be  an  electron 
used  for  each  produced,  so  one  of  the  equations 
must  be  reversed.  Experiment  shows  us  it  is  the 
second  because  hydrogen  gas  is  evolved  from  the 
solution.  The  first  equation  is  correct  as  written. 
Lithium  metal  dissolves  and  is  converted  to  ions. 
Thus, 

Li(s)  — ►-  Li+  +  e~       (81),  (83) 

2H+  +  2e-  — >-  H2(g)  (9),  (84) 

These  equations  now  correctly  express  the  ob- 
served facts  and  they  must  be  properly  com- 
bined. The  first  step  is  to  make  them  indicate  the 
same  number  of  electrons  produced  and  used. 
We  "balance  the  electrons."  By  inspection  we 
find  that  we  achieve  balance  by  doubling  the 
first  equation. 

2U(s)  — >-  2Li+  +  2e~     (81),  (85) 


2H+  +  2e 


2LiH 
H*f*J 


(9),  (86) 


218 


OXIDATION-REDUCTION   REACTIONS   I   CHAP.    12 


Now  we  add  the  two  equations  to  get  the  net 
reaction, 

2H+  +  2Li(s)  — *-  2Li+  +  U,(g)         (87) 

Note  that  the  electrons  cancel — we  took  care 
that  this  would  happen  because  we  know  elec- 
trons are  neither  consumed  nor  produced  in  the 
net  reaction. 

As  a  final  check,  let  us  verify  the  conservation 
of  charge: 


The  electrons  cancel  out.  The  atoms  and  charges 
are  properly  balanced,  but  there  is  one  remaining 
disturbance.  There  are  H+  ions  included  among 
the  reactants  as  well  as  the  products.  Cancelling 
the  excess,  we  obtain  the  final,  balanced  reaction: 

5H£(g)  +  2Mn04"  +  6H+  — ►- 

5S(s)  +  2MH+2  +  8H20    (95) 

Before  leaving  the  equation,  let  us  check  the 
electric  charge  balance: 


5HS(g)  +  2Mn04-  +    6H+ 
5(0)      +  2(-l)   +6(+l) 
-2      +    +6 
+4 


2H+  +  2Li(s)  — ►-  2Li+  +  H2(g) 
+2+0  +2+0 

+2     =      +2 

As  a  more  complex  case,  suppose  we  want  to 
write  the  equation  for  the  reaction  that  occurs 
when  hydrogen  sulfide  gas,  H2S,  is  bubbled  into 
an  acidified  potassium  permanganate  solution, 
KMn04.  When  we  do  this,  we  observe  that  the 
purple  color  of  the  MnCV  ion  disappears  and 
that  the  resulting  mixture  is  cloudy  (sulfur  par- 
ticles). From  Appendix  3  we  find  the  two  half- 
reactions 

H£(g)  — ►-  S(s)  +  2H+  +  2e~         (88) 
Mn+2  +  4H20  — ■*-  Mn04"  +  8H+  +  5e~    (89) 

Since  we  know  that  sulfur  is  formed,  we  will  use 
equation  (88)  as  it  is  written.  However,  the  purple 
Mn04~  is  being  changed  to  almost  colorless  Mn+2 
so  we  will  rewrite  equation  (89)  as  the  reverse  of 
what  we  obtained  from  Appendix  3: 

H2S(g)  — ►-  S(s)  +  2H+  +  2e~    (88),  (90) 
Mn04-  +  8H+  +  5e~  — >-  Mn+2  +  4H20    (91) 

Balancing  the  electrons  is  the  next  step,  but  this 
time  it  is  a  little  more  difficult.  We  see  that  if  we 
multiply  equation  (90)  by  5  and  equation  (97) 
by  2,  there  will  be  10  electrons  in  each  case. 
Arithmetic  teachers  call  this  finding  the  least 
common  multiple: 

5H2Sfs;  — >- 
2Mn04"  +  16H+  +  10e"  — ►- 


->-   5S(s)  +  2Mn+2  +  8H20  (95) 

5(0)   +  2(+2)  +   8(0) 
+4 
=    +4 

12-3.2    Balancing  Half-Reactions 

When  potassium  chlorate  solution,  KC103,  is  added  to 
hydrochloric  acid,  chlorine  gas  is  evolved.  Although  we 
can  find  the  half-reaction,  2C1~  =  C\2(g)  +  2e~,  in  Ap- 
pendix 3,  we  find  no  equation  with  CIO3"  ion  involved. 
We  can  surmise  that  CIO3"  is  accepting  electrons  and 
changing  into  chlorine.  Let  us  write  a  partial  half-reaction 
in  which  we  indicate  an  unknown  number  of  electrons 
and  in  which  we  have  conserved  only  chlorine  atoms: 


CIO3    +xe- 


hCUg) 


(96) 


From  experience  we  note  that  in  acid  solution  the  oxygen 
in  such  oxidizing  agents  as  MnOi"  and  Cr20f2  ends  up 
as  water  and  that  H+  is  consumed.  Let  us  include  this 
notion  by  showing  6H+  among  the  reactants  and  3H20 
among  the  products: 


CIO3"  +  6H+  +  xe~ 


\C\2(g)  +  3H20     (97) 


Finally,  we  have  to  remember  that  charge  is  conserved. 
Since  the  products  are  neutral  molecules,  x  will  have  to 
be  5  in  order  that  the  total  charge  represented  among  the 
reactants  is  zero.  Our  desired  half-reaction  is 

CIO3"  +  6H+  +  5e~  — ►-  $CU(g)  +  3H20      (98) 

Now  we  can  return  to  working  out  the  equation  for  the 
reaction  we  observed.  The  least  common  multiple  be- 
tween 1  and  5  is  5.  Writing  the  half-reactions  to  involve  5 
electrons  and  adding  them,  we  obtain 

5CI-  — >-  iCh(g)  +  Se~         (99) 
CIO3"  +  6H+  +  5e~  — >■  hCh(g)  +  3H20    (700) 


5C1-  +  CIO3"  +  6H+ 


3Cl2fgJ  +  3H20    (101) 


Again  demonstrate  to  yourself  that  atoms  and  charge  are 
conserved. 


5S(s)  +  10H+  +  \0e~ 
2Mn+2  +  8H20 


5HS(g)  +  2Mn04"  +  16H+  — >-  5S(s)  +  10H+  +  2Mn+2  +  8H20 


(92) 
(93) 

(94) 


SEC.     12-3    |    BALANCING    OXIDATION-REDUCTION    REACTIONS 


219 


12-3.3     Use  of  Oxidation  Number  in  Balancing 
Oxidation-Reduction  Reactions 

We  have  already  introduced  oxidation  numbers 
as  a  device  for  assigning  a  fictitious  charge  to  an 
atom  in  a  molecule.  According  to  this  scheme, 
oxidation-reduction  reactions  involve  changes  of 
oxidation  numbers.  Consideration  of  conserva- 
tion of  charge  reveals  that  there  must  be  a  bal- 
ance between  changes  of  oxidation  number.  Con- 
sequently, oxidation  numbers  provide  just  as 
good  a  basis  for  balancing  equations  as  do  half- 
reactions. 

ASSIGNING    OXIDATION    NUMBERS 

For  the  present,  we  will  limit  ourselves  to  mole- 
cules containing  hydrogen  and/or  oxygen  along 
with  the  element  to  which  we  wish  to  assign  an 
oxidation  number.  The  rules  we  will  utilize  are 
as  follows: 

(1)  The  oxidation  number  of  a  monatomic  ion 
is  equal  to  the  charge  on  the  ion. 

(2)  The  oxidation  number  of  any  substance  in 
the  elementary  state  is  zero. 

(3)  The  oxidation  number  of  hydrogen  is  taken 
to  be  + 1  (except  in  H2,  which  is  the  elemen- 
tary state). 

(4)  The  oxidation  number  of  oxygen  is  taken  to 
be  —2  (except  in  02;  ozone;  03;  and  perox- 
ides). 

(5)  The  other  oxidation  numbers  are  selected  to 
make  the  sum  of  the  oxidation  numbers 
equal  to  the  charge  on  the  molecule. 

(6)  Reactions  occur  such  that  the  net  change  of 
oxidation  numbers  is  zero.  (This  last  rule  is 
really  a  result  of  the  conservation  of  charge.) 

Do  not  worry  about  the  exceptions  included 
within  parentheses  in  rules  3  and  4.  Your  atten- 
tion will  be  called  to  them  later  when  substances 
involving  them  are  considered. 


EXERCISE  12-9 

Show  that  the  oxidation  number  of  nitrogen  is 
+  5  in  each  of  the  two  species  N03"  and  N206. 


BALANCING    REACTIONS 

Just  as  before,  the  first  step  in  balancing  a  reac- 
tion must  be  to  decide  the  products.  Again,  ex- 
periment provides  the  answer.  Let  us  reconsider 
one  of  the  same  examples  we  balanced  previously 
by  the  half-reaction  method.  For  these  we  al- 
ready know  the  products. 

In  the  second  example  of  Section  12-3.1,  we 
find  H2S  gas  reacts  with  Mn04~  to  give  solid 
sulfur  and  Mn+2: 

MntV  +  H£(g)  gives  S(s)  +  Mn«     (702) 

First,  we  assign  oxidation  numbers  to  each  ele- 
ment, using  rules  1-5.  We  find 

Mn04"  +  HS(g)         S(s)  +  Mn« 

+7  -2  0  +2 

with  changes, 


Oxidation 
number 


for  manganese +7 
for  sulfur 


-5 


+2 


-2  — *-  0 

+  2 


If  the  gain  in  oxidation  number  by  sulfur  is  to 
equal  the  loss  by  manganese,  then  five  atoms  of 
sulfur  must  react  with  two  atoms  of  manganese: 

2Mn04"  +  5H$(g)  gives  5S(s)  +  2Mn+2 

(not  balanced)    (103) 


2(+7) 


2(-5)=-10 


+  2(+2) 


5(-2) 


5(  +  2)  =  +10 


^5(0) 


Now  we  proceed  to  ensure  conservation  of  oxy- 
gen atoms.  There  are  eight  oxygen  atoms  on  the 
left  in  (103),  hence  we  must  add  eight  molecules 
of  H20  to  the  right.  (The  reaction  occurs  in 
aqueous  solution,  so  there  is  plenty  of  H20.) 

2Mn04"  +  5H2S(g)  gives 

5S(s)  +  2Mn+2  +  8H20  (not  balanced)    (104) 

Next  we  must  ensure  conservation  of  hydrogen 
atoms.  On  the  left,  there  are  10  hydrogen  atoms 
(in  5H2S)  and  on  the  right  16  (in  8H20).  In 
aqueous  solutions  (in  neutral  or  acidic  solution) 
we  assume  that  these  six  hydrogen  atoms  needed 
on  the  left  are  provided  by  H+: 

2Mn04~  +  5HS(g)  +  6H+  — >- 

5S(s)  +  2Mn+2  +  8H20    (95),  (105) 

The  equation  is  balanced  now  but  experience 
dictates  that  a  check  should  always  be  made  on 
the  basis  of  charge  balance: 


220 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.     12 


2(-l)  +  5(0)  +  6(+l)      5(0)  +  2(+2)  +  8(0)  Of  course,  the  oxidation  number  method  gives 

—  2  +6  +4  the  same  balanced  equation  as  the  half-reaction 

-|-4  =  -|-4  method. 


12-4     ELECTROLYSIS 


So  far  in  this  chapter  we  have  dealt  with  reac- 
tions that  proceed  spontaneously.  But  the  same 
ideas  and  names  are  applied  to  reactions  that  we 
force  to  take  place,  against  their  natural  tend- 
ency, by  supplying  energy  with  an  externally 
applied  electric  current.  Such  a  process  is  termed 
electrolysis  or  "separation  by  electricity." 

We  have  dealt  with  electrolysis  before — every 
time  we  discussed  or  measured  the  electrical  con- 
ductivity of  an  electrolyte  solution.  To  see  this, 
let's  consider  the  processes  that  occur  when  we 
cause  electric  charge  to  pass  through  an  aqueous 
solution  of  hydrogen  iodide. 

A  distinguishing  property  of  ionic  solutions  is 
electrical  conductivity,  just  as  it  is  a  distinguish- 
ing property  for  metals,  but  the  current-carry- 
ing mechanism  differs.  Electric  charge  moves 
through  a  metal  wire,  we  believe,  by  means  of 


electron  movement.  Electrons  flow  through  the 
wire  without  changing  the  metal  chemically. 
In  contrast,  the  movement  of  electric  charge 
through  an  aqueous  solution  of  an  electrolyte 
causes  significant  chemical  changes. 

Figure  12-7  shows,  on  the  right,  the  behavior 
of  an  aqueous  hydrogen  iodide  solution  during 
conduction.  The  two  carbon  rods  are  connected 
by  wires  to  the  terminals  of  a  2  volt  battery. 
Electrons  flow  from  the  battery  through  the  left 
carbon  rod,  entering  the  solution.  An  equal  num- 
ber of  electrons  leave  the  solution  through  the 
right  carbon  rod  to  return  to  the  battery.  The 
hydrogen  ion,  H+,  has  the  ability  to  accept  an 


Fig.  12-7.  A  schematic  view  of  electrolytic 
conduction. 


QUESTIONS    AND    PROBLEMS 


221 


electron  at  the  left  electrode,  where  electrons  are 
in  excess.  The  H+  ion  is  changed  chemically  to  a 
neutral  atom.  An  iodide  ion,  I-,  has  one  excess 
electron  that  can  be  released  at  the  right  elec- 
trode, where  electrons  are  in  deficiency.  The  I- 
ion  changes  chemically  to  a  neutral  atom.  The 
net  result  of  these  two  occurrences  is  the  process 
we  call  electrolysis. 

Let  us  sum  up  the  process  during  the  move- 
ment of  one  electron  through  the  entire  circuit 
shown  on  the  left  in  Figure  12-7. 

(1)  An  electron  on  the  left  carbon  rod  is  gone; 
another  electron  has  shown  up  at  the  right 
carbon  rod. 


charge.  This  drift-  of  the  ions  through  the 
solution,  positive  ions  in  one  direction  and 
negative  ions  in  the  other,  explains  the  con- 
duction in  aqueous  solutions. 
(4)  The  battery  performs  work  in  forcing  current 
to  flow  through  the  solution  and  in  causing 
chemical  changes  to  occur  that  would  not 
proceed  spontaneously. 


The  net  reaction  is 
2H+  +  21" 


H,fej  +  I2 


006) 


Reaction  (106)  is  just  an  oxidation-reduction  re- 
action and  it  is  readily  separated  into  the  two 
half-reactions 


2H+  +  2e~  =  H,(g) 

21-  =  \,(g)  +  2e~ 


E°  =      0.000  volt 
E°  =  -0.53  volt 


Overall  reaction    2H+  +  21"  =  H2(g)  +  h(g)        E°  =  -0.53  volt 


(106),  (107) 


(2)  One  W+(aq)  ion  and  one  l~(aq)  ion  are 
gone;  one  H  atom  and  one  I  atom  have  been 
formed. 

(3)  As  this  process  continues  to  take  place,  the 
H+  ions  in  the  solution  at  the  left  tend  to  be 
used  up  and  the  same  occurs  for  I"  ions  in 
the  solution  at  the  right.  Since  only  positive 
ions  are  used  up  at  the  left,  the  remaining 
negative  ions  are  electrically  repelled  from 
this  region  and  are  attracted  toward  the 
solution  at  the  right  where  positive  ions  are 
plentiful.  Here,  negative  ions  are  used  up,  so 
the  remaining  positive  ions  are  repelled  and 
they  are  attracted  toward  the  left  where  nega- 
tive ions  are  plentiful.  Iodide  ions,  I~  move 
from  left  to  right  through  the  solution, 
carrying  negative  charge.  At  the  same  time, 
hydrogen  ions,  H+,  move  from  right  to 
left  through  the  solution,  carrying  positive 


The  negative  value  of  E°  =  —0.53  volt  tells  us 
that  the  reaction  will  not  occur  spontaneously  as 
written.  This  voltage  tells  us  further  that  elec- 
trolysis will  occur  only  if  a  cell  with  a  voltage 
exceeding  —0.53  volt  is  placed  in  the  external 
circuit  so  as  to  oppose  the  voltage  generated  by 
the  cell  itself. 


EXERCISE  12-10 

From  Appendix  3,  estimate  the  minimum  voltage 
required  to  cause  electrolysis  of  1  M  HC1,  form- 
ing H-2(g)  and  02(g),  each  at  1  atmosphere  pres- 
sure. Show  that  at  this  voltage  electrolysis  to 
produce  H2(g)  and  C\2(g)  will  not  occur. 


QUESTIONS  AND  PROBLEMS 


1.  One  method  of  obtaining  copper  metal  is  to  let        2.  (a)  If  a  neutral  atom  becomes  positively  charged, 


a  solution  containing  Cu"*"2  ions  trickle  over  scrap 
iron.  Write  the  equations  for  the  two  half- 
reactions  involved.  Assume  the  iron  becomes 
Fe~2.  Indicate  in  which  half-reaction  oxidation 
is  taking  place. 


has  it  been  oxidized  or  reduced?  Write  a 
general  equation  using  M  for  the  neutral 
atom. 

(b)  If  an  ion  X'1  acquires  a  —2  charge,  has  it 


222 


OXIDATION-REDUCTION    REACTIONS    I    CHAP.     12 


been  oxidized  or  reduced?  Write  a  general 
equation. 

3.  Aluminum  metal  reacts  with  aqueous  acidic 
solutions  to  liberate  hydrogen  gas.  Write  the 
two  half-reactions  and  the  net  ionic  reaction. 

4.  When  copper  is  placed  in  concentrated  nitric 
acid,  vigorous  bubbling  takes  place  as  a  brown 
gas  is  evolved.  The  copper  disappears  and  the 
solution  changes  from  colorless  to  a  greenish- 
blue.  The  brown  gas  is  nitrogen  dioxide,  N02, 
and  the  solution's  color  is  due  to  the  formation 
of  cupric  ion,  Cu+2.  Using  half-reactions  from 
Appendix  3,  write  the  net  ionic  equation  for  this 
reaction. 

5.  Nickel  metal  reacts  with  cupric  ions,  Cu+2,  but 
not  with  zinc  ions,  Zn+2;  magnesium  metal  does 
react  with  Zn+2.  In  each  case  of  reaction,  ions 
of  +2  charge  are  formed.  Use  these  data  to 
expand  the  table  of  reactions  on  p.  206. 

6.  In  acid  solution  the  following  are  true :  H2S  will 
react  with  oxygen  to  give  H20  and  sulfur.  H2S 
will  not  react  in  the  corresponding  reaction  with 
selenium  or  tellurium.  H2Se  will  react  with  sulfur 
giving  H2S  and  selenium  but  it  will  not  react 
with  tellurium.  Arrange  the  hydrides  of  column 
VI,  H:0,  H2S,  H2Se,  and  H2Te,  in  order  of  their 
tendency  to  lose  electrons  to  form  the  elements, 
02,  S,  Se,  and  Te. 

7.  If  you  wish  to  replate  a  silver  spoon,  would  you 
make  it  the  anode  or  cathode  in  a  cell  ?  Use  half- 
reactions  in  your  explanation.  How  many  moles 
of  electrons  are  needed  to  plate  out  1.0  gram 
of  Ag? 

8.  Figure  12-5  shows  electrons  leaving  the  Cu(s) 
and  going  to  the  Ag(s).  Experimentally,  both 
half-cells  are  found  to  be  electrically  neutral 
before  current  flows  and  to  remain  so  as  the  cell 
operates.  Explain  this. 

9.  In  the  electrolysis  of  aqueous  cupric  bromide, 
CuBr2,  0.500  gram  of  copper  is  deposited  at  one 
electrode.  How  many  grams  of  bromine  are 
formed  at  the  other  electrode?  Write  the  anode 
and  cathode  half-reactions. 

Answer.  1.26  grams  of  Br2(l) 

10.  Complete  the  following  equations.  Determine 
the  net  potential  of  such  a  cell  and  decide 
whether  reaction  can  occur. 


(a)  Zn  +  Ag+  — ■*- 

(b)  Cu  +  Ag+  — >- 

(c)  Sn  +  Fe+2  — >■ 

(d)  Hg  +  H+  — ►- 

11.  For  each  of  the  following, 

(i)  write  the  half-reactions ; 
(ii)  determine  the  net  reaction ; 
(iii)  predict  whether  the  reaction  can  occur  giv- 
ing the  basis  for  your  prediction : 

(a)  Mg(s)  +  Sn+2  — >- 

(b)  Mn(s)  +  Cs+  — >- 

(c)  Cu(s)  +  C\2(g)  — ►- 

(d)  Zn(s)  +  Fe+2  — ■+- 

(e)  Fe(s)  +  Fe+3  — >- 

12.  A  half-cell  consisting  of  a  palladium  rod  dipping 
into  a  1  M  Pd(N03)2  solution  is  connected  with 
a  standard  hydrogen  half-cell.  The  cell  voltage 
is  0.99  volt  and  the  platinum  electrode  in  the 
hydrogen  half-cell  is  the  anode.  Determine  E° 
for  the  reaction 


Pd 


Pd+2  +  2e~ 


13.  Suppose  chemists  had  chosen  to  call  the  2I~  — »- 
I2  +  2e~  half-cell  potential  zero. 

(a)  What  would  be  E°  for  Na  — )-  Na+  +  e~l 

(b)  How  much  would  the  net  potential  for  the 
reaction  2Na  +  I2  — >-  2Na+  +  2I~  change? 

14.  If  a  piece  of  copper  metal  is  dipped  into  a  solu- 
tion containing  Cr+3  ions,  what  will  happen? 
Explain,  using  E°s. 

15.  What  would  happen  if  an  aluminum  spoon  is 
used  to  stir  an  Fe(N03)2  solution?  What  would 
happen  if  an  iron  spoon  is  used  to  stir  an  A1C13 
solution? 

16.  Can  1  M  Fe2(S04)3  solution  be  stored  in  a  con- 
tainer made  of  nickel  metal?  Explain  your  an- 
swer. 

17.  Suppose  water  is  added  to  each  of  the  beakers 
containing  copper  sulfate  in  the  two  electro- 
chemical cells  shown  in  Figure  12-4  (p.  204). 
What  change  will  occur  in  the  voltage  in  each 
cell?  Explain. 

18.  Determine  the  oxidation  numbers  of  carbon  in 
the  compounds  carbon  monoxide,  CO,  carbon 
dioxide,  C02,  and  in  diamond. 


QUESTIONS    AND    PROBLEMS 


223 


19.  Determine  the  oxidation  number  of  uranium  in 
each  of  the  known  compounds:  U03)  U308) 
U2Ofi,  U02,  UO,  K2U04,  Mg2U207. 

20.  By  use  of  half-reactions,  give  a  balanced  equa- 
tion for  each  of  the  following  reactions: 

(a)  H202  +  I"  +  H+ 

gives  H20  +  I2 

(b)  Cr.,CV2  +  Fe+2  +  H+ 

gives  Cr+3  +  Fe+3  +  H20 

(c)  Cu  +  NtV  +  H+ 

gives  Cu+2  +  NO  +  H20 

(d)  MnOr  +  Sn+2  +  H+ 

gives  Mn+2  +  Sn+4  +  H20 

21.  By  use  of  oxidation  numbers,  give  a  balanced 
equation  for  each  of  the  following  reactions: 

(a)  HBr  +  H2S04       gives  S02  +  Br2  +  H20 


(b)  N03-  +  CI-  +  H+ 

gives  NO  +  Cl2  +  H.O 

(c)  Zn  +  NO3-  +  H+ 

gives  Zn+2  +  N02  +  H.O 

(d)  BrO"  gives  Br~  +  Br03" 

22.  Use  oxidation  numbers  to  balance  the  reaction 
between  ferrous  ion,  Fe+2,  and  permanganate 
ion,  MnO^~,  in  acid  solution  to  produce  ferric 
ion,  Fe+3,  and  manganous  ion,  Mn+2. 

23.  Show  the  arbitrariness  of  oxidation  numbers  by 
balancing  the  reaction  discussed  in  Problem  22 
with  the  assumption  that  the  oxidation  number 
of  manganese  in  Mn04~  is  +2.  Compare  with 
the  result  obtained  in  Problem  22. 

24.  In  order  to  make  NafsJ  and  C\2(g),  an  electric 
current  is  passed  through  NaCl(l).  What  does 
the  energy  supplied  to  this  reaction  do? 


CHAPTER 


13 


Chemical 
Calculations 


The  sceptical  chemist  •  •  •  draws  conclusions  regarding  chemical  ma- 
terials •  •  •  chiefly  on  the  basis  of  quantitative  chemical  analysis 
which   •    •    •    is  the  touchstone  of  all  chemical  hypothesis. 

G.     T.     MORGAN,     1930 


Chemistry  is  a  quantitative  science.  This  means 
that  a  chemist  wishes  to  know  more  than  the 
qualitative  fact  that  a  reaction  occurs.  He  must 
answer  questions  beginning  "How  much  .  .  .?" 
The  quantities  may  be  expressed  in  grams,  vol- 
umes, concentrations,  percentage  composition, 
or  a  host  of  other  practical  units.  Ultimately, 
however,  the  understanding  of  chemistry  re- 
quires that  amounts  be  related  quantitatively  to 
balanced  chemical  reactions.  The  study  of  the 
quantitative  relationships  implied  by  a  chemical 
reaction  is  called  stoichiometry. 

Stoichiometric  calculations  are  based  upon 
two  assumptions.  First,  we  assume  that  only  a 
single  reaction  need  be  considered  to  describe  the 
chemical  changes  occurring.  Second,  we  assume 
that  the  reaction  is  complete.  For  example,  con- 
sider the  question,  How  much  iron  is  produced 
per  mole  of  Fe203  reacted  with  aluminum  in  the 
following  reaction? 


2Al(.sJ  +  Fe203(sj 


Al03(s)  +  2Fe(s)    (/) 


We  base  our  calculations  upon  the  two  assump- 
tions above,  whether  they  are  stated  or  not.  First, 
we  assume  this  is  the  only  reaction  that  occurs. 
Reaction  (2),  for  example,  is  assumed  to  be  un- 
important. 

2AlfsJ  +  3Fe203fsj  q=^r  6FeO(s)  +  Al203(sj    (2) 

Sometimes  such  an  assumption  is  based  upon 
experience,  sometimes  upon  hope.  Furthermore, 
it  is  assumed  that  a  mole  of  Fe203  reacts  com- 
pletely according  to  reaction  (/).  None  of  the 
Fe203  remains  unreacted  at  the  finish,  either  be- 
cause of  equilibrium,  because  of  mechanical 
losses  of  some  sort,  or  because  insufficient  alumi- 
num was  added.  In  practice  these  conditions  are 
sometimes  difficult  to  obtain. 

If  these  assumptions  are  valid,  however,  stoi- 
chiometric calculations  provide  a  reliable  basis 
for  quantitative  predictions.  It  is  important  to  be 
able  to  make  these  calculations  with  ease.  Fortu- 
nately, they  all  can  be  made  with  a  single  pattern 
based  upon  the  mole  concept. 


224 


SEC.     13-1    I    A    PATTERN    FOR    STOICHIOMETRIC    CALCULATIONS 


225 


13-1    A  PATTERN  FOR  STOICHIOMETRIC  CALCULATIONS 


The  equation  for  a  chemical  reaction  speaks  in 
terms  of  molecules  or  of  moles.  It  contains  the 
basis  for  stoichiometric  calculations.  However, 
in  the  laboratory  a  chemist  measures  amounts  in 
such  units  as  grams  and  milliliters.  The  first  step 
in  any  quantitative  calculation,  then,  is  to  con- 
vert the  measured  amounts  to  moles.  In  mole 
units,  the  balanced  reaction  connects  quantities 
of  reactants  and  products.  Finally,  the  result  is 
expressed  in  the  desired  units  (which  may  not 
necessarily  be  the  same  as  the  original  units). 


Let's  put  this  down  schematically.  Suppose 
two  substances,  A  and  B,  combine  according  to 
a  known  reaction.  We  wish  to  know  how  much 
B  will  react  with  (or,  be  produced  from)  a  meas- 
ured quantity  of  A.  The  solution  to  this  typical 
problem  of  stoichiometry  consists  of  three  steps. 

We  shall  apply  this  scheme  to  a  series  of  types 
of  calculations  to  show  its  general  applicability. 
The  calculations  are  all  connected  with  the  man- 
ufacture of  sulfuric  acid,  H2S04,  one  of  the  most 
important  commercial  chemicals. 


Step  I. 


Amount  of  A  in 
measured  units 


Convert  to  moles  of  A 
Mol  wt  of  A 


Moles  of  A 


(3) 


Step  II. 


Moles  of  A 


Convert  to  moles  of  B 
Balanced  reaction 


Moles  of  B 


(4) 


Step  III. 


Moles  of  B 


Convert  to  amount  of  B 
Mol  wt  of  B 


Amount  of  B  in 
desired  units 


(5) 


13-2    THE  MANUFACTURE  OF  SULFURIC  ACID 


Many  millions  of  tons  of  sulfuric  acid,  H2S04, 
are  produced  every  year.  Its  uses  are  so  wide  that 
the  amount  consumed  per  year  by  a  country  can 
be  taken  as  a  crude  index  of  the  technological 
development  of  that  country.  Two  manufactur- 
ing processes  have  widespread  industrial  im- 
portance and  both  will  be  described.  These 
processes  are  so  highly  perfected  that  the  cost 
of  this  useful  chemical  is  only  about  S22.00  per 
ton! 


EXERCISE  13-1 

If  H2S04  is  purchased  at  a  price  of  $22.00  per 
ton,  how  many  moles  are  obtained  for  a  penny? 
(Note:  1  pound  =  453.6  grams.) 


The  chemical  reactions  appear  simple.  They 
begin  with  pure  sulfur  (which  occurs  in  natural 
deposits  in  the  elemental  state).  First,  sulfur  is 
burned  to  give  gaseous  sulfur  dioxide,  S02.  Next, 
the  S02  is  further  oxidized,  catalytically,  to  sulfur 
trioxide,  SO.t.  Finally,  addition  of  water  forms 
sulfuric  acid.  The  reactions  are: 


8so,o;  +  40/gj 

8SO,teJ  +  8H,0(IJ 


8SO/*J 


(<5) 


=±  8so,oj 

(catalyst  needed)    (7) 


8H>S04(7J 


(8) 


Overall  reaction 

Ss(s)+  I202fgj  +  8H,.0(/J 


8H.S04(7J    (9) 


Now  let's  investigate  some  of  the  quantitative 
questions  that  are  connected  with  this  important 
process. 


226 


CHEMICAL  CALCULATIONS  I  CHAP.  13 


13-2.1    Weight-Weight  Calculations 

A  shovelful  of  sulfur  containing  1 .00  kg  is  placed 
in  the  hopper  to  be  converted  to  sulfuric  acid. 
What  weight  of  H2SOA  will  be  formed? 

This  practical  question  is  of  a  familiar  form. 
We  wish  to  calculate  the  weight  of  product  re- 
sulting from  a  specified  weight  of  reactant.  Our 
calculational  pattern  is  applicable. 

First,  a  reaction  must  be  assumed.  According 
to  the  intent  of  the  process,  the  overall  reaction 
is  (9): 


Srfs)  +  I202(g)  +  8H2Of/j 


8H2SO//j    (9,70) 


Our  calculation  proceeds  on  the  assumption  that 
equation  (70)  is  the  only  reaction  that  occurs  and 
that  the  entire  kilogram  of  sulfur  is  consumed 
in  it. 

Step  I.  We  must  convert  grams  of  sulfur  to 
moles.  The  molecular  weight  of  sulfur  is 
needed. 

Mol  wt  Sg  =  8  X  (atomic  wt  of  sulfur) 

=  8  X  (32.1)  =  256.8  g  mole    (77) 

Now  the  number  of  moles  of  S8  in  1 .00  kg  of 
sulfur  is 

Moles  S    =  (wt  sulfur)  =  (1-00  X  10»g) 
8        (mol  wt)        (256.8  g  mole) 


=  3.89  moles 


(12) 


Step  II.  The  next  step  is  to  decide  how  many 
moles  of  H2SO4  can  be  produced  from  3.89 
moles  of  S8.  The  balanced  reaction  (70)  tells 
us  that 

one  mole  of  S8  forms      eight      moles  H2SO4 

hence 

3.89  moles  of  S8  form  8  X  (3.89)  moles  H2S04 
3.89  moles  of  S8  form      37.7       moles  H2SOA  (13) 

Step  III.  The  amount  of  sulfuric  acid  formed  is 
31.1  moles.  How  much  does  this  weigh? 

Wt  H2S04  =  (moles  H2S04Xmol  wt  of  H2S04) 
=  (31.1  moles)(98.1  g/mole) 
=  3051  g 

Wt  H,S04  =  3.05  kg 

Answer.  1.00  kg  of  sulfur  produces  3.05  kg  H2SO4 
by  reaction  (10).  (14) 


13-2.2    Weight-Gas  Volume  Calculations 

The  first  step  in  the  manufacture  of  H2S04  is  to 
burn  sulfur  to  sulfur  dioxide.  Sulfur  burns  spon- 
taneously in  air,  liberating  heat. 

S*(s)  +  802(g)  z^±  SSOJg) 

AH  =  -70.96  kcalAnole  S02    (5),  (75) 

An  important  economic  feature  of  the  modern 
processes  is  the  utilization  of  this  heat  in  another 
step  in  which  heat  is  absorbed. 

Of  course,  a  plant  designer  must  anticipate  the 
weights  and  volumes  of  the  constituents  at  each 
stage  of  the  process.  Hence,  he  must  be  able  to 
answer  such  a  question  as,  "What  weight  of  sulfur 
will  burn  to  produce  100,000  liters  of  pure  SO2  at 
500° C  and  one  atmosphere  pressure  ?" 

Again,  we  must  assume  reaction  (75)  can  be 
carried  out  exclusively  and  completely.  This  time 
the  calculation  begins  with  a  specified  amount  of 
a  product  and  we  wish  to  calculate  the  corre- 
sponding amount  of  a  reactant.  Note  that  we  are 
immediately  confronted  with  a  question  of  sig- 
nificant figures.  How  many  significant  figures  are 
intended  in  the  volume,  100,000  liters?  In  the 
absence  of  other  information,  let's  let  common 
sense  dictate.  Would  it  be  of  value  to  make  the 
calculation  to  six  significant  figures?  Undoubt- 
edly not.  Two,  or  at  most  three,  significant  fig- 
ures will  probably  suffice  for  the  purposes  of  the 
plant  designer.  Other  conditions,  such  as  the 
weight  of  sulfur  and  the  temperature,  can't  be 
easily  controlled  to  more  than  this  accuracy. 
Let's  carry  three  significant  figures  to  be  sure  we 
have  enough.  The  volume  of  S02  is,  then, 
1.00  X  105  liters.* 

Step  I.  We  must  convert  the  specified  volume  of 
S02  into  moles.  A  convenient  way  to  do  this 
is  to  calculate  the  volume  this  gas  would  oc- 
cupy under  conditions  at  which  we  know  the 
volume  occupied  by  one  mole  of  gas.  For  ex- 
ample, we  know  that  one  mole  of  gas  occupies 
22.4  liters  at  0°C  and  one  atmosphere  pressure. 
Increasing  the  temperature  of  a  gas  at  constant 


*  This  volume,  1.00  X  105  liters,  is  about  the  volume 
of  a  small  room — a  practical  dimension  for  a  reaction 
chamber. 


SEC.     13-2    I    THE    MANUFACTURE    OF    SULFURIC    ACID 


227 


pressure  increases  the  volume  in  proportion  to 
the  absolute  temperature: 

0°C  =  273°K 
500°C  =  273  +  500  =  773°K 
1.00  mole  of  gas  occupies  22.4  liters 

at  273°K,  1  atm 
1.00  mole  of  gas  occupies  22.4  X  m  liters 

at  773°K,  1  atm 
1.00  mole  of  gas  occupies  63.4  liters 
at  773°K,  1  atm 

or 

63.4  liters  S02  at  773°K,  1  atm  contain 
1.00  mole  S02 

hence 
1.00  X  10s  liters  contain 

1'°°  *  10&  X  1.00  mole  S02 
63.4 

1.00  X  105  liters  contain 

1.58  X  103  moles  S02  at  773°K,  1  atm    (16) 

Step  II.  Next  we  decide  how  many  moles  of  S8 
are  needed  to  produce  1.58  X  103  moles  of 
S02.  The  balanced  reaction  (75)  indicates 

8  moles  S02  are  produced  from  1  mole  S8 
1  mole  S02  is  produced  from  £  mole  S8 
1.58  X  103  moles  S02  are  produced  from 
Kl-58  X  103)  moles  S8 

7.55  X  10s  moles  S02  are  produced  from 
198  moles  Si  (77) 

Step  III.  How  much  do  198  moles  of  S8  weigh? 

Wt  S8  =  (moles  S8)(mol  wt  S8) 

=  (198  moles)(256.8  g/mole) 
Wt  S8  =  50.8  X  103  g 

Answer.  1.00  X  705  liters  of  SOi  are  produced 
from  50.8  kilograms  of  sulfur  by  re- 
action (75).  (75) 

13-2.3    Gas  Volume-Gas  Volume  Calculations 

After  sulfur  dioxide  is  produced  by  combustion 
of  sulfur,  further  oxidation  is  needed  in  the 
manufacture  of  H2S04.  The  reaction,  producing 
sulfur  trioxide,  SO3,  is  exothermic;  heat  is  re- 
leased: 

SO,(g)  +  hO,(g)  ^=  SCVgj 

AH  =  -23.5  kcal/mole  S03    (7),  (79) 


Yet  the  reaction  is  quite  slow,  even  at  high  tem- 
peratures. Evidently  the  rate  is  controlled  by  a 
high  activation  energy.  In  fact,  the  practical  use 
of  reaction  (79)  depends  upon  the  presence  of  a 
catalyst  to  provide  a  reaction  path  with  a  lower 
activation  energy.  The  two  important  commer- 
cial methods  for  manufacture  of  H2S04  differ 
principally  in  the  choice  of  catalyst  for  this  step. 

The  older  process  is  called  the  lead  chamber 
process.  It  uses  a  mixture  of  gaseous  oxides  of 
nitrogen — nitric  oxide,  NO,  and  nitrogen  diox- 
ide, N02 — as  the  catalyst.  This  process  has  been 
in  use  and  under  development  for  over  200  years. 
It  is  named  after  the  large  room-like  chambers 
lined  with  lead  in  which  the  gaseous  reactions 
are  carried  out.  The  lead  walls  react  with  the 
acid  and  become  coated  with  an  inert  protective 
coating  of  lead  sulfate. 

The  newer  process  uses  a  solid  catalyst  for 
reaction  (79).  Either  finely  divided  platinum  or 
vanadium  pentoxide,  V2Os,  is  effective.  Because 
catalysis  occurs  where  the  gas  contacts  the  sur- 
face of  the  catalyst,  this  process  is  called  the 
contact  process. 


EXERCISE  13-2 

Reaction  (79)  is  carried  out  at  a  high  temperature 
(about  500°C  in  the  contact  process).  How  does 
temperature  affect  equilibrium,  according  to  Le 
Chatelier's  Principle?  In  view  of  your  answer, 
propose  an  explanation  of  why  the  temperature 
is  kept  high. 


Reaction  (79)  requires  the  reaction  of  oxygen 
from  air  and  sulfur  dioxide.  What  volume  of  air, 
at  500°C  and  one  atmosphere  pressure,  is  needed 
to  react  with  the  1.00  X  70s  liters  of  S02  produced 
from  50.8  kilograms  of  sulfur? 

Step  I.  In  considering  a  chemical  reaction  be- 
tween gases,  we  can  apply  Avogadro's  Hy- 
pothesis: Equal  volumes  of  gases  contain 
equal  numbers  of  molecules  (at  the  same  pres- 
sure and  temperature).  The  volume  of  the  S02, 
1.00  X  10s  liters,  is  already  a  measure  of  the 
number  of  moles  of  S02. 


228 


CHEMICAL    CALCULATIONS    |    CHAP.     13 


H2S0+    + 
catalyst (NV.N02)  Steam 


Lead- lined  chamber- 


Air 


NO+NOz  + 
H20 


Catalyst 
recovery 


H1SQ4  prodi 


FUNCTION: 


Burn    sulfur 

S6  +  802  —  8502 


Catalyst   (NO, 
N02)  added 


Fig.  75-7.  The  "lead  chamber"  process  for  H:SO,  man- 
ufacture. 

Step  II.  By  reaction  (79),  1  mole  of  S02  reacts 
with  \  mole  of  02.  By  Avogadro's  Hypothesis, 
1  liter  of  S02  reacts  with  \  liter  of  02  (if  they 
are  at  the  same  temperature  and  pressure). 
Hence  1.00  X  105  liters  of  S02  react  with 
1(1.00  X  105)  liters  of  02  if  both  S02  and  02 
are  pure  and  measured  at  500°C  and  one  at- 
mosphere. 

Answer.  We  need  0.500  X  10h  liters  of  pure  0>  at 
500°C  and  one  atmosphere  to  oxidize 
LOO  X  105  liters  of  S02.  (20) 

Step  III.  Now  we  must  convert  to  the  desired 
units.  We  need  the  number  of  moles  of  oxygen 
that  are  present  in  0.500  X  105  liters  (at  500°C, 
1  atm),  but  we  wish  to  use  air  instead  of  pure 
oxygen.  If  air  contains  about  20%  oxygen  (by 
volume),  then  it  takes  5  liters  of  air  (at  500°C, 


Ca  ta  ly  tic    oxida  tion 
of  S02j   reaction  yvith  H20 

S02  +  !io2 — -sos 

S03  +  H20 H2S04 

1   atm)  to  provide  the  amount  of  oxygen  in 

1  liter  of  pure  oxygen  (at  500°C,   1   atm).  If 

0.500  X  105  liters  of  pure  02  are  needed,  then 

5(0.500  X  105)  =  2.50  X  105  liters  of  air  are 

needed. 

Answer.  2.50  X  10h  liters  of  air  (at  500°  C,  1  atm) 
react  with  1.00  X  10h  liters  of  S02  (at 
500°C,  1  atm).  (21) 

13-2.4    Weight-Liquid  Volume  Calculations 

The  last  step  in  the  preparation  of  commercial 
sulfuric  acid  is  to  allow  the  sulfur  trioxide  to 
react  with  steam: 

S03(g)  +  H20(g)  +±  H,SCVZJ    (9),  (22) 

This  results  in  a  concentrated  sulfuric  acid  solu- 
tion that  contains  98%  H2S04.  It  is  a  viscous, 
colorless  liquid.  When  it  is  mixed  with  water,  so 
much  heat  is  liberated  that  the  operation  must 
be  carried  out  very  cautiously.  The  sulfuric  acid 


SEC.     13-2    |    THE    MANUFACTURE    OF    SULFURIC    ACID 


229 


is  slowly  poured  into  the  water,  not  the  reverse. 
The  density  of  this  concentrated  sulfuric  acid 
solution  (98%)  is  1.84  grams/ml  and  its  con- 
centration is  18.3  M. 

One  of  the  important  uses  of  sulfuric  acid  is 
that  of  an  oxidizing  agent.  For  example,  when 
heated,  it  will  even  dissolve  carbon.  The  reac- 
tion is 

C  +  2H,.S04  +±  CO.  +  2H20  +  2S02    (23) 


EXERCISE  13-3 

Verify  that  reaction  (23)  is  an  oxidation-reduc- 
tion reaction  and  that  the  oxidation  number 
change  of  carbon  is  balanced  by  the  oxidation 
number  change  of  the  sulfur. 


How  many  liters  of  concentrated  sulfuric  acid 
would  be  consumed  in  reaction  (23)  to  oxidize 
J. 00  kg  of  carbon  ? 

Fig.   13-2.  The  "contact"  process  for  HjSOt 

manufacture. 


Step  I.  We  wish  to  oxidize  1.00  kg  of  carbon. 
The  number  of  moles  of  carbon  is 

K.   ,  .  (wt  C)         1.00  X  10>g 

Moles  carbon  =  r1 £-  =  — — — ; — f 

(at  wt  C)       12.01  g/mole 

=  83.2  moles 

Moles  carbon  =  83.2  (24) 

Step  II.  Now  we  must  decide  how  many  moles 
of  sulfuric  acid  are  needed.  Reaction  (23) 
shows  that 

1  mole  C  reacts  with  2  moles  H2S04 

83.2  moles  C  react  with  2(83.2)  moles  H2S04 

83.2  moles  C  react  with  166  moles  H£0A         (25) 

Step  III.  We  need  166  moles  of  H2S04.  What 
volume  of  concentrated  sulfuric  acid  (18.3  M) 
is  needed? 

18.3  moles  H2S04  are  present  in  1.00  liter 
98  %  H2S04 

166  moles  H2S04  are  present  in  — —  liters 

98  %  H2S04 
166  moles  H>SOA  are  present  in  9.07  liters 


A  SQ** 

S02  +  02+  impurities  Jy         02+  H2 


H2S04 


S02 


Air 


FUNCTION :         Burn,    sulfur.  Remove   impurities 

Ss  +  BOj — 8S02  that  might 

"poison."  catalyst 


H2504  +  H20 


Dry 
S02  -  02 
mixture 


H2S0^-S05  . 
"fuming" sulfuric   acid 

Catalyticatly       Absorb  S03(y) 
burn.  S02  ,  in  H2SO+ 

S02  +%o2~so3 


230 


CHEMICAL    CALCULATIONS    I    CHAP.     13 


Answer.  1.00  kg  of  carbon  is  oxidized  by  9.07 
liters  of  concentrated  HtSO*  in  reac- 
tion (23).  (26) 

13-2.5     Liquid  Volume-Volume  Calculations 

A  second  major  use  of  sulfuric  acid  of  commerce 
is  in  reactions  with  bases.  In  laboratory  use  it  is 
diluted  to  a  much  lower  concentration  and  can 
be  used  as  a  standard  acid.  A  typical  problem 
would  be  the  titration  of  a  base  solution  of 
unknown  concentration  using  a  sulfuric  acid 
solution  of  known  concentration.  For  example, 
"What  is  the  concentration  of  a  sodium  hydroxide 
solution  if  25.43  ml  of  the  NaOH  solution  just 
reacts  with  18.51  ml  of  0.1250  M  H^SOi  (to  pro- 
duce a  neutral  solution)  ?" 

Step  I.  We  are  given  the  concentration  and  vol- 
ume of  H2S04  solution.  How  many  moles  of 
H2SO4  are  present? 

Moles  H0SO4 
present        =  (volume)(concentration) 

=  (18.51  X  10-3  liter)(0.1250  NT) 
Moles  H2S04  =  2.314  X  10-3  mole 

Step  II.  We  are  interested  in  the  reaction  be- 
tween H+(aq)  and  OW~(aq).  Sulfuric  acid 
gives  2  moles  of  H+(aq)  per  mole  of  H2S04 
dissolved  in  water. 


Hence 


2H+(aqj  +  SOr*(aq)       (27) 


Moles  H~(aq)  =  2(moles  H,S04) 
=  2(2.314  X  10-') 
Moles  H+(aq)  =  4.628  X  10~3  (28) 

Now  we  are  concerned  with  the  acid-base 
reaction 


H+(aq)  +  OH-(aq) 

By  reaction  (29), 


H20 


(29) 


one  mole  of  OH~(aq)  ion  reacts  with  one 
mole  of  H+(aq)  ion.  4.628  X  10~3  mole  of 
OH~(aq)  ion  reacts  with  4.628  X  10~3  mole  oj 
H+(aq)  ion.  (30) 

Step  III.  We  can  now  calculate  the  concentra- 
tion of  hydroxide  ion.  We  now  know  that 
4.628  X  10-3  mole  of  OR-(aq)  is  contained  in 
25.43  ml  of  sodium  hydroxide  solution: 


Concentration 
hydroxide  ion 


(moles  OH~  ion) 
(volume) 

roH_1  =  (4.628  X  10~3  mole) 
L         J       (25.43  X  10-3  liter) 

[OH-]  =  0.1820  M 

Answer.  The  sodium  hydroxide  solution  has  a  con- 
centration of  0.1820  M.  (31) 


QUESTIONS  AND  PROBLEMS 

1 .  In  Experiment  7,  would  the  ratio  between  moles 
of  copper  atoms  used  and  moles  of  silver  atoms 
formed  change  if  silver  sulfate,  Ag2S04,  had  been 
used  rather  than  silver  nitrate,  AgN03?  Explain. 

2.  Although  sodium  carbonate  is  needed  in  the 
manufacture  of  glass,  very  little  is  found  in 
nature.  It  is  made  using  two  very  abundant 
chemicals,  calcium  carbonate  (marble)  and  so- 
dium chloride  (salt).  The  process  involves  many 
steps,  but  the  overall  reaction  is 

CaC03  +  2NaCl  — -»-  Na2C03  +  CaCl2 

(a)  How  many  grams  of  sodium  chloride  react 
with  1.00  kg  of  calcium  carbonate? 


(b)  How  many  grams  of  sodium  carbonate  are 
produced? 

3.  Some  catalysts  used  in  gasoline  manufacture 
consist  of  finely  divided  platinum  supported  on 
an  inert  solid.  Suppose  that  the  platinum  is 
formed  by  the  high  temperature  reaction  between 
platinum  dioxide,  Pt02,  and  hydrogen  gas  to 
form  platinum  metal  and  water. 

(a)  What  is  the  oxidation  number  of  platinum 
in  platinum  dioxide? 

(b)  Is  hydrogen  an  oxidizing  or  reducing  agent 
in  this  reaction? 

(c)  How  many  grams  of  hydrogen  are  needed  to 
produce  1.0  gram  of  platinum  metal? 


QUESTIONS    AND    PROBLEMS 


231 


(d)  How  many  moles  of  water  are  produced 
along  with  1.0  gram  of  Pt? 

(e)  How  many  grams  of  water  are  produced 
along  with  1.0  gram  of  Pt? 

Answer,  (e)  0.18  gram  of  H:0 

4.  Hydrazine,  NjH4,  and  hydrogen  peroxide,  H.Oj, 
are  used  together  as  a  rocket  fuel.  The  products 
are  N2  and  H^O.  How  many  grams  of  hydrogen 
peroxide  are  needed  per  1.00  X  103  grams  of 
hydrazine  carried  by  a  rocket  ? 

Answer.  2.12  X  103  grams  of  H02 

5.  Iodine  is  recovered  from  iodates  in  Chile  salt- 
peter by  the  reaction 

HS03-  +  IO3-  gives  I,  +  SO4"2  +  H+  +  HO 

(a)  How  many  grams  of  sodium  iodate,  NaI03, 
react  with  1.00  mole  of  KHSO3? 

(b)  How  many  grams  of  iodine,  I2,  are  pro- 
duced ? 

6.  The  hourly  energy  requirements  of  an  astronaut 
can  be  satisfied  by  the  energy  released  when  34 
grams  of  sucrose  are  "burned"  in  his  body.  How 
many  grams  of  oxygen  would  need  to  be  carried 
in  a  space  capsule  to  meet  this  requirement? 

sucrose   +  oxygen  gives  carbon  dioxide  +  water 
CnH«Ou+     02     gives  C02  +  H.O 

7.  The  chlorine  used  to  purify  your  drinking  water 
was  possibly  made  by  electrolyzing  molten 
NaCl  to  produce  liquid  sodium  and  gaseous 
chlorine. 

(a)  How  many  grams  of  sodium  chloride  are 
needed  to  produce  355  grams  of  chlorine 
gas? 

(b)  What  volume  would  this  gas  occupy  at  STP? 

8.  A  reaction  involved  in  the  production  of  iron 
from  iron  ore  is 

Fe203  +  CO  gives  Fe  +  C02 

AH  =  -4.3  kcal/mole  Fe2Os 

(a)  How  many  grams  of  CO  must  react  to  re- 
lease 13  kcal? 

(b)  How  many  liters  of  CO(STP)  are  needed  to 
produce  1.0  kg  of  Fe? 

9.  More  C8Hi8,  a  hydrocarbon  that  is  useful  in 
gasoline,  can  be  obtained  from  petroleum  if  this 
reaction  takes  place: 

Ci6H32(g)  +  mjg)  — *-  2C8H18fgJ 


(a)  How  many  grams  of  C8Hi8  can  be  made 
using  224  liters  of  H2  at  STP? 

(b)  What  pressure  conditions  favor  production 
ofQH^gJ? 

Answer,  (a)  1.14  X  103  g  CgH18 

10.  How  many  liters  of  oxygen  gas,  at  STP,  will  be 
released  by  decomposing  14.9  grams  of  NaOCl 
to  produce  02(g)  +  C\~(aq)  (as  in  Experiment 
14a)? 

11.  A  compound  found  in  kerosene,  a  mixture  of 
hydrocarbons,  is  decane,  Q0H22.  A  stove  might 
burn  1 .0  kg  of  kerosene  per  hour.  Assume  kero- 
sene is  Q0H22  and  answer  the  following: 

(a)  How  many  liters  (STP)  of  oxygen  are  needed 
per  hour? 

(b)  How  many  liters  (STP)  of  carbon  dioxide 
are  produced  per  hour? 

12.  How  many  grams  of  zinc  metal  are  needed  to 
react  with  hydrochloric  acid  to  produce  enough 
hydrogen  gas  to  fill  an  11.2  liter  balloon  at  STP? 
What  would  be  the  volume  of  this  balloon  at 
27°C  and  680  mm  Hg  pressure?  How  many 
grams  of  zinc  would  be  needed  if  sulfuric  acid 
were  used  ? 

1 3.  How  many  liters  of  air  (STP)  are  needed  to  burn 
2.2  liters  (STP)  of  methane,  CH4,  gas  in  your 
laboratory  burner?  How  much  heat  is  released? 
The  AH  for  combustion  of  CH4  is  —210 
kcal  mole  of  CH4.  Assume  air  is  20  %  oxygen. 

14.  In  the  reaction 

NHafeJ  +  02(g)  gives  NO(g)  +  H20(g), 
if  4.48  liters  of  ammonia  gas  measured  at  STP 
are  used,  how  many  liters  of  oxygen  measured 
at  STP  will  be  needed  to  react  with  all  the 
ammonia? 

Answer.  5.60  liters  of  02  at  STP 

15.  The  following  reaction  is  carried  out  with  all  gas 
volumes  measured  at  the  same  pressure  and  tem- 
perature: 

CMl0(g)  +  02(g)  gives  C02(g)  +  H20(g) 

(a)  How  many  liters  of  oxygen  are  required  to 
produce  2.0  liters  of  C02? 

(b)  If  15  liters  of  oxygen  are  used,  how  many 
liters  of  butane,  C4Hi0,  will  be  burned? 

(c)  If  8.0  liters  each  of  oxygen  and  butane  are 
mixed,  how  many  liters  of  CO.  are  produced 
(assume  complete  reaction)? 

Answer,  (a)  3.2  liters  of  0> 


232 


CHEMICAL    CALCULATIONS    I    CHAP.     13 


16.  What  volume  of  Cl2  gas  at  37°C  and  753  mm 
could  be  obtained  from  58.4  liters  of  HC1,  also 
measured  at  37°C  and  753  mm,  if  the  following 
reaction  could  be  carried  effectively  to  comple- 
tion? 

UC\(g)  +  02(g)  gives  H20(g)  +  C\2(g) 

17.  Suppose  105  liters  of  NH3  and  285  liters  of  02 
are  allowed  to  react  until  the  reaction : 

NH3teJ  +  02(g)  gives  H20(g)  +  NO/gj 

is  complete.  The  temperature  and  pressure  are 
maintained  constant  at  200°C  and  0.30  atmos- 
phere during  all  volume  measurements.  What 
gas  and  what  volume  of  it  measured  at  the  stated 
conditions  remains  unreacted? 

18.  A  6  volt  lead  storage  battery  contains  700  grams 
of  pure  H2S04(  I)  dissolved  in  water. 

(a)  How  many  grams  of  solid  sodium  carbonate, 
Na2C03,  would  be  needed  to  neutralize  this 
acid  (giving  C02  gas  and  H20)  if  it  were 
spilled  ? 

(b)  How  many  liters  of  2.0  M  Na2COs  solution 
would  be  needed? 

19.  Nitric  acid,  HN03,  is  made  by  the  process 

3NQ2(g)  +  H2OfIJ  ++  2HN03([j  +  NOfe) 


Commercial  concentrated  acid  contains  68  %  by 
weight  HN03  in  water.  The  solution  is  15  M. 
How  many  liters  of  concentrated  acid  are  needed 
to  react  with  0.100  kg  of  copper  metal? 

Cu(s)  +  H+(aq)  +  NO^(aq)  gives 

Cxx+Haq)  +  N02(gj  +  H20(/; 

Answer.  0.42  liter  of  HN03 

20.  How  many  grams  of  silver  metal  will  react  with 
2.0  liters  of  6.0  M  HN03?  The  reaction  is 

Ag(s)  +  H+(aq)  +  N03_  (aq)  gives 

Ag+(aq)  +  NO(g)  +  H20(JJ 

21.  A  measured  volume,  10.00  liters,  of  the  waste 
process  water  from  a  cotton  mill  require  23.62 
ml  of  0.1000  M  hydrochloric  acid  to  produce  a 
neutral  solution.  What  is  the  hydroxide  ion  con- 
centration in  the  waste  ? 

22.  What  weight  of  silver  chloride  may  be  obtained 
from  1 .0  liter  of  1 .0  M  AgN03,  if  1 2  ml  of  0. 1 5  M 
NaCl  are  added? 

23.  How  many  milliliters  of  a  0.050  M  KMn04  solu- 
tion are  required  to  oxidize  2.00  grams  of  FeS04 
in  a  dilute  acid  solution? 

Answer.  53  ml  of  KMn04  solution 


CHAPTER 


14 


Why  We  Believe 
in  Atoms 


From  the  time  of  Dalton  •  •  •  the  history  of  the  atom  has  been  a  march 
of  triumph.  Wherever  the  concept  of  the  atom  was  employed  for  the  in- 
terpretation of  observational  measurements,  it  supplied  lucid  explanation; 
conversely,  such  success  became  overwhelming  evidence  for  the  existence 
of  the  atom. 

HANS    REICHENBACH,     1951 


In  Chapter  2  you  were  introduced  to  atoms  and 
in  Chapter  6  they  were  described  in  more  detail. 
You  were  told  that  the  atom  contains  charged 
particles,  that  it  has  a  nucleus  made  up  of  neu- 
trons and  protons,  and  that  the  nucleus  is  sur- 
rounded by  electrons.  The  atom  is  incredibly 
small  but  the  nucleus  is  even  smaller.  But  also 
you  were  told  that  every  theory  (including  the 
atomic  theory)  should  be  thought  about  and 
criticized — the  evidence  upon  which  it  is  based 
should  be  examined  and  understood.  It  is  one 
thing  to  ask  "Do  we  believe  in  atoms?"  and 
quite  another  to  ask  "Why  do  we  believe  in 
atoms?"  In  this  chapter  we  shall  try  to  answer 
this  last,  harder  question. 

Let's  begin  with  an  unpretentious  example  that 
shows  how  we  make  such  decisions  in  day-to-day 
living. 

A  new  tenant  is  told  by  his  neighbor  that  the 
garbage  collector  comes  every  Thursday,  early  in 
the  morning.  Later,  in  answer  to  a  question  from 


his  wife  about  the  same  matter,  the  tenant  says, 
"I  have  been  told  there  is  a  garbage  collector  and 
that  he  comes  early  Thursday  morning.  We  shall 
see  if  this  is  true."  The  tenant,  a  scientist,  ac- 
cepts the  statement  of  the  neighbor  (who  has  had 
opportunity  to  make  observations  on  the  sub- 
ject). However,  he  accepts  it  tentatively  until  he 
himself  knows  the  evidence  for  the  conclusion. 

After  a  few  weeks,  the  new  tenant  has  made  a 
number  of  observations  consistent  with  the  ex- 
istence of  a  Thursday  garbage  collector.  Most 
important,  the  garbage  does  disappear  every 
Thursday  morning.  Second,  he  receives  a  bill 
from  the  city  once  a  month  for  municipal  serv- 
ices. And  there  are  several  supplementary  ob- 
servations that  are  consistent — often  he  is 
awakened  at  5:00  a.m.  on  Thursdays  by  a  loud 
banging  and  sounds  of  a  truck.  Occasionally  the 
banging  is  accompanied  by  gay  whistling,  some- 
times by  a  dog's  bark. 

The  tenant  now  has  many  reasons  to  believe 

233 


234 


WHY    WE    BELIEVE    IN    ATOMS    I    CHAP.     14 


in  the  existence  of  the  garbage  collector.  Yet  he 
has  never  seen  him.  Being  a  curious  man  and  a 
scientist,  he  sets  his  alarm  clock  one  Wednesday 
night  to  ring  at  5:00  a.m.  Looking  out  the  win- 
dow Thursday  morning,  his  first  observation  is 
that  it  is  surprisingly  dark  out  and  things  are 
difficult  to  see.  Nevertheless,  he  discerns  a 
shadowy  form  pass  by,  a  form  that  looks  like  a 
man  carrying  a  large  object. 

Seeing  is  believing!  But  which  of  these  pieces 
of  evidence  really  constitutes  "seeing"  the  gar- 
bage collector?  Which  piece  of  evidence  is  the 
basis  for  "believing"  there  is  a  garbage  collector? 
The  answer  is,  all  of  the  evidence  taken  together, 
constitutes  "seeing."  And  all  of  the  evidence 
taken  together,  furnishes  the  basis  for  accepting 
the  "garbage  collector  theory  of  garbage  dis- 
appearance." The  direct  vision  of  a  shadowy 
form  at  5:00  a.m.  would  not  constitute  "seeing 
a  garbage  collector"  if  the  garbage  didn't  dis- 
appear at  that  time.  (The  form  might  have  been 
the  paper  boy  or  the  milkman.)  Neither  would 
the  garbage  disappearance  alone  consist  of  "see- 
ing" the  garbage  collector.  (Perhaps  a  dog  comes 
by  every  Thursday  and  eats  the  garbage.  Remem- 
ber, a  dog's  bark  was  heard!)  No,  the  tenant  is 
convinced  there  is  a  garbage  collector  because 
the  assumption  is  consistent  with  so  many  ob- 
servations, and  it  is  inconsistent  with  none.  Other 
possible  explanations  fit  the  observations  too, 
but  not  as  well  (the  tenant  has  never  heard  a  dog 
whistle   gaily).   The   garbage   collector   theory 


passes  the  test  of  a  good  theory — it  is  useful  in 
explaining  a  large  number  of  experimental  ob- 
servations. This  was  true  even  before  the  tenant 
set  eyes  on  the  shadowy  form  at  5:00  a.m. 

Yet  we  must  agree,  there  are  advantages  to  the 
"direct  vision"  type  of  experiment.  Often  more 
detailed  information  can  be  obtained  this  way. 
Is  the  garbage  collector  tall?  Does  he  have  a 
mustache?  Could  the  garbage  collector  be  a 
woman?  This  type  of  information  is  less  easily 
obtained  from  other  methods  of  observation.  It 
is  worthwhile  setting  the  alarm  clock,  even  after 
we  have  become  convinced  there  is  a  garbage 
collector. 

At  the  beginning  of  this  course  you  were  a  new 
tenant.  You  were  told  that  chemists  believe  in 
atoms  and  you  were  asked  to  accept  this  proposal 
tentatively  until  you  yourself  knew  the  evidence 
for  it.  Since  that  time,  we  have  used  the  atomic 
theory  continuously  in  our  discussions  of  chemi- 
cal phenomena.  The  atomic  theory  passes  the 
test  of  a  good  theory:  it  is  useful  in  explaining  a 
large  number  of  experimental  observations.  We 
have  become  convinced  there  are  atoms. 

Now  we  are  going  to  review  the  types  of  evi- 
dence that  form  the  basis  for  belief  in  the  atomic 
theory.  We  shall  include  a  number  of  experi- 
ments that  are  close,  in  concept,  to  the  "direct 
vision"  type.  These  are  particularly  convincing 
and  they  provide  detailed  information  that  is  less 
readily  obtained  in  other  ways. 


14-1    CHEMICAL  EVIDENCE  FOR  THE  ATOMIC  THEORY 


Let  us  begin  by  looking  again  at  the  kinds  of 
evidence  we  already  have  for  the  existence  of 
atoms — the  evidence  from  chemistry.  We  shall 
consider,  in  turn,  the  definite  composition  of 
compounds,  the  simple  weight  relations  among 
compounds,  and  the  reacting  volumes  of  gases. 
Each  behavior  provides  experimental  support 
for  the  atomic  theory. 

14-1.1    The  Law  of  Definite  Composition 

Compounds  are  found  to  have  definite  composi- 
tion, no  matter  how  prepared.  For  example, 


2.016  grams  of  hydrogen  are  found  combined 
with  16.00  grams  of  oxygen  in  the  compound 
water  whether  the  water  is  prepared  by  burning 
hydrogen  in  oxygen,  by  decomposing  gaseous 
nitrous  acid,  by  heating  barium  chloride  dihy- 
drate,  or  by  some  other  process: 

H*(g)  +  hOJg)  — >■  H2OfeJ 

2.016  g  hydrogen/16.00  g  oxygen    (/) 

2HN02(gJ  — >■  NOfgJ  +  N02(g)  +  H2OfgJ 

2.016  g  hydrogen/ 16.00  g  oxygen    (2) 


SEC.    14-1    I    CHEMICAL    EVIDENCE    FOR    THE    ATOMIC    THEORY 


235 


/  molecule    0? 


I      mole        02 


32   g 


+ 

+ 
+ 

+ 


Z  molecules  H2 
Z     moles      H2 


1^"        w 

2    molecules  HzO 
2      moles       H30 


36    g 


M,0 


g   oxygen  _   5£  __fl 
g  hydr-ogen~    -f         1 


+ 

1  molecule   Oz 

+ 

1  molecule.  H2 

1     mole        02 

+ 

I    mole     Hz 

32    g           02 

+ 

2    g         Hz 

1  molecule  H202 
1  mole  H2Oz 
34    g       H2Oz 


g  oxygen    _  32  _  16 
g  hydrogen-   2    ~  1 

^BaCl2  2H20]fsJ  — >-  \BaCUJs)  +  U20(g)  Fig.  14-1.  Simple  multiple  proportions  of  oxygen  to 

2.016  g  hydrogen   16.00  g  oxygen     (5)  hydrogen  in  H.0  and  H4)» 

The  atomic  theory  provides  a  ready  explana- 
tion for  the  definite  composition  of  chemical 
compounds.  It  says  that  compounds  are  com- 
posed of  atoms,  and  every  sample  of  a  given 
compound  must  contain  the  same  relative  num- 
ber of  atoms  of  each  of  its  elements.  Since  the 
atoms  of  each  element  have  a  characteristic 
weight,  the  weight  composition  of  a  compound 
is  always  the  same.*  Thus,  the  definite  composi- 
tion of  compounds  provides  experimental  sup- 
port for  the  atomic  theory. 

14-1.2    The  Law  off  Simple  Multiple  Proportions 

In  many  cases,  two  elements  brought  together 
under  different  conditions  can  form  two  or  more 
different  compounds.  In  addition  to  water,  hy- 
drogen and  oxygen  can  form  a  compound  called 
hydrogen  peroxide,  H202,  in  which  the  weight  of 

*  This  statement  applies  if  the  naturally  occurring  dis- 
tribution of  isotopes  is  hot  disturbed. 


hydrogen  is  ^  the  weight  of  oxygen  in  the  com- 
pound. This  weight  ratio  in  water  is  £,  exactly 
twice  as  large.  What  a  simple  relationship!  The 
weight  ratio  is  one  to  sixteen  in  one  compound 
of  hydrogen  and  oxygen  and  one  to  eight  in 
another.  Such  simple  numerical  relationships  are 
always  found  among  different  compounds  of  a 
set  of  elements.  This  is  explained  very  clearly 
within  the  atomic  theory.  Each  molecule  of  hy- 
drogen peroxide  contains  two  atoms  of  hydrogen 
and  two  atoms  of  oxygen.  The  ratio  of  the  num- 
ber of  hydrogen  atoms  to  oxygen  atoms  is 
2/2  =  1.  In  contrast,  a  molecule  of  water  con- 
tains two  atoms  of  hydrogen  and  only  one  atom 
of  oxygen.  The  ratio  of  hydrogen  atoms  to 
oxygen  atoms  is  2/1  =  2.  Since  there  are  twice 
as  many  hydrogen  atoms  per  oxygen  atom  in 
water  as  in  hydrogen  peroxide,  of  course  the 
weight  ratio  of  hydrogen  to  oxygen  in  water  is 
twice  that  in  hydrogen  peroxide. 


236 


WHY    WE    BELIEVE    IN    ATOMS   I    CHAP.     14 


In  general,  different  compounds  of  the  same 
two  elements  have  different  atomic  ratios.  Since 
these  atomic  ratios  are  always  ratios  of  integers, 
1/1,  1/2,  2/1,  2/3,  etc.,  the  weight  ratios  will  be 
simple  multiples  of  each  other.  Thus  the  atomic 
theory  explains  the  observation  that  different 
compounds  of  the  same  two  elements  have  rela- 
tive compositions  by  weight  that  are  simple 
multiples  of  each  other. 

This  success  of  the  atomic  theory  is  not  sur- 
prising to  a  historian  of  science.  The  atomic 
theory  was  first  deduced  from  the  laws  of  chemical 
composition.  In  the  first  decade  of  the  nineteenth 
century,  an  English  scientist  named  John  Dalton 
wondered  why  chemical  compounds  display  such 
simple  weight  relations.  He  proposed  that  per- 
haps each  element  consists  of  discrete  particles 
and  perhaps  each  compound  is  composed  of 
molecules  that  can  be  formed  only  by  a  unique 
combination  of  these  particles.  Suddenly  many 
facts  of  chemistry  became  understandable  in 
terms  of  this  proposal.  The  continued  success  of 
the  atomic  theory  in  correlating  a  multitude  of 
new  observations  accounts  for  its  survival.  To- 
day, many  other  types  of  evidence  can  be  cited 
to  support  the  atomic  postulate,  but  the  laws  of 
chemical  composition  still  provide  the  corner- 
stone for  our  belief  in  this  theory  of  the  structure 
of  matter. 


EXERCISE  14-1 

Two  compounds  are  known  that  contain  only 
nitrogen  and  fluorine.  Careful  analysis  shows  that 
23.67  grams  of  compound  I  contain  19.00  grams 
of  fluorine  and  that  26.00  grams  of  compound 
II  contain  19.00  grams  of  fluorine. 

(a)  For  each  compound,  calculate  the  weight  of 
nitrogen  combined  with  19.00  grams  of  fluo- 
rine. 

(b)  What  is  the  ratio  of  the  calculated  weight  of 
nitrogen  in  compound  II  to  that  in  I? 

(c)  Compound  I  is  NF3.  This  compound  has  one 
atom  of  nitrogen  per  three  atoms  of  fluorine. 
How  many  atoms  of  nitrogen  are  there  per 
three  atoms  of  fluorine  for  each  of  the  mo- 
lecular formulas  N2F2  and  N2F4?  Compare 


these  atom  ratios  to  the  weight  ratio  ob- 
tained in  part  (b)  and  convince  yourself  that 
compound  II  could  have  the  formula  N2F4 
but  not  N2F2. 


14-1.3    The  Law  of  Combining  Volumes 

Gases  are  found  to  react  in  simple  proportions 
by  volume,  and  the  volume  of  any  gaseous  prod- 
uct bears  a  whole-number  ratio  to  that  of  any 
gaseous  reactant.  Thus,  two  volumes  of  hydrogen 
react  with  exactly  one  volume  of  oxygen  to  pro- 
duce exactly  two  volumes  of  water  vapor  (all  at 
the  same  temperature  and  pressure).  These  in- 
teger relationships  naturally  suggest  a  particle 
model  of  matter  and,  with  Avogadro's  Hypothe- 
sis, are  readily  explained  on  the  basis  of  the 
atomic  theory. 

Once  again  it  is  no  surprise  that  the  simple 
integer  volume  ratios  are  readily  explained  with 
the  atomic  theory.  The  atomic  theory  was  de- 
vised for  this  purpose,  as  is  indicated  in  Chap- 
ter 2. 

To  summarize,  we  find  that  the  weight  and 
volume  relations  that  are  observed  in  chemical 
changes  provide  an  experimental  foundation  for 
the  atomic  theory.  All  of  contemporary  chemical 
thought  is  based  upon  the  atomic  model  and, 
hence,  every  successful  chemical  interpretation 
strengthens  our  belief  in  the  usefulness  of  this 
theory. 

14-1.4    Chemical  Evidence  for  the  Electrical 
Nature  of  Atoms 

You  have  been  told  that  the  atomic  nucleus  bears 
a  positive  charge  and  is  surrounded  by  a  number 
of  negatively  charged  particles  called  electrons. 
Also,  the  nucleus  is  supposed  to  contain  most  of 
the  mass  of  the  atom  and  to  be  made  of  protons 
and  neutrons,  each  of  which  has  nearly  two 
thousand  times  the  mass  of  the  electron.  How 
do  we  know  that  atoms  are  built  this  way?  How 
do  we  know  that  there  is  such  a  particle  as  an 
electron?  Again,  weight  relations  associated  with 
chemical  reactions  provide  key  evidence. 
In  Chapter  12  we  discussed  the  operation  of 


SEC.     14-1     I    CHEMICAL    EVIDENCE    FOR    THE    ATOMIC    THEORY 


237 


Hs(l) 


Ha  +z(aq) 


Ma  CI  (l)  tn  molten  CaCl, 


T     IT-I'I'I1'1! 


Al203(l)  in    molten  Na3ALF6 


6 .  03  g Hg  produced 

6.03 

—  .0300  mole 


2  01 


1.  38  g  Ma  produced 

1.38 
3  Q    =    0600  mole 


.538   g  A.I  produced 
.538 


Z6.9 


.0200  mole 


Hg 

.0300 

3 


No. 

Al 

0600 

.0200 

6 

Z 

an  electrochemical  cell.  We  successfully  inter- 
preted the  chemical  changes  brought  about  by 
the  movement  of  electric  charge  in  terms  of  the 
atomic  theory.  To  understand  the  full  impact  of 
these  experiments  on  the  development  of  the 
atomic  theory,  we  must  turn  back  the  scientific 
clock  to  the  views  held  in  the  nineteenth  century. 
When  Michael  Faraday  first  performed  his  elec- 
trolysis experiments  (in  the  early  1830's),  the 
atomic  theory  had  been  proposed  but  no  one  had 
yet  suggested  the  existence  of  electrons.  There 
was  no  reason  to  suspect  that  electricity  con- 
sisted of  individual  units.  Faraday  observed  that 
the  quantity  of  electricity  necessary  to  deposit  a 
given  weight  of  an  element  from  solutions  of  its 
different  compounds  was  always  equal  to  a  con- 
stant, or  some  simple  multiple  of  this  constant. 
For  example,  the  amount  of  electricity  that  will 
deposit  6.03  grams  of  metallic  mercury  from  a 
solution  of  mercuric  perchlorate,  Hg(CIO^)2,  will 
deposit  the  same  number  of  grams  of  mercury 
from  a  solution  of  mercuric  nitrate,  Hg(N03)2. 


Fig.   14-2.  Weights  of  different  elements  deposited  by 

a  given  amount  of  electricity. 


On  the  other  hand,  this  same  amount  of  elec- 
tricity will  deposit  exactly  twice  as  much  mer- 
cury, 2  X  (6.03)  =12.1  grams,  from  a  solution 
of  mercurous  perchlorate,  Hg2(004)2.  If  we  re- 
state Faraday's  experimental  finding  in  terms  of 
the  atomic  theory,  we  see  that  the  number  of 
atoms  of  mercury  deposited  by  a  certain  quan- 
tity of  electricity  is  a  constant  or  a  simple 
multiple  of  this  constant.  Apparently  this  certain 
quantity  of  electricity  can  "count"  atoms.  A 
simple  interpretation  is  that  there  are  "packages" 
of  electricity.  During  electrolysis,  these  "pack- 
ages" are  parcelled  out,  one  to  an  atom,  or  two 
to  an  atom,  or  three. 

The  second  of  Faraday's  observations  was  that 
the  weights  of  different  elements  that  were  de- 
posited by  the  same  amount  of  electricity  formed 
simple  whole-number  ratios  when  divided  by  the 
atomic  weights  of  these  elements.  For  example, 


238 


WHY    WE    BELIEVE    IN    ATOMS   I    CHAP.    14 


suppose  electric  current  is  passed  through  the 
three  electrolysis  cells  pictured  in  Figure  14-2. 
The  two  ammeters  have  the  same  reading,  show- 
ing that  the  current  entering  the  cell  at  the  right 
is  identical  to  that  leaving  the  cell  at  the  left. 
Thus,  the  electric  circuit  guarantees  that  the  same 
amount  of  electricity  passes  through  each  of  the 
three  cells. 

In  the  first  cell  the  net  reaction  is  the  produc- 
tion of  metallic  mercury  and  gaseous  oxygen 
through  electrolysis  of  aqueous  mercuric  nitrate: 

2Hg+*(aq)  +  2H20  — *- 

2Hg(l)  +  02(g)  +  4U+(aq)     (4) 

After  current  has  passed  through  the  cell  for  a 
definite  time,  the  weight  of  the  mercury  produced 
is  found  to  be  6.03  grams. 

In  the  second  cell  molten  sodium  chloride  is 
electrolyzed.*  The  net  reaction  is 

NaClflJ  —^  Naflj  +  \&jg)  (5) 

The  same  current  that  produced  6.03  grams  of 
mercury  is  found  to  produce  1.38  grams  of 
molten  sodium. 

The  third  cell  represents  another  industrial 
process,  the  electrolytic  process  for  manufactur- 
ing aluminum.  Here,  A1203  is  electrolyzedf  and 
the  net  reaction  in  the  cell  is 


A\203(l) 


2Al(l)  +  %02(g) 


(6~) 


Here  we  find  that  the  same  current  that  produced 
6.03  grams  of  mercury  produces  0.538  gram  of 
aluminum. 

Thus  after  the  same  amount  of  electricity  is 
passed  through  the  three  cells,  the  weights  of 
metals  produced  are  found  to  be 

6.03  g  Hg 

1.38  gNa 
0.538  g  Al 

How  are  these  weights  related?  Faraday  realized 


*  In  practice,  calcium  chloride  must  be  added  to  such 
a  cell  to  lower  the  melting  point  of  the  salt  mixture  and, 
even  then,  the  temperature  must  be  high  (600°C).  This  is 
the  commercial  method  for  manufacturing  metallic  so- 
dium. 

t  This  is  the  basis  for  the  commercial  manufacture  of 
aluminum.  Another  salt  is  added  as  solvent  to  lower  the 
melting  point.  A  mixture  of  AUOj  and  NaiAlFe  (cryolite) 
.  can  be  electrolyzed  at  950°C 


that  simple  numbers  result  in  such  a  case  if  each 
weight  is  divided  by  the  appropriate  atomic 
weight: 

T^tJ-  "  5Sr^T   =  00300  mole 
at  wt  Hg      201  g/mole 

wt  Na  1.38  g  nft,m       , 

.     .  VT    =  rr^ — t-S-t-  =  0.0600  mole 
at  wt  Na      23.0  g/mole 


wt  Al 


0.538  g 


at  wt  Al      26.9  g/mole 


=  0.0200  mole 


These  numbers  are  simply  related  to  each 
other  as  shown  by  the  ratios: 


Hg 

Na 

Al 

0.0300 

0.0600 

0.0200 

3 

6 

2 

Within  the  atomic  theory,  this  result  means 
that  a  certain  amount  of  electricity  will  deposit 
a  fixed  number  of  atoms,  or  some  simple  mul- 
tiple of  this  number,  whatever,  the  element.  Thus, 
in  both  of  Faraday's  experiments,  we  find  that 
an  atom  can  carry  only  a  fixed  quantity  of 
charge,  or  some  simple  multiple  of  this  quantity. 
Therefore,  electric  charge  comes  in  packages.  An 
atom  can  carry  one  package,  two  packages,  pos- 
sibly three  packages,  of  charge,  but  not  1.5872 
packages.  Whatever  the  package  of  charge  is,  it 
is  the  same  for  all  atoms.  The  realization  that 
electric  charge  comes  in  packages  led  to  the 
proposal  that  electricity  is  composed  of  particles. 
Since  atoms  carry  electric  charges,  atoms  must 
contain  these  particles. 


EXERCISE  14-2 


As  current  is  passed  through  the  cells  shown  in 
Figure  14-2,  the  oxygen  produced  in  the  first  cell 
is  collected  and  its  volume  is  compared  with  the 
volume  of  chlorine  produced  in  the  center  cell 
(the  volumes  being  compared  at  identical  tem- 
peratures and  pressures).  The  volume  of  chlorine 
is  found  to  be  exactly  double  that  of  oxygen. 
Applying  Avogadro's  Hypothesis,  explain  how 
this  result  shows  that  electricity  can  "count" 
atoms. 


sec.    14-2 


SEEING        PARTS    OF    ATOMS 


239 


14-2     "SEEING"  PARTS  OF  ATOMS 


Despite  the  convincing  support  for  the  atomic  theory 
provided  by  chemical  evidence,  there  is  intuitive  appeal 
to  evidence  that  is  closer  to  the  "direct  vision"  type.  From 
such  experiments  comes  a  much  more  detailed  view  of 
the  atom  and  its  make-up. 


14-2.1    "Seeing"  Electrons 

The  Faraday  experiments  were  the  original  basis  for  the 
suggestion  that  electricity  consists  of  individual  charges 
called  electrons.  Other  experiments  involving  the  passage 
of  electricity  through  gases  provide  further  evidence  that 
electrons  do  exist. 

Consider  the  apparatus  shown  in  Figure  14-3.  A  glass 
tube  is  fitted  with  electrodes  so  that  a  potential  difference 
of  10,000  volts  can  be  applied  across  a  space  filled  with  a 
desired  gas  at  various  pressures.  Suppose  neon,  for  ex- 
ample, is  placed  in  the  tube.  With  the  voltage  applied, 
the  gas  will  begin  to  conduct  electricity  when  its  pressure 
is  reduced  to  about  0.01  atmosphere.  The  tube  then  glows 
with  the  familiar  color  of  a  "neon  sign."  If  a  different 
gas  is  used,  the  color  is  different,  but  otherwise,  the  be- 
havior is  about  the  same.  If  the  pressure  is  reduced  still 
further  to  about  10-6  atmosphere,  the  glow  from  the  gas 


To    vacu. 
pump 


Fig.  14-3.  An  electric  discbarge  through  a  gas. 


Fig.  14-4.  An  electric  discharge  tube,  very  low  pres- 
sure. Electrons  travel  from  the  negative 
electrode  to  the  positive  electrode;  some  of 
them  pass  through  the  triangular  hole  to 
produce  a  triangular  spot  on  the  fluores- 
cent screen. 


voltage  can  be  applied  to  plates  A  and  P2  when  switch  S 
is  closed.  When  switch  Si  is  open,  the  fluorescence  appears 
at  position  A,  just  as  in  Figure  14-4.  When  Si  is  closed, 
however,  the  fluorescent  spot  moves  to  position  B.  Ap- 
parently the  fluorescent  spot  is  caused  by  particles  that 
are  attracted  to  the  positively  charged  electrode,  P2. 
Therefore  these  particles  must  be  negatively  charged. 
Light  cannot  be  deflected  in  this  way,  hence  the  fluo- 
rescent glow  cannot  be  caused  by  light.  Such  experiments 
with  discharge  tubes  show  that  negatively  charged  par- 
ticles exist.  These  particles  are  now  known  to  be  electrons. 
This  experiment  does  have  some  features  of  a  "direct 
vision"  observation.  First,  the  glowing  spot  is  directly 
visible.  Second,  it  is  easy  to  imagine  an  invisible  stream 
of  particles  hurtling  through  the  triangular  hole  in  the 
electrode  to  crash  against  the  fluorescent  screen  in  a  burst 


disappears  but  there  remains  a  fluorescent  glow  from  the 
glass  walls  of  the  tube.  The  apparatus  shown  in  Figure 
14-4  demonstrates  that  the  fluorescent  glow  is  either 
caused  by  particles  or  by  light  rays  that  travel  from  the 
negative  electrode  and  past  the  positive  electrode  (in  this 
apparatus,  through  the  triangular  hole  in  the  center  of 
the  electrode).  When  the  tube  operates,  a  glowing  area 
appears  on  the  glass  wall  at  position  A  directly  opposite 
the  hole  in  the  positive  electrode  and  just  the  same  shape 
as  the  hole.  Because  the  glowing  area  is  "shadowed"  by 
the  positive  electrode,  it  must  be  caused  by  rays  that  travel 
in  straight  lines. 

Figure  14-5  shows  this  same  apparatus  fitted  with  an 
auxiliary  pair  of  electrode  plates,  Pi  and  Pt.  An  electrical 


Fig.  14-5.  An  electric  discharge  tube  with  deflection 
electrodes. 


240 


WHY    WE    BELIEVE    IN    ATOMS    I    CHAP.    14 


of  light.  Third,  the  experiment  conveys  detailed  informa- 
tion about  these  particles,  information  difficult  to  obtain 
any  other  way.  The  electric  charge  on  a  particle  is  clearly 
evident  from  the  deflection  experiments.  Accurate  meas- 
urements of  such  deflections  even  lead  to  a  measure  of 
the  ratio  of  electron  charge  to  electron  mass. 

Yet  there  is  a  real  difference  between  this  experiment 
and  the  "direct  vision"  of  the  shadowy  form  of  the  gar- 
bage collector.  We  don't  see  the  electron  directly;  rather 
we  see  a  burst  of  light  on  the  fluorescent  screen.  The  light 
isn't  considered  to  be  the  electron:  the  burst  of  light  is 
caused  by  molecular  damage  to  the  screen— damage  re- 
sulting from  the  electron  crash.  A  more  apt  comparison 
from  our  analogy  would  be  our  seeing  some  footprints 
in  the  garden.  We  assume  the  footprints  are  caused  by 
the  garbage  collector.  Then  from  certain  properties  of  the 
footprints— size,  depth,  spacing— we  form  a  detailed 
image  of  his  height,  weight,  stride.  You  will  find  that  this 
is  typical  of  most  of  the  experiments  that  might  be  called 
"seeing"  atoms  and  their  components.  We  see  their  "foot- 
prints"—bursts  of  light  on  a  screen,  marks  on  a  photo- 
graphic plate,  discharges  in  a  Geiger  counter,  etc.  These 
"footprints"  substantiate  in  every  way  the  atomic  theory 
and  furnish  detailed  information  on  the  nature  of  atoms 

THE    RATIO    OF    ELECTRON    CHARGE 
TO    ELECTRON    MASS,    e/m 

By  the  action  of  a  magnetic  field  electrons  may  be  made 
to  follow  a  curved  path.  Such  experiments  lead  to  a  deter- 
mination of  the  ratio  of  the  electric  charge  of  an  electron 
e,  to  its  mass,  m. 

Consider  the  apparatus  shown  in  Figure  14-6.  The 
equipment  is  similar  to  that  shown  in  Figure  14-4  except 
a  fluorescent  screen  within  the  tube  reveals  the  trajectory 
of  the  particles  that  pass  through  the  slot  in  the  positive 
electrode.  When  a  magnetic  field  is  added,  the  electron 
trajectory  is  curved.  A  mathematical  analysis  of  the 
curvature  permits  an  interpretation  of  this  experiment 
that  leads  to  a  determination  of  e/m. 


Fig.  14-6.  The  effect  of  a  magnetic  field  on  the  elec- 
tron beam. 


To    vacuum 
put 


Fluorescent-  screen 


The  calculations  can  be  made  only  if  the  strength  of 
the  magnetic  field  is  known  and  if  the  field  is  uniform. 
Therefore,  apparatus  more  suitable  than  that  shown  in 
Figure  14-6  is  needed.  An  electric  current  flowing  through 
a  wire  coil  generates  a  magnetic  field  that  is  easily  meas- 
ured and  readily  made  uniform  (by  making  the  coil  large 
compared  to  the  apparatus). 

Substituting,  then,  a  large  coil  for  the  magnet  shown 
in  Figure  14-6,  we  can  proceed  with  our  measurement. 
The  beam  of  electrons  passes  through  the  positive  elec- 
trode and  strikes  the  far  end  of  the  tube,  producing  a 
fluorescent  spot.  When  the  magnetic  field  is  turned  on 
(by  passing  current  through  the  coil),  the  fluorescent  spot 
moves.  The  spot  moves  because  a  charged  particle  moving 
in  a  uniform  magnetic  field  has  a  path  which  is  an  arc  of  a 
perfect  circle.  From  the  deflection  of  the  spot  and  the 
length  of  the  apparatus,  the  radius  of  this  circular  path 
can  be  determined.  This  radius  we  shall  call  r. 

This  radius  is  useful  to  us  because  it  is  related  to  the 
mass,  m,  charge,  e,  and  velocity,  v,  of  the  electron.  It  is 
determined  also  by  the  strength  of  the  magnetic  field,  B, 
as  follows: 


e        B 


(7) 


This  equation  shows  that  the  greater  the  mass  or  velocity 
of  the  particle,  the  less  curved  is  its  path  (a  small  value 
of  r  describes  a  highly  curved  path).  On  the  other  hand, 
the  path  of  the  particle  becomes  more  curved  if  the  mag- 
netic field  is  made  stronger. 

We  can  rearrange  equation  (7)  to  the  form 


m       r       B 


(8) 


Magnet 


Equation  (8)  shows  us  how  to  calculate  the  charge/mass 
ratio  for  the  electron  if  r  is  measured  and  both  v  and  B 
are  known. 

However,  the  velocity,  v,  of  the  electron  is  still  un- 
known. We  must  calculate  this  quantity  from  the  work 
done  on  the  electron  as  it  was  accelerated,  moving  from 
the  negative  electrode  to  the  positive  electrode.  The  work 
done  on  the  electron  is  the  product  of  the  charge  on  the 
electron  times  the  voltage  difference,  V,  between  the 
electrodes: 

Work  done  on  electron  =  e  X  V  (9) 

This  work  is  used  to  accelerate  the  electron,  giving  it 
kinetic  energy.  Also,  we  know  that  the  kinetic  energy  of 
the  electron  can  be  expressed  in  terms  of  its  mass  and 
velocity : 

Kinetic  energy  of  a  moving  electron  =  \rn\~1    (10) 

We  must  equate  (9)  and  (10):  the  work  done,  eV,  equals 
the  kinetic  energy  the  electron  acquires,  \rn\-: 

eV=$m\*  (11) 

The  voltage,   V,  we  obtain  from  a  voltmeter  reading. 
Expressions  (8)  and  (//)  both  relate  the  ratio  e/m  to 


SEC.   14-2  I  "seeing"  parts  of  atoms 


241 


the  electron  velocity  and  the  measured  quantities,  r,  B, 
and  V.  We  can  calculate  e/m  if  v  is  eliminated  from  the 
two  equations.  This  is  an  algebraic  process  that  can  be 
done  several  ways.  Here  is  one  way. 
Since  (//)  involves  v2,  let  us  square  expression  (8): 


g2        v2 
m2  ~  r-B2 

Now  let  us  multiply  both  sides  of  (12)  by  m/e: 

e       my2 


X 


r-B- 


Now  we  can  rearrange  (//)  to  the  form 

e 
And,  finally,  we  can  substitute  (14)  into  (/J): 

m         e         r-B2  r2B2 


(12) 


(13) 


(14) 


2V 
r2B2 


(15) 


We  measure  V  (from  the  voltmeter  reading),  r  (from 
the  deflection  of  the  spot),  B  (from  the  current  through 
the  magnet  coil  windings),  substitute  them  into  (15),  and 
calculate 

coulombs 


-  =  1.759  X  10s 

m  gram 


THE    CHARGE    ON    THE    ELECTRON 


(16) 


Experiments  like  those  described  in  Figures  14-3  to  14-6 
establish  that  the  electron  is  a  negatively  charged  particle 
and  that  it  is  present  in  all  substances.  Further  confirma- 
tion of  the  particulate  nature  of  electricity  comes  from 
experiments  that  were  conducted  by  an  American  physi- 


Fig.  14-7.  Millikun's  oil-drop  apparatus  for  determin- 


ing the  electron  charge. 


Telescope 


Adjustable 

high 

voltage 


cist,  Robert  Millikan,  in  1906  to  determine  the  charge  on 
the  electron.  The  apparatus  used  for  his  experiment  is 
shown  schematically  in  Figure  14-7. 

Tiny  droplets  of  oil  or  some  other  liquid  are  sprayed 
into  the  upper  part  of  the  apparatus.  A  few  droplets 
fall  through  a  small  hole  into  the  lower  chamber.  During 
its  production,  an  oil  drop  is  very  likely  to  become 
charged  by  friction.  When  an  oil  drop  enters  the  lower 
chamber,  a  voltage  is  applied  to  the  metal  plates.  If  the 
oil  drop  is  charged,  its  fall  can  be  completely  stopped  by 
adjusting  the  voltage  so  that  the  electrical  force  on  the 
charged  drop  is  just  equal  and  opposite  to  the  force  of 
gravity. 

Millikan  made  thousands  of  determinations  of  the 
charge  on  drops  of  oil,  glycerol,  and  mercury.  The  charge 
on  the  drop  was  sometimes  positive  and  sometimes  nega- 
tive, but  in  every  case  its  magnitude  was  some  integral 
multiple  of  1.602  X  10~19  coulomb.*  In  no  case  was  the 
charge  any  less  than  this.  These  experiments  give  a  clear 
demonstration  that  the  fundamental  unit  of  electricity 
must  be  a  charge  of  1.602  X  10~19  coulomb.  If  the  elec- 
tron carries  this  fundamental  unit  of  electricity  (as  we 
believe  it  does),  the  value  of  the  charge  on  the  electron 
must  be  1.602  X  10~19  coulomb: 


e  =  1.602  X  10-'9  coulomb 


(17) 


We  may  use  this  value  of  the  charge  on  the  electron 
to  calculate  the  mass  of  an  electron.  To  do  so,  it  is  neces- 
sary to  know  the  ratio  of  (electron  charge/electron 
mass)  =  e/m.  This  ratio  is  measured  with  apparatus 
based  on  principles  displayed  in  Figures  14-4  and  14-6. 
Using  the  result  e/m  =  1.759  X  10s  coulombs/g,  the 
mass  of  an  electron  is  found  to  be 


=  1.602  X  IP"19  coulomb  /electron 
1.759  X  lO^coulomb/g 

m  =  9.11  X  10-28  g/electron 


(/*) 


EXERCISE  14-3 

Suppose  five  measurements  of  oil-drop  charges  give  the 
values  listed  below: 

4.83  X  10"19  coulomb 
3.24  X  10-'9 
9.62  X  10"19 
6.44  X'10-» 
4.80  X  1019 


*  The  coulomb  is  a  unit  of  electric  charge.  Its  magni- 
tude can  be  appraised  by  its  relation  to  the  ampere.  One 
ampere  is  an  electric  current  of  one  coulomb  of  charge 
passing  a  point  in  a  wire  every  second.  One  mole  of  elec- 
trons has,  then,  96,500  coulombs  of  charge.  In  a  wire 
carrying  10  amperes,  it  takes  about  two  and  one  half 
hours  for  one  mole  of  electrons  to  pass  any  point. 


242 


WHY    WE    BELIEVE    IN    ATOMS    |    CHAP.     14 


(a)  Divide  each  charge  by  the  smallest  value  to  investi- 
gate the  relative  magnitudes  of  these  charges. 

(b)  Assuming  each  measurement  has  an  uncertainty  of 
±0.04  X  10-19,  decide  what  electron  charge  is  indi- 
cated by  these  experiments  alone. 


14-2.2    "Seeing"  Positive  Ions 

Experiments  can  be  conducted  in  which  positive  ions  are 
detected  and  their  properties  measured  (charge  and  mass). 
These  experiments  are  similar  to  those  we  have  described 
for  electrons.  A  gas  discharge  tube,  such  as  was  shown 
in  Figure  14-3,  can  be  used  because  measurements  show 
that  positive  ions  are  present  as  well  as  electrons.  Whereas 
electrons  are  accelerated  toward  the  positive  electrode, 
the  positively  charged  ions  are  accelerated  in  the  opposite 
direction,  toward  the  negative  electrode.  These  ions  can 
be  removed  from  the  apparatus  as  a  beam  in  the  same 
way  that  the  electron  beam  is  removed  in  the  apparatus 
of  Figure  14-4.  In  such  fashion,  we  obtain  a  beam  of  posi- 
tive ions.  By  deflecting  these  beams  in  electric  and  mag- 
netic fields,  the  charges  and  masses  of  the  positive  ions 
can  be  measured. 

The  results  of  experiments  of  this  type  show  two  very 
important  differences  from  measurements  on  electrons. 

(1)  The  charge/mass  ratio  for  positive  ions  changes  when 
the  gas  in  the  tube  is  changed.  When  the  (e/m)  meas- 


urement is  made  for  electrons,  the  same  value  is 
obtained  no  matter  what  gas  is  introduced. 
(2)  The  charge/mass  ratio  for  positive  ions  is  very  much 
smaller  than  (elm)  for  electrons.  These  facts  are  in- 
terpreted to  mean  that  the  positive  ions  are  ions 
formed  from  the  gas  in  the  tube.  The  electric  charge 
is  considered  to  arise  from  the  removal  of  one  or  more 
electrons  from  an  atom  or  a  molecule.  Thus  the  value 
of  the  ratio  (charge/mass)  for  positive  ions  depends 
upon  the  gas  because  each  type  of  atom  (or  molecule) 
has  a  distinctive  mass. 

"weighing"  positive  ions,  the  mass 
spectrograph 

A  mass  spectrograph  is  an  instrument  with  which  the 
masses  of  individual  atomic  or  molecular  ions  can  be 
measured.  One  type  of  mass  spectrograph  is  shown  in 
Figure  14-8.  Positive  ions  are  accelerated  through  a 
slotted  negative  electrode  and  then  passed  through  a 
uniform  magnetic  field.  The  left  view  in  Figure  14-8  shows 
the  apparatus  supported  between  the  pole  faces  of  a 
strong  magnet.  The  right  view  is  an  enlargement  of  the 
spectrograph  with  the  magnetic  field  directed  vertically 
through  the  figure.  This  is  the  view  of  the  mass  spectro- 


Fig.  14-8.  A  mass  spectrograph  and  the  mass  spec- 
trum of  neon. 


Pole  face 


Pole  face 

To  vacuum  pump 

Vacuu.m   charrCber 


To  vacuum. 


A 


Magnet 


Electron    beam. 


Photographic  plate 
rC__r*"  To  vacf.um.  pump 


Neon  gas  en-try 


Plate  III.  A  simple  spectrograph  and  the  spectrum  of  a  hot  tungsten  ribbon. 


^_ 


I 

0) 

e 


8 

M 


H 


0\ 


»s»\ 


0\  ■ 


CO 


CO 


»0 

^0 


v. 


^> 


8- 


h 
> 


q 
* 
n 


^ 


N 


^ 


SEC.    14-2 


SEEING       PARTS    OF    ATOMS 


243 


graph  seen  by  an  ant  sitting  on  one  pole  face  of  the  mag- 
net looking  toward  the  other  pole  face. 

The  positive  ions  can  be  produced  with  a  glow  dis- 
charge tube  like  that  shown  in  Figure  14-3.  More  usually, 
however,  gaseous  atoms  or  molecules  are  bombarded 
with  an  electron  beam  as  shown  in  Figure  14-8.  If  the 
bombarding  electrons  have  enough  energy,  they  cause 
positive  ion  formation  when  collisions  occur  with  gas 
molecules.  The  figure  shows  neon  gas  entering  at  the 
bottom.  The  gas  passes  through  the  electron  beam  and 
some  of  the  atoms  collide  with  electrons  to  form  neon 
ions.  Both  Ne+  and  Ne+2  ions  are  formed  and  they  are 
accelerated  by  the  slotted  electrode.  As  the  positive  ions 
enter  the  magnetic  field,  they  follow  a  circular  path.  They 
have  a  large  radius  of  curvature  if  the  mass  is  high,  a  low 
radius  of  curvature  if  the  charge  is  high.  Thus  each  posi- 
tive ion  follows  a  circular  path  fixed  by  its  mass  and 
charge.  After  circling  through  an  arc  of  180°,  the  ions  are 
collected  on  a  photographic  plate.  The  impact  of  the  ions 
with  the  photographic  plate  causes  a  reaction  that  leads 
to  a  darkening  of  the  sensitized  surface,  just  as  exposure 
to  light  does.  Such  a  record  shows  a  spot  for  each  ion  at 
a  position  fixed  by  the  charge/mass  ratio.  Measurement 
of  the  position  of  each  spot  reveals  the  masses  of  the 
ions.  The  record  is  called  a  mass  spectrum. 

When  neon  gas  is  put  in  the  spectrograph,  the  mass 
spectrum  consists  of  two  widely  separated  groups  of  three 
spots  each.  The  three  spots  corresponding  to  large  radii 
are  caused  by  neon  ions  with  a  single  positive  charge, 
while  the  three  spots  corresponding  to  small  radii  are 
caused  by  doubly  charged  ions.  For  each  ionic  charge 
there  are  three  slightly  separated  spots  which  indicate  that 
neon  consists  of  atoms  with  three  different  masses. 
This  shows  that  ordinary  neon  consists  of  three  different 
isotopes.  The  relative  abundances  of  these  isotopes  can 
be  determined  by  measuring  the  intensity  of  the  spots 
caused  by  each  of  the  ion  beams. 


EXERCISE  14-4 

When  chlorine,  Cls,  is  examined  in  a  mass  spectrograph, 
CI2",  Cl+,  and  Cl+2  ions  are  formed.  Remembering  that 
there  are  two  isotopes  in  chlorine,  35  (75%)  and  37 
(25%),  describe  qualitatively  the  appearance  of  the  mass 
spectrum.  Which  ion  will  produce  lines  at  the  largest 
radius?  Which  at  the  smallest  radius?  How  many  lines 
will  each  ion  produce? 


Figure  14-8  shows  a  mass  spectrograph  that  uses  photo- 
graphic detection.  Nonphotographic  detection  is  also 
possible.  The  ions,  after  being  sorted  according  to  mass 
and  charge,  can  be  "counted"  by  a  charge  measuring 
device.  The  advantage  of  such  a  detector  is  that  tbe  result 
can  be  presented  continuously  on  a  paper  chart,  thus 
eliminating  the  cumbersome  and  slow  photographic 
process. 


THE    RATIO    OF    CHARGE    TO    POSITIVE 

ion  mass,  e/m 

In  a  mass  spectrograph,  the  factors  that  determine  the 
trajectory  of  the  ions  are  the  same  as  those  we  discussed 
when  we  considered  the  measurement  of  (film)  for  the 
electron.  In  that  discussion  we  derived  equation  (79): 
e       2V 

m  =  7*  (79) 

where  e   =  electron  charge, 

m  =  electron  mass, 

V  =  accelerating  voltage, 

B  =  magnetic  field  strength. 

For  positive  ions,  the  charge  is  e  or  2e  or,  in  general, 
some  integer,  n,  times  e,  the  electron  charge.  We  might 
write  a  capital  M  for  mass  to  indicate  that  the  mass  of  a 
positive  ion  is  involved.  Then  we  can  solve  (19)  for  the 
mass  as  a  function  of  V,  B  (experimental  conditions  we 
control),  n  (which  will  be  one,  two,  three,  or  some  low 
integer),  and  r,  the  radius  that  we  measure  on  the  photo- 
graphic plate: 

neriBi 


M  = 


2V 


(20) 


With  equation  (20),  we  can  verify  quantitatively  our 
identifications  of  the  ionic  masses  and  their  charges.  The 
two  sets  of  three  spots  immediately  suggest  that  one  set 
is  caused  by  three  isotopes,  each  with  the  same  ionic 
charge,  and  the  second  set  by  the  same  three  isotopes  with 
a  different  ionic  charge.  Measurement  of  the  radii  shows 
that  one  set  has  a  larger  radius  by  just  the  square  root  of 
two.  This  ratio  is  consistent  with  the  assignment  of  +1 
to  one  set  of  ions  and  +2  to  the  other  set.  Calculations 
based  on  the  assumption  that  n  =  1  for  the  outer  set, 
together  with  the  accurately  measured  radii  of  the 
paths  corresponding  to  the  three  spots  on  the  plate 
give  isotopic  masses  corresponding  to  20,  21,  and  22 
grams/mole,  the  three  stable  isotopes  of  neon.  If  the  true 
value  of  n  for  this  outer  set  of  spots  had  been  n  =  2,  the 
calculated  masses  would  correspond  to  isotopic  masses  of 
40,  42,  and  44  grams/mole.  The  choice  between  the  two 
assumptions  is  usually  easily  made  on  the  basis  of  chemi- 
cal arguments  about  the  possible  atoms  or  molecules 
present. 

EXERCISE  14-5 

Suppose  a  mass  spectrograph  is  used  to  measure  the 
charge/mass  ratio  for  fluorine  ions.  Fluorine  has  only  one 
stable  isotope  and  its  atomic  weight  is  19.0  grams/mole. 
From  the  measured  charge/mass  ratio,  5.08  X  10s  cou- 
lombs/gram, and  the  assumption  that  the  ion  has  one 
electron  charge,  calculate  the  mass  of  one  ion.  Repeat  the 
calculation  assuming  the  ion  has  two  electron  charges. 
Now  calculate  Avogadro's  number  from  the  weight  of  a 
mole  of  fluorine  ions,  using  each  of  your  two  calculations. 
Which  assumption  about  ion  charge  do  you  prefer? 
Could  the  other  be  correct  as  well? 


244 


WHY    WE    BELIEVE    IN    ATOMS    I    CHAP.     14 


Lead  block 


Fig.  14-9.  Rutherford's  apparatus  for  observing  the 
scattering  of  alpha  particles  by  a  metal  foil. 
{The    entire    apparatus    is    enclosed    in    a 

vacuum  chamber.) 


14-2.3     "Seeing"  the  Nucleus:  Structure 
off  the  Atom 

So  far  we  have  described  experiments  that  indicate  that 
atoms  exist  and  that  indicate  atoms  are  composed  of 
charged  particles.  We  know  also  that  all  the  positively 
charged  part  of  the  atom  is  located  in  a  very  small  but 
dense  region  which  we  call  the  nucleus.  The  negatively 
charged  electrons  spend  most  of  their  time  at  relatively 
great  distances  from  the  nucleus.  The  story  of  how  this 
nuclear  model  of  the  atom  was  first  proposed  gives  a 
fascinating  view  of  how  science  progresses. 

The  first  detailed  model  of  the  atom,  proposed  by 
J.  J.  Thomson  in  1898,  was  based  upon  the  expectation 
that  the  atom  was  a  sphere  of  positive  electricity  in  which 
electrons  were  embedded  like  plums  in  a  pudding.  This 
picture  of  the  atom  was  not  particularly  satisfying  be- 
cause it  was  not  useful  in  predicting  or  explaining  the 
chemical  properties  of  the  atom.  Finally,  in  1911,  a  series 
of  experiments  performed  in  the  McGill  University  labo- 
ratory of  Ernest  Rutherford  showed  that  Thomson's 
picture  of  the  atom  had  to  be  abandoned. 

The  experiment  conducted  by  Rutherford  and  his  co- 
workers involved  bombarding  gold  foil  with  alpha  par- 
ticles, which  are  doubly  charged  helium  atoms.  The 
apparatus  used  in  their  experiment  is  shown  in  Figure 
14-9.  The  alpha  particles  are  produced  by  the  radioactive 
decay  of  radium,  and  a  narrow  beam  of  these  particles 
emerges  from  a  deep  hole  in  a  block  of  lead.  The  beam 
of  particles  is  directed  at  a  thin  metal  foil,  approximately 
10,000  atoms  thick.  The  alpha  particles  are  detected  by 
the  light  they  produce  when  they  collide  with  scintilltaion 
screens,  which  are  zinc  sulfide-covered  plates  much  like 
the  front  of  the  picture  tube  in  a  television  set.  The  screen 


Unde-viat-ed  alpha  particle  beam 


is  mounted  on  an  arm  in  such  a  way  that  it  can  be  moved 
around  in  a  circle  whose  center  is  in  line  with  the  point 
where  the  alpha  particles  strike  the  foil.  A  telescope  is 
mounted  behind  the  screen  so  that  the  very  small  flashes 
of  light  produced  when  individual  alpha  particles  strike 
the  scintillation  screen  can  be  detected  and  counted.  The 
apparatus  operates  in  a  vacuum  chamber  in  order  that 
no  deflections  are  caused  by  the  impact  of  alpha  particles 
upon  gaseous  molecules. 

The  first  observation  made  with  this  apparatus  was  that 
apparently  all  the  alpha  particles  passed  through  the  foil 
undeflected.  Let  us  see  if  this  result  is  consistent  with  the 
model  of  the  atom  proposed  by  Thomson.  You  will  recall 
that  Thomson's  picture  of  the  atom  assumed  that  the 
positive  charge  is  distributed  evenly  throughout  the  entire 
volume  of  the  atom  with  the  negative  electrons  embedded 
in  it.  Since  the  electrons  weigh  so  little,  the  positive  part 
accounts  for  nearly  all  of  the  mass  of  the  atom.  Thus  the 
Thomson  model  pictures  the  atom  as  a  body  of  uniform 
density. 

Imagine  what  our  thin  metal  foil  would  be  like  if  it 
were  to  be  made  up  of  Thomson  atoms.  The  physical 
properties  of  a  solid  suggest  that  the  atoms  lie  very  close 
together,  so  the  metal  foil  would  look  something  like  the 
diagram  shown  in  Figure  14-10.  Of  course,  the  real  foil 
is  10,000  atoms  thick.  What  would  happen  to  the  alpha 
particles  if  they  were  shot  into  a  solid  of  such  uniform 
density?  At  first  we  might  think  that  they  would  be 
stopped  or  deflected  back  upon  colliding  with  the  atoms. 
Since  it  was  observed  that  the  alpha  particles  went  straight 
through  the  metal  foil,  we  must  reconsider  the  problem. 
When  we  shoot  at  a  paper  target  with  a  high-powered 
rifle,  the  projectile  forces  its  way  through  the  paper.  The 
alpha  particles  produced  by  radium  have  very  high  kinetic 
energy  and  are  very  much  like  bullets  from  a  high- 
powered  rifle.  Perhaps  the  very  high  kinetic  energy  allows 
an  alpha  particle  to  force  its  way  right  through  the  atoms 
of  the  metal  foil.  Since  a  rifle  bullet  fired  into  paper  passes 
through  undeflected,  it  seems  reasonable  to  conclude  that 
the  alpha  particle  would  also  pass  through  the  metal  foil 
undeflected. 


SEC.     14-3    I    MEASURING    DIMENSIONS    OF    ATOMS    AND    MOLECULES 


245 


Alpha  particle 


Fig.  14-10.  The   scattering   of   alpha   particles   by   a 

metal  foil  made  of  Thomson  atoms. 


Atomic  nucleus 


In  summary,  in  the  Thomson  model  a  metal  foil  is 
considered  to  have  essentially  uniform  density.  If  this  is 
true,  there  is  no  way  for  bombarding  alpha  particles  to 
be  deflected  through  large  angles.  At  best,  the  alpha  par- 
ticles might  suffer  slight  deflections  from  many  collisions 
with  many  atoms.  The  model  predicts  the  scattering  dis- 
tribution shown  in  Figure  14-10. 

The  first  results  of  Rutherford's  experiments  seemed 
to  be  quite  consistent  with  the  Thomson  picture  of  the 
atom.  On  more  careful  examination,  an  astounding  dis- 
covery was  made.  By  moving  the  screen  around  the  metal 
foil,  Rutherford  and  his  co-workers  were  able  to  observe 
that  a  very  few  scintillations  occurred  at  many  different 
angles;  some  of  these  angles  were  nearly  as  large  as  180°. 
It  was  as  if  some  of  the  alpha  particles  had  rebounded 
from  a  head-on  collision  with  an  immovable  object.  In 
the  words  of  Rutherford,  "It  is  about  as  incredible  as  if 
you  had  fired  a  15  inch  shell  at  a  piece  of  tissue  paper 
and  it  came  back  and  hit  you."  It  was  impossible  to  ex- 
plain the  simultaneous  observation  of  large-angle  and 
small-angle  deflections  by  using  the  Thomson  atom. 

In  order  to  explain  his  experimental  results,  Rutherford 
designed  a  new  picture  of  the  atom.  He  proposed  that  the 
atom  occupies  a  spherical  volume  approximately  10~8  cm 
in  radius  and  at  the  center  of  each  atom  there  is  a  nucleus 
whose  radius  is  about  10~12  cm.  He  further  proposed  that 
this  nucleus  contains  most  of  the  mass  of  the  atom,  and 
that  it  also  has  a  positive  charge  that  is  some  multiple 
of  the  charge  on  the  electron.  The  region  of  space  outside 
the  nucleus  must  be  occupied  by  the  electrons.  We  see 
from  Figure  14-1 1  that  Rutherford's  picture  requires  that 
most  of  the  volume  of  the  atom  be  a  region  of  very  low 
density. 

Using  this  kind  of  model  of  the  atom,  we  can  account 
for  the  alpha  particles  that  are  deflected  through  both 


Fig.   14-11.  The  scattering  of  alpha  particles  by  a  foil 
made  of  Rutherford  nuclear  atoms. 


large  and  small  angles.  If  we  allow  alpha  particles  to 
impinge  upon  a  metal  foil  composed  of  atoms  based  on 
Rutherford's  model,  only  a  few  of  tne  particles  would  be 
appreciably  deflected  by  the  foil.  The  heavy,  fast  moving 
alpha  particles  can  brush  past  the  lighter  electrons  with- 
out being  deflected.  Since  most  of  the  volume  of  the  metal 
foil  is  relatively  empty  space,  the  greatest  number  of 
alpha  particles  pass  through  the  metal  undeflected.  It  is 
possible,  however,  foi  a  few  particles  to  be  scattered 
through  very  large  angles.  Since  both  the  alpha  particle 
and  the  nucleus  of  the  atom  are  positively  charged,  they 
exert  a  force  of  repulsion  on  each  other.  This  force  be- 
comes large  only  when  the  alpha  particle  comes  quite 
close  to  the  nucleus.  Since  the  nucleus  in  question  is  much 
heavier  than  the  alpha  particle,  it  can  deflect  the  alpha 
particle  considerably,  just  as  a  steel  post  can  deflect  a 
rifle  bullet. 

Besides  providing  a  qualitative  picture  of  the  atom, 
Rutherford's  experiments  provided  a  way  of  measuring 
the  charge  of  the  nucleus.  The  force  that  a  nucleus  exerts 
on  an  alpha  particle  depends  upon  the  magnitude  of  the 
charge  on  the  nucleus.  Rutherford  showed  how  to  relate 
the  number  of  alpha  particles  scattered  at  any  angle  to 
the  magnitude  of  the  charge  on  the  nucleus.  The  first 
measurements  of  the  nuclear  charge  by  this  method  were 
not  very  accurate,  but  they  did  show  that  different  ele- 
ments have  different  nuclear  charges.  By  1920,  however, 
the  alpha  particle  scattering  experiments  were  so  refined 
that  they  could  be  used  to  determine  nuclear  charge 
accurately. 


14-3     MEASURING  DIMENSIONS  OF  ATOMS  AND  MOLECULES 


There  are  several  ways  by  which  sizes  of  atoms 
in  molecules  and  in  solids  can  be  estimated. 
These  methods  are  classified  as  "spectroscopic" 


methods  because  they  involve  the  interaction  of 
light  with  matter.  The  measurements  show 
atomic  size  in  the  sense  that  they  show  how 


246 


WHY    WE    BELIEVE    IN    ATOMS   I    CHAP.     14 


closely  the  atoms  pack  together.  These  packing 
distances,  as  measured  spectroscopically,  have 
provided  the  dimensions  for  the  atomic  models 
you  have  seen. 

14-3.1     Light  and  the  Frequency  Spectrum 

Light  can  be  characterized  by  its  frequency  or 
its  wavelength.  To  understand  the  meaning  of 
these  terms,  consider  water  waves  approaching 
and  breaking  on  a  beach.  Figure  14-12  shows 
two  measurements  we  might  make,  the  distance 
between  crests  and  the  time  between  waves.  The 
distance  between  crests  is  called  wavelength  and 
it  can  be  expressed  in  centimeters.  The  time  be- 
tween waves,  r,  indicates  how  often  waves  pass 
a  fixed  point.   Usually  the  reciprocal,   1/r,  is 


Fig.  14-12.  Waves  can  be  characterized  by  wavelength 
or  by  time  between  waves. 


specified.  This  quantity,  with  the  dimension 
waves  per  second,  is  called  frequency  and  is  sym- 
bolized v  ("nu").  Light,  which  is  an  electromag- 
netic disturbance  traveling  through  space,  has 
properties  much  like  water  waves.  The  electro- 
magnetic disturbance  varies  periodically,  as  does 
the  water  disturbance,  hence  it  can  be  charac- 
terized by  its  frequency.  Also,  light  travels 
through  space  with  definite  distance  between  the 
"wave  crests"  where  the  electromagnetic  dis- 
turbance is  greatest.  This  distance  is  called  the 
wavelength. 

Figure  14-13  shows  a  spectrograph — an  in- 
strument that  reveals  the  frequency  composition 
of  light.  The  light  entering  a  narrow  slit  is 
focused  into  a  beam  by  the  lens.  This  beam  is 
passed  through  the  prism.  All  of  the  light  is 
refracted  (bent)  by  the  angular  prism,  but  dif- 
ferent frequencies  (colors)  are  bent  through 
different  angles.  The  result  is  that  the  frequency 


Time 
betrtveen 
waves 


SEC.     14-3    |    MEASURING    DIMENSIONS    OF    ATOMS    AND    MOLECULES 


247 


Blue 
Violet- 
Ultraviolet 


Hot 

tungsten 

ribbon 


Fig.  14-13.  A  simple  spectrograph. 


composition  of  the  light  entering  the  slit  can  be 
learned  from  the  pattern  focused  on  a  photo- 
graphic film.  The  light  source  is  a  tungsten  rib- 
bon heated  to  a  temperature  near  1000°C  by  an 
electric  current. 

This  separation  of  light  into  its  component 
frequencies  produces  a  spectrum.  This  spectrum 
is  recorded  on  the  photographic  film  because  the 
darkening  of  the  film  (on  development)  is  deter- 
mined by  the  light  intensity.  From  the  spectrum 
we  learn  that  different  colors  correspond  to  dif- 
ferent frequencies.  Blue  light  is  found  to  have  a 
frequency  of  about  7.5  X  10u  waves  per  second 


Ph  o  togrraph  ic  plate 


or,  in  more  usual  terminology,  v  =  7.5  X  1014 
cycles  per  second.  The  red  light  has  a  lower 
frequency;  v  is  about  4.3  X  10u  cycles  per  sec- 
ond. 

The  experiment  shown  in  Figure  14-13  points 
up  another  extremely  important  fact.  The  photo- 
graphic film  is  darkened  at  larger  angles  than 
that  at  which  blue  light  appears  and  at  smaller 
angles  than  that  at  which  red  light  appears.  This 
implies  that  the  light  emitted  by  the  hot  ribbon 
includes  frequencies  that  are  not  detected  by  the 
human  eye.  The  frequencies  lower  than  the  fre- 
quency of  red  light  are  called  infrared  frequen- 

Fig.  14-14.  The  complete  light  spectrum. 


G-arnma    rays 
X-rays 

Ultraviolet 
Visible 
Infra  red 


Mic  ro  wa  ve 
Radio 


Frequency.         '22 
cycles/second 


W1 


w 


JO' 


10 " 


w2 


Molecular  rotation 
b/folecular   vibration 
Electron    excitation 


248 


WHY    WE    BELIEVE    IN    ATOMS    i    CHAP.     14 


cies.  The  frequencies  higher  than  the  frequency  of 
violet  light  are  called  ultraviolet  frequencies. 

Scientists  have  now  realized  that  the  electro- 
magnetic phenomenon  called  light  extends  over 
an  enormous  range  of  frequencies — much  wider 
than  the  rather  narrow  region  in  which  the 
human  eye  is  sensitive.  Figure  14-14  shows  the 
range  that  is  commonly  studied  and  the  familiar 
names  given  to  various  spectral  regions. 

There  are  three  spectral  regions  or  ranges  of 
light  frequencies  that  are  particularly  useful  to 
chemists  in  learning  atomic  sizes.  We  shall  dis- 
cuss each  briefly. 


14-3.2     X-Ray  Diffraction  Patterns 

X-Rays  are  light  waves  of  frequencies  near  1018 
cycles  per  second  and  wavelengths  near  10~8  cm. 
Such  light  waves,  when  reflected  from  the  surface 
of  a  crystal,  give  patterns  on  a  photographic 
film.  The  pattern  is  fixed  by  the  spacings  of  the 
atoms  of  the  crystal  and  their  spatial  arrange- 
ment. The  pattern  is  obtained  only  with  X-rays 
because  it  results  from  scattering  effects  that 
occur  only  if  the  wavelength  of  the  light  is  close 
to  the  atomic  separations  within  the  crystal. 
Therefore,  a  knowledge  of  the  wavelength  of  the 
X-ray  light  permits  an  interpretation  of  the  pat- 
tern in  terms  of  atomic  packing. 

Figure  14-15  shows  three  X-ray  diffraction 
patterns  obtained  from  small  crystalline  particles 
of  metallic  copper,  aluminum,  and  sodium.  The 
qualitative  similarity  of  the  patterns  given  by 
copper  and  aluminum  shows  that  they  have  the 
same  crystal  packing.  Careful  measurements  of 
the  spacing  of  the  lines  indicate  that  the  atoms 


Fig.  14-15.  X-Ray  diffraction  patterns  of  finely  di- 
vided metallic  copper,  aluminum,  and 
sodium. 


CtM. 

At 


in  copper,  though  occupying  the  same  relative 
positions  as  the  atoms  in  aluminum,  are  closer 
together.  In  contrast,  the  pattern  of  lines  pro- 
duced by  sodium  does  not  resemble  either  of  the 
preceding  patterns.  Sodium  atoms  are  packed 
differently  in  the  metallic  sodium  crystal. 

The  X-ray  diffraction  method  is  applicable  to 
solids  and  provides  such  detailed  views  of  crystal 
geometry  as  those  shown  for  sodium  chloride 
solid  in  Figure  5-10,  p.  81. 


14-3.3     Microwave  Spectroscopy: 
Molecular  Rotation 

Radio  waves  are  light  rays  of  macroscopic  wavelengths 
(that  is,  wavelengths  of  many  meters).  Using  techniques 
similar  to  those  used  in  generating  radio  waves,  "micro- 
wave" light  can  be  produced  with  wavelengths  in  the 
range  1  mm  to  10  cm.  In  this  spectral  region,  gaseous 
molecules  absorb  light  because  of  excitation  of  rotational 
movements  of  the  molecules.  The  frequencies  of  these 
rotational  motions,  shown  in  Figure  7-8  (p.  118),  depend 
upon  the  distance  of  the  atoms  from  the  molecular  center 
of  gravity. 

Microwave  spectroscopy  is  applicable  only  to  gases  but 
it  is  capable  of  extremely  high  accuracy.  Interatomic  dis- 
tances and  structures  have  been  so  measured  for  many 
molecules  containing  only  a  few  atoms. 


14-3.4     Infrared  Spectroscopy: 
Molecular  Vibration 

Light  found  in  the  spectrum  just  beyond  the  red 
end  of  the  visible  spectrum  is  called  infrared 
light.  The  frequencies  are  in  the  range  2  X  1013 
to  12  X  1013  cycles  per  second  (with  wavelengths 
between  1.5  X  10-3  to  2.5  X  lCM  cm).  Mole- 
cules absorb  light  in  this  spectral  region,  and 
analysis  of  the  frequencies  shows  that  the  ab- 
sorptions are  associated  with  the  excitation  of 
vibrational  motions.  These  to-and-fro  motions  of 
the  atoms  occur  at  natural  frequencies  just  like 
the  natural  vibrational  frequencies  of  a  ball-and- 
spring  model  of  a  molecule.  These  natural  fre- 
quencies are  fixed  by  the  masses  of  the  atoms, 
the  molecular  shape,  and  the  strengths  of  the 
chemical  bonds  that  link  the  atoms  together. 
Again  the  frequencies  absorbed  by  gaseous  mole- 
cules provide  information  about  the  molecular 


SEC.    14-3    I    MEASURING    DIMENSIONS    OF    ATOMS    AND    MOLECULES 


249 


moments  of  inertia,  the  molecular  geometry,  and 
the  chemical  bonds.  In  addition,  infrared  study 
can  be  extended  easily  to  the  liquid  and  solid 
state,  hence  it  finds  widespread  use  in  chemistry. 
Figure  14-16  contrasts  the  infrared  absorption 
spectra  of  hydrogen  bromide  gas,  HBr,  and 
deuterium  bromide  gas,  DBr.  The  horizontal 
scale  shows  frequency.  For  a  given  frequency, 
the  vertical  scale  shows  the  percentage  of  light 
of  that  frequency  transmitted  by  the  sample. 
Thus  a  reading  of  100%  means  all  of  the  light  is 
transmitted;  hence,  no  light  is  absorbed  at  that 
frequency.  Plainly,  gaseous  HBr  absorbs  in  only 
one  spectral  region,  that  near  7.9  X  1013  cycles 
per  second.  This  one  absorption  corresponds  to 
the  vibrational  excitation  of  the  chemical  bond 
in  HBr.  There  is  only  one  bond,  hence  only  one 
absorption.  The  spectrum  of  gaseous  DBr  is 
similar  but  the  absorption  occurs  near  5.9  X  1013 
cycles  per  second.  Of  course,  the  chemistries  of 
DBr  and  HBr  are  identical,  hence  the  chemical 
bond  in  DBr  is  identical  to  that  in  HBr;  never- 


Fig.  14-16.  Infrared    absorption    spectra    of   gaseous 
HBr  and  DBr. 


theless,  the  vibration  frequencies  of  these  two 
molecules  differ  because  the  atom  masses  differ. 
Since  the  deuterium  atom  is  heavier  than  the 
hydrogen  atom,  it  vibrates  more  slowly. 

More  complicated  molecules,  with  two  or 
more  chemical  bonds,  have  more  complicated 
absorption  spectra.  However,  each  molecule  has 
such  a  characteristic  spectrum  that  the  spectrum 
can  be  used  to  detect  the  presence  of  that  par- 
ticular molecular  substance.  Figure  14-17,  for 
example,  shows  the  absorptions  shown  by  liquid 
carbon  tetrachloride,  CC14,  and  by  liquid  carbon 
disulfide,  CS2.  The  bottom  spectrum  is  that  dis- 
played by  liquid  CC14  containing  a  small  amount 
of  CS2.  The  absorptions  of  CS2  are  evident  in  the 
spectrum  of  the  mixture,  so  the  infrared  spec- 
trum can  be  used  to  detect  the  impurity  and  to 
measure  its  concentration. 

The  value  of  infrared  spectra  for  identifying 
substances,  for  verifying  purity,  and  for  quan- 
titative analysis  rivals  their  usefulness  in  learning 
molecular  structure.  The  infrared  spectrum  is  as 
important  as  the  melting  point  for  characterizing 
a  pure  substance.  Thus  infrared  spectroscopy 
has  become  an  important  addition  to  the  many 
techniques  used  by  the  chemist. 


I 

I 

K 

I 


100% 

50% 
0% 

- 

HBr  (9) 

........  i             i             i 

lOxlo13 

8xW13 

6xl013 

4xl013 

2xl013 

loo%r 


0% 


50%-       DBr    ($) 


Wxio13 


8xl015 


6xW13 


4  *1013 


2  xjo 


u 


Frequency  ,     cycles/second 


250 


WHY    WE    BELIEVE    IN    ATOMS   I    CHAP.    14 


100% 
50% 

:       cs2  (i) 

o% 

WxlO13 

8  xW13 

6 xlO13 

4xl013 

2X10 

13 

s 

0 

.5 

M 

I 


100%, 
50% 

:       cci+  (i) 

0% 

lOxlO13 

8xl013 

GxlO13 

4xW13 

2  xlO' 

'3 

ioo% 

V  \f                "N 

\/~^\             / 

50% 

-           CCI4  +  5%  CS2  (I) 

\) 

0% 

cs2 

t 

cs2 

lOxlO13              8  xlO13 

6  xlO13 

-4  xlO13 

2  xlO13 

Frequency ,     cycles /second 

Fig.  14-17.  Infrared  absorption  spectra  of  liquid  carbon    tetrachloride,    CCU,    carbon    disulfide,    CS«,   and   a 
mixture  of  the  two. 


QUESTIONS  AND  PROBLEMS 

1.  A  compound  of  carbon  and  hydrogen  is  known 
that  contains  1.0  gram  of  hydrogen  for  every  3.0 
grams  of  carbon.  What  is  the  atomic  ratio  of 
hydrogen  to  carbon  in  this  substance? 

2.  There  are  two  known  compounds  containing 
only  tungsten  and  carbon.  One  is  the  very  hard 
alloy,  tungsten  carbide,  used  for  the  edges  of 
cutting  tools.  Analysis  of  the  two  compounds 
gives,  for  one,  1.82  grams  and,  for  the  other, 
3.70  grams  of  tungsten  per  0.12  gram  of  carbon. 
Determine  the  empirical  formula  of  each. 


3.  John  Dalton  thought  the  formula  for  water  was 
HO  (half  a  century  passed  before  the  present 
formula  for  water  was  generally  accepted).  What 
relative  weights  did  he  then  obtain  for  the  weight 
of  oxygen  and  hydrogen  atoms? 

4.  Nitrogen  forms  five  compounds  with  oxygen  in 
which  1.00  gram  of  nitrogen  is  combined  with 
0.572,  1.14,  1.73,  2.28,  2.85  grams  of  oxygen, 
respectively.  Show  that  the  relative  weights  of 
the  elements  in  these  compounds  are  in  the  ratio 
of  small  whole  numbers.  Explain  these  data  us- 
ing the  atomic  theory. 


QUESTIONS    AND    PROBLEMS 


251 


5.  Using  Appendix  3,  list  two  metals  that  could 
have  given  the  same  number  of  moles  as  alumi- 
num did  in  the  experiment  shown  in  Figure  14-2. 

6.  If  n  coulombs  will  deposit  0.1 19  gram  of  tin  from 
a  solution  of  SnS04,  how  many  coulombs  are 
needed  to  deposit  0.1 19  gram  of  tin  from  a  solu- 
tion of  Sn(S04)2? 

7.  Suppose  two  more  cells  were  attached  to  the 
three  in  Figure  14-2.  In  one  cell,  at  one  of  the 
electrodes  copper  is  being  plated  from  CuS04 
solution  and  at  one  of  the  electrodes  in  the  other 
cell,  bromine,  Rr-iig),  is  being  converted  to  bro- 
mide ion,  Br~.  How  many  grams  of  Cu  and  Br~ 
would  be  formed  during  the  same  operation 
discussed  in  the  figure? 

8.  Carbon  monoxide  absorbs  light  at  frequen- 
cies near  1.2  X  10",  near  6.4  X  1013,  and  near 
1.5  X  1015  cycles  per  second.  It  does  not  absorb 
at  intermediate  frequencies. 

(a)  Name  the  spectral  regions  in  which  it  ab- 
sorbs (see  Figure  14-14). 

(b)  Explain  why  carbon  monoxide  is  colorless. 

9.  The  wavelength  and  frequency  of  light  are  re- 
lated by  the  expression  X  =  c/v,  where  X  = 
wavelength  in  centimeters,  v  =  frequency  in 
cycles  per  second,  and  c  =  velocity  of  light  = 
3.0  X  1010  cm/second.  Calculate  the  wavelength 
corresponding  to  each  of  the  three  frequencies 
absorbed  by  CO  (see  Problem  8).  Express  each 
answer  first  in  centimeters,  and  then  in  Ang- 
stroms (1  A  =  10-8  cm). 

Answer.  1.5  X  1015  cycles/sec: 

2.0#X  10~6  cm/cycle  =  2.0  X  10s 
A/cycle. 

10.  The  oxygen  molecule  carries  out  molecular  vi- 
bration at  a  frequency  of  2.4  X  1013  cycles/sec- 
ond. If  the  pressure  is  such  that  an  oxygen 
molecule  has  about  109  collisions  per  second, 
how  many  times  does  the  molecule  vibrate  be- 
tween collisions? 


11.  When  several  oil  drops  enter  the  observation 
chamber  of  the  Millikan  apparatus,  the  voltage 
is  turned  on  and  adjusted.  One  drop  may  be 
made  to  remain  stationary,  but  some  of  the 
others  move  up  while  still  others  continue  to 
fall.  Explain  these  observations. 

12.  Dust  particles  may  be  removed  from  air  by  pass- 
ing the  air  through  an  electrical  discharge  and 
then  between  a  pair  of  oppositely  charged  metal 
plates.  Explain  how  this  removes  the  dust. 

1 3.  How  many  electrons  would  be  required  to  weigh 
one  gram?  What  would  be  the  weight  of  a 
"mole"  of  electrons? 

14.  About  how  many  molecules  would  there  be  in 
each  cubic  centimeter  of  the  tube  shown  in  Fig- 
ure 14-3  when  the  glow  appears?  When  the  glow 
disappears  again  because  the  pressure  is  too 
low? 

15.  Describe  the  spectrum  produced  on  a  photo- 
graphic plate  in  a  mass  spectrograph  if  a  mixture 
of  the  isotopes  of  oxygen  (160,  "O,  and  180)  is 
analyzed.  Consider  only  the  record  for  + 1  and 
+2  ions. 

16.  Hydroxylamine,  NH2OH,  is  subjected  to  elec- 
tron bombardment.  The  products  are  passed 
through  a  mass  spectrograph.  The  two  pairs  of 
lines  formed  indicate  charge/mass  ratios  of 
0.0625,  0.0588  and  0.1250,  0.1176.  How  can  this 
be  interpreted? 

17.  Platinum  and  zinc  have  the  same  number  of 
atoms  per  cubic  centimeter.  Would  thin  sheets 
of  these  elements  differ  in  the  way  they  scatter 
alpha  particles?  Explain. 

18.  Assume  that  the  nucleus  of  the  fluorine  atom  is 
a  sphere  with  a  radius  of  5  X  10~13  cm.  Calculate 
the  density  of  matter  in  the  fluorine  nucleus. 

19.  An  average  dimension  for  the  radius  of  a  nucleus 
is  1  X  10~12  cm  and  for  the  radius  of  an  atom  is 
1  X  10-8  cm.  Determine  the  ratio  of  atomic  vol- 
ume to  nuclear  volume. 


CHAPTER 


15 


Electrons  and  the 
Periodic  Table 


It  is  the  behavior  and  distribution  of  the  electrons  around  the  nucleus  that 
gives  the  fundamental  character  of  an  atom:  it  must  be  the  same  for  mole- 
cules. 

c.  a.  coulson,   195  1 


We  have  seen  that  much  is  known  about  the 
structure  of  the  atom.  A  small  nucleus  containing 
protons  and  neutrons  accounts  for  most  of  the 
mass  of  the  atom.  Electrons  occupy  the  space 
around  the  nucleus  like  bees  around  a  hive. 
In  the  electrically  neutral  atom,  the  number  of 
electrons  is  equal  to  the  number  of  protons. 

Looking  back  to  Chapter  6,  we  have  discov- 
ered marvelous  regularity  among  the  elements. 
Of  the  100  or  so  elements,  six  are  unique  in  their 
absence  of  chemical  reactivity.  Those  six  ele- 
ments, the  inert  gases,  provide  the  key  to  the 
most  important  correlation  of  chemistry,  the 
periodic  table.  Not  only  do  these  elements  fur- 
nish the  cornerstone  for  the  periodic  table  but, 
also,  their  electron  populations  seem  to  play  a 
dominant  role  in  the  chemistry  of  the  other  ele- 
ments in  the  table.  An  element  just  preceding  an 
inert  gas  in  the  table  (one  of  the  halogens)  has  a 
strong  tendency  to  acquire  an  extra  electron. 
The  resulting  negatively  charged  ion  has,  then, 
the  number  of  electrons  possessed  by  an  atom  of 
its  inert  gas  neighbor.  In  striking  contrast,  an 
252 


element  just  following  an  inert  gas  (one  of  the 
alkalies)  releases  electrons  quite  readily.  The  re- 
sulting positively  charged  ion  has,  then,  the 
number  of  electrons  possessed  by  an  atom  of  its 
inert  gas  neighbor.  In  each  type  of  element,  the 
halogens  and  the  alkalies,  the  chemistry  can  be 
discussed  in  terms  of  the  tendency  of  atoms  to 
acquire  or  release  electrons  so  as  to  reach  the 
special  stability  of  the  inert  gases.  The  impor- 
tance of  this  tendency  is  revealed  in  the  dramatic 
differences  that  exist  between  the  chemistry  of 
the  halogens  and  the  chemistry  of  the  alkalies. 
This  special  stability  associated  with  the  inert 
gas  electron  populations  was  found  to  pervade 
the  chemistry  of  every  element  of  the  third  row 
of  the  periodic  table  (see  Section  6-6.2).  Each 
element  forms  compounds  in  which  it  contrives 
to  reach  an  inert  gas  electron  population.  Ele- 
ments with  a  few  more  electrons  than  an  inert 
gas  are  apt  to  donate  one  or  two  electrons  to 
some  other  more  needy  atom.  Elements  with  a 
few  less  electrons  than  an  inert  gas  are  apt  to 
acquire  one  or  two  electrons  or  to  negotiate  a 


SEC.     15-1    I    THE    HYDROGEN    ATOM 


253 


communal  sharing  with  other  atoms.  In  all  cases, 
the  number  of  electrons  transferred  or  shared  is 
understandable  in  terms  of  the  inert  gas  stability. 
In  this  chapter  we  shall  explore  our  current 
understanding  of  this  behavior.  We  are  guided 
in  this  exploration  by  a  regularity  presented  in 
Chapter  6  and  reproduced  in  Table  15-1.  The 


Table  15-1 

REGULARITY  AMONG  THE  ELECTRON 
POPULATIONS  OF  THE  INERT  GASES 


INERT  GAS 

ELECTRONS 

DIFFERENCES 

helium 

2 

2 

neon 

10 

10  -    2  =    8 

argon 

18 

18  -  10  =    8 

krypton 

36 

36  -  18  =  18 

xenon 

54 

54  -  36  =  18 

radon 

86 

86  -  54  =  32 

regularity  of  the  differences  2,  8,  8,  18,  18,  32  is  a 
clue  of  magnificent  proportions.  The  electron 
populations  of  the  inert  gas  atoms  have  startling 


regularity.  This  clue  has  led  scientists  to  a  de- 
tailed and  quantitative  understanding  of  the 
atomic  properties  that  give  rise  to  the  periodic 
table. 


EXERCISE  15-1 

To  see  that  these  numbers  have  regularity,  con- 
sider the  series  of  numbers  2-8-18-32.  (We  shall 
forget,  for  the  time  being,  that  8  and  18  each 
appear  twice  in  the  series.) 

(a)  If  you  were  to  consider  this  series  incom- 
plete, would  you  expect  the  next  number 
(after  32)  to  be  even  or  odd? 

(b)  The  numbers  2-8-18-32  were  obtained  by 
subtracting  electron  populations  (that  is,  by 
taking  differences).  Take  differences  again 
and  use  them  to  predict  the  next  number 
beyond  32  in  the  series. 

(c)  Divide  the  numbers  2-8-18-32  by  two.  Use 
these  numbers  as  a  basis  for  predicting  the 
next  number  beyond  32  in  the  series  by  an- 
other method  than  taking  differences. 


15-1    THE  HYDROGEN  ATOM 


Just  as  the  inert  gases  form  the  cornerstone  of 
the  structure  we  call  the  periodic  table,  the  sim- 
plest atom,  hydrogen,  provides  the  key  that  un- 
locks the  door  of  this  structure.  Atoms  of  every 
other  element  mimic  the  hydrogen  atom.  To  see 
that  this  is  so,  we  must  examine  the  interaction 
of  hydrogen  atoms  with  light.  The  light  emitted 
(or  absorbed)  by  hydrogen  atoms  is  called  the 
atomic  hydrogen  spectrum.  This  spectrum  ex- 
plains the  existence  of  the  periodic  table. 

15-1.1    Light— A  Form  off  Energy 

Before  we  can  analyze  the  spectrum  of  hydrogen 
atoms,  we  must  become  more  familiar  with  light. 
In  Chapter  14  light  was  characterized  by  fre- 
quency or  wavelength.  (Reread  Section  14-3.1.) 
Now  we  shall  consider  another  property  of  light 


— a  property  less  obvious  than  the  "water  wave" 
characteristics.  Light  is  a  form  of  energy. 

The  statement,  "Light  is  a  form  of  energy," 
is  consistent  with  quite  a  bit  of  common  experi- 
ence. Many  of  you  have  used  a  hand  lens  to 
focus  light  rays  on  a  paper,  setting  it  afire.  This 
experiment  is  put  to  work  in  the  huge  solar 
furnaces  that  achieve  temperatures  of  many 
thousands  of  degrees  and  that  melt  the  most 
refractory  material.  The  temperature  rise  in  the 
paper  or  in  the  refractory  material  is  caused  by 
the  absorption  of  light.  A  temperature  rise  means 
energy  has  been  absorbed.  This  energy  must  have 
been  carried  by  the  light. 

We  do  not  need  a  hand  lens  to  "feel"  the 
energy  of  light  rays.  Just  remember  those  lazy 
afternoons  you  spent  last  summer  soaking  up 
the  warmth  of  the  sun.  The  afternoon  pleasure 


254 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


came  from  the  absorption  of  the  energy  by  the 
tissues  of  your  skin.  The  sunburn  pain  you  suf- 
fered that  night  was  caused  by  the  chemical 
reactions  energized  by  the  treacherous  light  rays. 
Here  we  have  a  personal  basis  for  claiming  that 
light  carries  energy. 

As  one  other  familiar  reference,  consider  pho- 
tosynthesis. You  have  undoubtedly  heard  many 
times  that  this  is  the  chemical  process  by  which 
a  plant  "stores"  the  energy  of  the  sun.  Much  is 
known  about  the  chemical  reactions  of  photo- 
synthesis and  it  is  indeed  true  that  they  result  in 
formation  of  chemical  compounds  with  higher 
heat  content  than  the  starting  substances.  These 
reactions  will  not  occur  in  the  absence  of  light — 
the  light  supplies  the  energy  required  to  raise  the 
reactants  to  the  higher  heat  content  of  the 
products. 

We  have  all  of  this  familiar  experience  to  build 
upon,  but  it  is  all  qualitative.  We  need  a  quan- 
titative relationship.  How  much  energy  is  carried 
by  light?  The  answer  is  simple  in  form,  but  not 
in  concept.  Light,  too,  comes  in  packages.  Each 
package,  called  a  photon,  contains  an  amount  of 
energy  determined  by  the  frequency.  This  state- 
ment is  contained  in  the  famous  equation 

E  =  hv  (/) 

The  quantity  h  is  called  Planck's  constant.  It  is 


merely  a  conversion  factor  that  reexpresses  fre- 
quency, v,  in  energy  units.  The  experimental 
evidence  that  led  to  equation  (/)  furnishes  a 
fascinating  story — a  story  that  you  will  hear  in 
your  physics  class.  Our  interest  is  in  the  applica- 
tion of  this  equation  to  the  interpretation  of  the 
atomic  hydrogen  spectrum.  This  spectrum  gives 
a  record  of  the  frequencies  that  are  emitted  by  a 
hydrogen  atom.  By  equation  (7),  then,  the  spec- 
trum also  gives  information  about  the  energies  a 
hydrogen  atom  can  possess. 

15-1.2    The  Light  Emitted  by  Hydrogen  Atoms 

Figure  15-1  shows  again  the  spectrograph  de- 
scribed in  Section  14-3.1.  This  time  the  light 
source  is  a  gas  discharge  tube  such  as  was  shown 
in  Figure  14-3  on  p.  239.  Just  as  before,  some  of 
the  light  emitted  by  the  source  passes  through  a 
narrow  slit  and  is  focused  to  a  beam  by  the  lens. 
Again  this  light  beam  is  refracted  (bent)  by  the 
angular  prism.  The  spectrum,  or  frequency  com- 
position of  the  light,  appears  on  the  photo- 
graphic film. 
When  hydrogen  gas  is  admitted  to  the  dis- 


Fig.  15-1.  The  spectrum  of  light  emitted  by  a  hydro- 
gen discharge  tube. 


Photographic    ptaire 


SEC.     15-1    I    THE    HYDROGEN    ATOM 


255 


Hot  tungsten    ribbon   emits 
ligrh-t  airatl  frequencies 


Visible  region 


v^L 


Ultra  vio  le  t 

region 

JJ) 


Hydrogen     discharge    i~u.be  emits 
lighT  eft  special  frequencies  only 

Fig.  15-2.  Contrast  between  the  continuous  spectrum 
of  a  hot  tungsten  ribbon  and  the  line 
spectrum    of   a    hydrogen    discharge    tube. 


charge  tube  and  a  high  voltage  is  applied,  light 
is  emitted.  To  the  eye,  the  light  appears  magenta 
in  color  but  the  spectrograph  tells  a  surprising 
story!  Instead  of  a  fairly  continuous  darkening 
across  the  photographic  film  (as  obtained  from 
the  hot  tungsten  ribbon)  the  film  shows  a  series 
of  lines!  Each  line  corresponds  to  a  particular 
frequency  emitted  by  hydrogen  atoms.  Each 
space  between  two  lines  on  the  film  corresponds 
to  a  frequency  range  in  which  no  light  is  emitted 
by  the  hydrogen  atom. 

Fig.  15-3.  The  spectrum  of  the  hydrogen  atom. 
Visible  region  Ultraviolet-  region 


VO    «0    On  N.*>«xo         Energy  ~\      CQ  Q^Os 

S    £  ^  ££'£     of  photons,    Vj 
(k  cal/m.  ole)    <\ 

vO    ^0   Q  ^Vrv^>,   Frequency,  ^ 
g   3  S  S§£  §  (cycles/sec)  \ 

*    $  ^  ^  £§    Wavelength,  ^ 


o     ^ 


ty  ^Ov^o)Q 


S<s^S 


?  s>*3 


<i    ©\OCOn 


Two  peculiarities  of  the  hydrogen  discharge 
spectrum  are  immediately  evident — they  were 
evident  to  scientists  as  long  ago  as  1840.  First, 
hydrogen  atoms  are  very  particular  about  the 
frequencies  they  emit.  Only  special  frequencies 
are  observed  in  the  spectrum.  Second,  the  fre- 
quencies corresponding  to  the  lines  on  the  film 
are  spaced  systematically.  There  are  two  groups 
of  lines,  one  group  in  the  visible  part  of  the  light 
spectrum  and  another  group  in  the  ultraviolet 
part.  Within  each  group  there  is  a  regular  de- 
crease in  the  spacing  between  successive  lines  as 
the  frequency  increases.  The  measured  frequen- 
cies are  shown  in  Figure  15-3. 


EXERCISE  15-2 

Complete  the  following  table  for  all  of  the  ultra- 
violet lines  listed  in  Figure  15-3.  Plot  the  energy 
spacing  against  the  arbitrary  spacing  number 
assigned  in  the  last  column  to  convince  yourself 
that  there  is  regularity  in  the  spacings  of  these 
lines.  Make  the  same  sort  of  a  table  for  the  lines 
in  the  visible  group. 


ENERGY  PER 

ENERGY 

SPACING 

GROUP 

MOLE  OF  PHOTONS 

SPACING 

NUMBER 

235.2  kcal 

Ultraviolet 

278.8 

43.6  kcal 

1 

lines 

294.0 

2 
3 

To  see  how  these  results  might  be  explained, 
let  us  translate  these  spectral  lines  into  energy 
terms.  The  hydrogen  atom  just  before  light  emis- 
sion has  some  amount  of  energy — let  us  call  it 
E2.  Light  of  frequency  ■>  is  emitted,  carrying 
away  energy  hv.  The  hydrogen  atom  is  left  with 
less  energy — let  us  call  it  Ex.  As  is  habitual,  we 
assume  energy  is  conserved,  so  the  energy  lost 
by  the  hydrogen  atom  must  be  exactly  equal  to 
that  carried  away  by  the  light: 

E2  -  £,  =  hv  (2) 

To  "explain"  why  hydrogen  atoms  emit  the 
line  spectrum,  we  must  seek  a  model  with  the 


256 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


No-tch.     tt  1 


ZSS.Z  Ac  at 


same  sort  of  properties.  What  sort  of  a  system 
has  the  property  that  its  initial  energy,  E2,  tells 
it  how  much  energy  it  can  release  so  as  to  make 
the  difference  E2  —  Ex  just  exactly  one  of  the 
special  energies,  hvl  Fortunately,  we  have  just 
such  a  system  at  hand,  though  it  deals  in  weights, 
not  energies. 

Picture  your  triple  beam  balance.  The  front 
beam  has  a  sliding  hanger  that  permits  you  to 
balance  any  weight  placed  on  the  pan  up  to  the 
full  scale  reading  on  the  beam.  This  front  beam 
might  be  compared  to  the  hot  tungsten  ribbon. 
The  beam  can  be  used  to  balance  any  small 
weight  change  made  on  the  pan:  the  hot  tungsten 
ribbon  emits  light  of  any  energy  (that  is,  of  any 
frequency). 

But  now  consider  the  other  two  beams — they 
are  quite  different.  They  are  notched  so  that  only 
particular  hanger  positions  can  be  selected.  Each 
position  corresponds  to  a  particular  weight  and 
intermediate  weights  cannot  be  balanced  without 
the  front  beam.  The  notched  beam  has  properties 
in  common  with  the  hydrogen  atom.  The  weight 
added  to  the  notched  beam  is  fixed  by  the 
notches — intermediate  weights  cannot  be  meas- 
ured. In  the  hydrogen  atom  the  energy  is  some- 
how "hung  on  a  notched  beam"  and  the  amount 
of  energy  it  can  absorb  or  release  must  corre- 
spond to  the  energy  difference  between  two  of 
these  "notches." 

If  we  pursue  this  analogy,  we  can  see  how 


Fig.  15-4.  The  notches  implied  by  light  emission  at 
235.2  kcal. 


scientists  have  deduced  the  energy  properties  of 
the  hydrogen  atom  from  its  line  spectrum.  We 
can  use  the  observed  energies,  listed  in  Figure 
15-3,  to  construct  a  notched  beam  that  would 
"deliver"  the  observed  spectrum.  This  beam 
must  have  a  scale  calibrated  not  for  weight  but 
for  energy.  There  must  be  notches  on  this  scale 
to  match  the  observed  lines  in  the  spectrum.  We 
shall  find  it  convenient  to  begin  with  the  line 
observed  at  wavelength  1216  A  or,  in  energy 
units,  235.2  kcal/mole  of  photons.  This  energy 
represents  the  difference  in  energy  between  two 
notches.  If  we  number  the  notches  #1  and  #2, 
changing  an  imaginary  hanger  from  notch  #1  to 
notch  #2  delivers  235.2  kcal,  as  shown  in  Figure 
15-4. 

Now  consider  the  other  lines  in  the  ultraviolet 
group.  Exercise  15-2  showed  that  they  are  related 
to  the  one  at  235.2  kcal  in  their  systematic  spac- 
ings.  Let's  seek  regularity  among  these  energies 
by  making  a  simple  assumption.  We  shall  assume 
that  each  line  in  the  ultraviolet  group  corre- 
sponds to  a  movement  of  the  hanger  from  notch 
#1  to  a  new  notch.  This  assumption  is  arbitrary 
and  will  be  kept  only  if  it  turns  out  to  be  helpful. 
(Remember  the  lost  child  and  his  rule,  "Cylin- 
drical objects  burn"?) 


SEC.     15-1    I    THE    HYDROGEN    ATOM 


257 


Fig.  15-5.  The    notches   required    by    the    ultraviolet 
group  of  lines  if  all  changes  begin  in  notch 


So  we  can  add  four  more  notches  to  the  beam, 
as  shown  in  Figure  15-5.  Notch  #3  is  cut  at  a 
scale  position  278.8  kcal  relative  to  notch  #1. 
Notch  #4  is  at  294.0  kcal,  and  so  on. 

The  appearance  of  this  notched  beam  is  un- 
familiar, to  say  the  least.  Can  we  find  any  basis 


for  corroboration  or  contradiction  of  this  inter- 
pretation? To  seek  such  evidence,  let's  base  a 
prediction  upon  the  model.  One  such  prediction 
concerns  the  energies  that  this  notched  beam 
would  deliver  if  we  investigated  movements  of 
the  hanger  from  notch  #2  to  notch  #3.  Notch  #2 
is  235.2  kcal  from  notch  #1,  and  notch  #3  is 


Fig.  15-6.  The  energy  change  implied   by  a  change 
from  notch  #2  to  notch  #3. 


Notch.    #  / 


3      45  6 


(278.8  -235.2) 
-43.6  Ucat 


258 


ELECTRONS    AND    THE    PERIODIC    TABLE   I    CHAP.    15 


278.8  kcal  from  notch  #1.  Therefore,  the  energy 
difference  going  from  notch  #2  to  #3  is  the  dif- 
ference between  these  two  numbers: 

Energy  change  from  notch  #2  to  #3 

=  278.8  -  235.2  =  43.6    (5) 

Calculation  (3)  indicates  that  our  notched  beam 
model  implies  light  would  be  emitted  with  en- 
ergy at  43.6  kcal  as  well  as  at  235.2  kcal  and 
278.8  kcal.  We  refer  back  to  the  spectrum  in 
Figure  15-3  and,  indeed,  there  is  light  emitted  at 
43.6  kcal — this  is  one  of  the  group  of  lines  in  the 
visible  region! 

With  this  encouragement,  let's  calculate  the 
other  lines  implied  by  changes  beginning  in 
notch  #2. 


NOTCH 

ENERGY  CHANGE 

CHANGE 

CALCULATED 

OBSERVED 

#2  — >-#3 

278.8 

-  235.2  =  43.6  kcal 

43.6  kcal 

#2  —^#4 

294.0 

-  235.2  =  58.8 

58.8 

#2  — *-#5 

301.1 

-  235.2  =  65.9 

65.9 

#2  -+#6 

304.9 

-  235.2  =  69.7 

69.7 

Plainly,  the  agreement  between  calculation 
and  experiment  is  too  good  to  be  accidental. 
Our  notched  beam  is  a  useful  basis  for  interpret- 
ing the  spectrum  of  the  hydrogen  atom. 


EXERCISE  15-3 

Plot  the  energy  of  each  line  of  the  ultraviolet 
group  against  notch  number,  n,  using  the  higher 
of  the  two  notch  numbers  assigned  to  that  line 
in  Figure  15-5.  For  example,  plot  235.2  kcal  on 
the  vertical  axis  against  2  on  the  horizontal  axis. 
Assign  to  notch  #1  the  arbitrary  value  zero  and 
draw  a  smooth  curve  through  all  of  the  points, 
including  the  point  for  notch  #1.  Estimate  the 
energy  value  that  would  be  observed  for  a  notch 
with  very  high  notch  number,  as  suggested  by 
the  curve.  (Call  this  "notch  #  infinity,"  n  =  <x>.) 


300 


200 


Energy  based  cm 
notch**  1=0 


lOO 


0 -i 


301.1 
294.0 


278.8 


235.2 


\-12.5 
\-19.6 

-34.d 


73.4 


-313.6 


n 


-100 


JSneryy  based  on 
notch**  00=0 


-200 


-300 


Fig.  15-7.  The  energy  level  scheme  of  the  hydrogen 
atom. 


Historically,  the  visible  emission  lines  shown  in  Figure 
15-3  were  the  first  atomic  hydrogen  lines  discovered.  They 
were  found  in  the  spectrum  of  the  sun  by  W.  H.  Wollaston 
in  1802.  In  1862,  A.  J.  Angstrom  announced  that  there 
must  be  hydrogen  in  the  solar  atmosphere.  These  lines 
were  detected  first  because  of  the  lesser  experimental  dif- 
ficulties in  the  visible  spectral  region.  They  are  called  the 
"Balmer  series"  because  J.  J.  Balmer  was  able  to  formu- 
late a  simple  mathematical  relation  among  the  frequencies 
(in  1?<?^).  The  ultraviolet  series  shown  in  Figure  15-3  was 


SEC.    15-1    I    THE    HYDROGEN    ATOM 


259 


actually  predicted  prior  to  its  discovery  (in   1915  by 
T.  Lyman). 

Hence  our  analysis  and  prediction,  selected  for  logical 
clarity,  reverse  the  actual  chronology.  Our  prediction  of 
the  visible  spectrum  from  the  ultraviolet  spectrum  is  more 
straightforward  than  was  the  reverse  prediction.  Notice 
that  the  ultraviolet  spectrum  utilizes  all  of  the  notches 
involved  in  producing  the  visible  spectrum.  In  contrast, 
none  of  the  visible  frequencies  involves  notch  #1,  the  key 
to  the  ultraviolet  spectrum. 


15-1.3    The  Energy  Levels  off  a  Hydrogen  Atom 

Scientists  deduced  the  notched  energy  scale  of 
the  hydrogen  atom  from  the  spectrum  in  just  the 
same  way  we  did. 

Of  course,  more  sophisticated  language  is  gen- 
erally used.  For  example,  the  energy  possibilities 
are  usually  shown  on  a  vertical  scale  and  they 
are  called  energy  levels  rather  than  notches.  Fig- 
ure 15-7  shows  the  energy  level  scheme  of  the 
hydrogen  atom.  Each  energy  level  is  character- 
ized by  an  integer,  n,  the  lowest  level  being  given 
the  number  1 . 

Two  energy  scales  are  shown  in  Figure  15-7. 
On  the  left  is  a  scale  based  upon  the  energy  zero 
for  the  n  =  1  level.  On  this  scale,  the  levels 
appear  at  energies  corresponding  to  our  notched 
beam  spacings.  The  right-hand  scale  is  displaced 
upward  so  that  the  n  =  1  level  corresponds  to  a 
negative  energy,  —313.6  kcal.  The  zero  has  been 
moved  upward  by  this  amount  so  that  the  zero 
of  the  energy  scale  corresponds  to  the  "notch  # 
infinity"  that  you  estimated  in  Exercise  15-3. 
Obviously,  the  positioning  of  the  energy  levels  is 
not  affected  by  this  arbitrary  change  of  the  zero 
of  energy,  so  either  scale  can  be  used.  For  rea- 
sons that  will  become  evident  later  in  this  chap- 
ter, the  right-hand  scale  is  more  convenient  and 
is  the  one  generally  used. 


EXERCISE  15-4 

Calculate  the  energy  change  that  occurs  between 
notch  #1  and  notch  #2,  using  the  right-hand  scale 
in  Figure  15-7.  (Remember  that  the  energy 
change  is  the  energy  of  the  final  level  minus  the 
energy  of  the  initial  level.  Pay  careful  attention 
to  algebraic  signs.)  Repeat  the  calculation  for  the 


energy  change  between  notch  #2  and  #3.  Com- 
pare your  calculations  with  the  numbers  shown 
in  Figure  15-6. 


Long  after  this  energy  level  diagram  for  the 
hydrogen  atom  had  been  established,  scientists 
still  pondered  its  significance.  Finally,  in  the  late 
1920's,  a  mathematical  scheme  was  developed 
that  explained  the  facts.  The  mathematical 
scheme  is  called  quantum  mechanics. 

15-1.4    Quantum  Mechanics  and  the 
Hydrogen  Atom 

Prior  to  the  development  of  quantum  mechanics, 
the  spectrum  of  the  hydrogen  atom  posed  quite  a 
dilemma.  To  see  the  problem,  and  how  it  was 
resolved,  let's  go  back  about  fifty  or  sixty  years 
and  trace  the  history  of  this  problem.  This  is  a 
valuable  example  because  it  shows  how  science 
advances. 

By  the  year  1912  it  was  known  that  the  hydro- 
gen atom  is  composed  of  a  proton  and  an  elec- 
tron. These  two  particles  are  attracted  to  each 
other  by  reason  of  their  electric  charges.  Physi- 
cists felt  they  should  be  able  to  calculate  the 
properties  of  such  a  combination.  After  all,  the 
laws  of  motion  of  macroscopic  bodies  had  been 
studied  for  centuries.  The  behavior  of  electrically 
charged  bodies  was  also  thoroughly  understood 
on  the  macroscopic  scale.  Yet  scientists  could 
not  explain  why  a  hydrogen  atom  exists,  let 
alone  why  it  would  have  only  particular  values  of 
energy.  In  fact,  the  laws  that  had  been  deduced 
for  macroscopic  bodies  gave  the  firm  (though 
incorrect)  prediction  that  the  nuclear  atom  is  un- 
stable and  the  electron  should  collapse  into  the 
nucleus. 

At  this  point  a  Danish  physicist,  Niels  Bohr, 
decided  to  take  a  fresh  start.  In  effect,  he  faced 
the  fact  that  an  explanation  is  a  search  for  like- 
nesses between  a  system  under  study  and  a  well- 
understood  model  system.  An  explanation  is  not 
good  unless  the  likenesses  are  strong.  Niels  Bohr 
suggested  that  the  mechanical  and  electrical  be- 
havior of  macroscopic  bodies  is  not  a  completely 
suitable  model  for  the  hydrogen  atom.  He  pro- 


260 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


ceeded  to  seek  a  new  model  that  did  not  contra- 
dict the  known  facts. 

He  began  by  supposing  that  the  structure  of 
an  atom  (the  arrangement  of  the  electrons 
around  the  nucleus)  is  determined  by  its  energy. 
To  agree  with  the  facts,  Bohr  proposed  that  only 
special  atomic  structures  can  exist — he  called 
these  special  structures  "stationary  states."  Each 
such  state  is  characterized  by  a  particular  energy, 
and  since  a  set  of  special  atomic  structures  exists, 
a  corresponding  set  of  energies  will  be  found. 
Here  Bohr  departed  from  the  older  atomic 
models  (those  of  classical  physics)  that  permitted 
structures  corresponding  to  all  possible  energies. 

The  most  stable  state  of  the  atom  would  be 
expected  to  be  the  one  in  which  the  atom  has  the 
lowest  energy.  Bohr  reasoned  that  since  we  ob- 
serve that  the  nuclear  atom  does  exist  then  it 
must  be  a  fundamental  fact  of  nature  that  an 
atom  can  exist  in  its  most  stable  state  indefinitely. 
Even  though  this  fact  could  not  be  rationalized 
(remember,  the  earlier  laws  of  physics  predicted 
the  atom  should  collapse)  it  had  to  be  accepted 
because  it  was  a  result  of  experiments. 

Bohr  also  proposed  that  although  the  lowest 
energy  state  of  the  atom  is  its  most  stable  state, 
the  atom  can  be  excited  to  its  higher  allowed 
energy  states  (by  absorbing  light  or  through  a 
violent  collision  with  other  atoms  or  electrons). 
The  excited  atom  does  not  remain  in  this  condi- 
tion for  long;  it  loses  its  extra  energy  by  emitting 
light.  Since  only  certain  levels  of  energy  exist, 
only  certain  energy  changes  can  occur.  The  en- 
ergy change  of  the  atom  must  be  equal  to  the 
energy  of  the  light  emitted,  in  accord  with  equa- 
tion (2),  E2  —  Ei  =  hv.  Consequently,  the  fre- 
quency of  the  light  emitted  by  an  atom  is  entirely 
determined  by  the  values  of  the  allowed  energies 
of  its  electrons. 

These  ideas  were  so  revolutionary  that  they 
would  not  have  been  accepted  except  for  the  fact 
that  Bohr  was  able  to  propose  a  way  to  calculate 
exactly  the  energy  levels  for  the  hydrogen  atom. 
Within  ten  years  Bohr's  calculational  methods 
were  completely  replaced  by  better  techniques, 
but  his  postulate  that  only  certain  atomic  energy 
states  are  possible  has  been  repeatedly  shown  to 
be  correct. 


In  this  example  there  is  much  food  for  thought  con- 
cerning the  development  of  science.  The  wide  applicability 
of  the  laws  of  motion  and  of  electromagnetics  made  it 
natural  for  scientists  to  assume  that  these  same  laws, 
without  change,  applied  to  the  atom.  True,  this  repre- 
sented an  extrapolation,  for  the  laws  were  deduced  on  the 
macroscopic  level.  Yet,  the  same  laws  that  described  the 
motions  of  the  planets  also  described  the  motions  of 
tennis  balls — why  not  also  the  motions  of  electrons? 
Many  experimental  facts  said  no,  but  physicists  held  to 
the  expectation  that  a  way  would  be  found  to  explain 
these  facts  within  the  framework  of  the  known  (and 
almost  sacred)  laws.  When  Bohr  finally  broke  away  from 
the  established  laws,  he  still  used  them  for  guidance, 
proposing  only  those  changes  required  by  the  discordant 
facts.  Perhaps  the  most  ironic  part  is  that  Bohr's  principal 
weapon  in  gaining  acceptance  for  his  new  attack  was  his 
mathematical  success  in  predicting  the  energy  levels  of 
hydrogen,  though  his  model  has  since  been  discarded 
completely.  The  model  he  used  proved  to  fit  only  the 
hydrogen  atom  and  no  other. 

Don't  be  too  eager,  though,  to  scoff  at  this  example. 
Instead,  you  may  rest  assured  that  some  of  the  theories 
you  will  find  in  this  book  are  waiting  to  be  swept  aside 
as  we  learn  more  about  nature.  The  rub — no,  the  excite- 
ment— in  the  game  is  that  we  don't  know  which  theories 
are  fated  to  go.  That  remains  for  some  of  you  to  discover! 


15-1.5    The  Hydrogen  Atom  and 
Quantum  Numbers 

The  modern  theory  of  the  behavior  of  matter, 
called  quantum  mechanics,  was  developed  by 
several  workers  in  the  years  1925-1927.  For  our 
purposes  the  most  important  result  of  the  quant- 
um mechanical  theory  is  that  the  motion  of  an 
electron  is  described  by  the  quantum  numbers 
and  orbitals.  Quantum  numbers  are  integers 
that  identify  the  stationary  states  of  an  atom; 
the  word  orbital  means  a  spatial  description  of 
the  motion  of  an  electron  corresponding  to  a  par- 
ticular stationary  state. 

THE    PRINCIPAL    QUANTUM    NUMBER,    n 

If  we  return  to  our  notched  beam  analogy,  as 
shown  in  Figure  15-5,  we  find  that  we  numbered 
the  notches,  1,  2,  3,  •••.  These  numbers  serve 
as  natural  identifying  designations.  They  are  the 
"quantum  numbers"  of  the  balance  beam. 

In  Exercises  15-2  and  15-3  you  observed  that 
the  energy  levels  of  hydrogen  vary  systematically 
with    the   quantum   (or   notch)   number.    The 


SEC.     15-1    I    THE    HYDROGEN    ATOM 


261 


smooth  curves  suggest  that  the  energy  could  be 
conveniently  expressed  mathematically  in  terms 
of  n,  the  quantum  number.  Using  the  right-hand 
scale  in  Figure  15-7,  we  see  that  for  any  value 
of  n,  E  is  always  negative.  As  n  becomes  larger, 
the  energy  rises  and  approaches  zero  on  the 
scale.  Investigation  of  the  actual  energy  values 
shows  that  the  energy  levels  of  Figure  15-7  are 
exactly  determined  by  n  according  to  the  relation 

£„  =  — *  kcal  mole  (4) 

ir 

En  =  energy  level  with  quantum  number  n 
n  =  1,2,3,  ••■  x 

The  number  n  is  called  the  principal  quantum 
number. 

The  mathematical  relationship  (4)  is  the  one  Bohr  was 
able  to  deduce.  Current  quantum  mechanical  methods 
also  deduce  this  relationship,  of  course,  but  with  a  model 
that  is  in  fundamental  discord  with  the  one  used  by  Bohr. 

ORBITALS    AND    THE    PRINCIPAL 
QUANTUM    NUMBER,    n 

Quantum  mechanics  provides  a  mathematical 
framework  that  leads  to  expression  (4).  In  addi- 
tion, for  the  hydrogen  atom  it  tells  us  a  great 
deal  about  how  the  electron  moves  about  the 
nucleus.  It  does  not,  however,  tell  us  an  exact 
path  along  which  the  electron  moves.  All  that 
can  be  done  is  to  predict  the  probability  of  find- 
ing an  electron  at  a  given  point  in  space.  This 
probability,  considered  over  a  period  of  time, 
gives  an  "averaged"  picture  of  how  an  electron 
behaves.  This  description  of  the  electron  motion 
is  what  we  have  called  an  orbital. 

Thus,  an  orbital  description  of  the  motion  of 
an  electron  contains  the  same  information  con- 
veyed by  the  holes  made  by  darts  in  a  dartboard. 
After  the  board  has  been  used  in  many  games, 
the  distribution  of  holes  shows  how  successful 
earlier  players  had  been  in  their  scoring.  There 
are  many  holes  near  the  bullseye  and,  moving 
away  from  it,  there  is  a  regular  decrease  in 
the  number  of  holes  per  square  centimeter  of 
dartboard.  At  any  given  distance  from  the  bulls- 
eye,  the  "density"  of  dart  holes  (number  per 
square  centimeter)  is  a  measure  of  the  probability 
that  the  next  throw  will  land  there. 


We  see  that  the  holes  in  the  dartboard  tell  us 
only  the  probability  that  a  given  throw  will  land 
a  particular  distance  from  the  bullseye.  It  does 
not  tell  us  the  order  in  which  the  holes  were  made 
in  the  dartboard.  In  the  case  of  the  electron  dis- 
tribution, the  orbital  tells  us  the  probability  that 
an  experiment  designed  to  locate  the  electron 
will  find  it  a  particular  distance  from  the  nucleus. 
It  does  not  tell  us  how  the  electron  moves  from 
point  to  point—  its  trajectory. 

Though  quantum  mechanics  does  not  tell  us 
the  electron  trajectory,  it  does  tell  us  how  the 
orbital  changes  as  n  increases.  It  also  indicates 
that  for  each  value  of  n  there  are  n2  different 
orbitals.  For  the  hydrogen  atom,  the  n2  orbitals 
for  a  particular  value  of  n  all  have  the  same  en- 
ergy, 

313.6  kcal 
n-     mole 

S    ORBITALS 

Consider  the  lowest  energy  level  of  a  hydrogen 
atom,  n  =  1 .  We  have  just  learned  that  there  are 
n2  levels  with  this  energy,  and  since  n  =  1,  there 
is  but  one  level.  It  corresponds  to  an  electron 
distribution  that  is  spherically  symmetrical 
around  the  nucleus,  as  shown  in  Figure  15-8.  It 
is  called  the  Is  orbital.*  An  electron  moving  in 
an  s  orbital  is  called  an  s  electron. 

The  term  "spherically  symmetric"  and  the 
picture  of  an  s  orbital  (Figure  15-8)  must  be 
clearly  understood.  They  indicate  that  if  we  were 
to  look  for  the  electron  somewhere  on  the  sur- 
face of  a  sphere  of  a  particular  radius,  ru  which 
has  the  nucleus  at  its  center,  the  probability  of 
finding  the  electron  at  any  one  point  on  the  r\ 
sphere  is  the  same  as  the  probability  of  finding 
it  at  any  other  point  on  the  rx  sphere.  The  same 
would  be  true  at  a  different  radius,  r2,  but  the 
probability  of  finding  the  electron  somewhere  on 
the  r2  sphere  would  not  be  the  same  as  that  on 
the  rx  sphere.  The  chance  of  finding  the  electron 
does  depend  upon  the  radius  of  the  sphere  on 


*  In  the  symbol  \s,  the  number  1  tells  us  that  n  =  1. 
The  letter  s  tells  us  that  the  orbital  is  spherically  sym- 
metric. Since  the  letter  s  has  been  used  for  an  orbital  that 
is  spherically  symmetric,  we  might  as  well  think  of  it  as 
an  abbreviation :  s  =  spherical. 


262 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


1  S    orbita.1 

y 


-313.6/4  =  -78.4  kcal/mole.  One  of  them  is 
again  spherically  symmetric  and  it  is  called  the 
25  orbital.  Figure  15-8  shows  the  atomic  2s 
orbital.  Here  we  find  the  reasonable  result  that 
the  higher  energy  of  the  2s  electron  permits  the 
electron  to  spend  more  time  far  from  the  nucleus. 
For  every  value  of  n,  there  is  one  spherically 
symmetric  orbital.  As  n  increases,  the  ns  orbitals 
place  the  electron,  on  the  average,  farther  and 
x      farther  from  the  nucleus. 


2  s    orbi-tal 
Fig.  15-8.  Atomic  orbitals:  Is  and  2s  orbitals. 


which  we  search.  A  \s  electron  can  be  found  any- 
where from  right  at  the  nucleus  to  a  great  dis- 
tance away — but  it  is  most  likely  to  be  found 
approximately  10~8  cm  from  the  nucleus. 

The  next  energy  level  corresponds  to  n  =  2. 
According  to  our  rule,  there  are  n2  =  22  =  4  dif- 
ferent spatial  arrangements  with  the  same  energy, 


p    ORBITALS 

There  is  only  one  orbital  corresponding  to  n  =  1 , 
the  Is  orbital.  For  n  =  2,  there  are  four  different 
spatial  arrangements  and  we  have  described  one 
of  them,  the  2s  orbital.  The  other  three  are  called 
2p  orbitals.  An  electron  in  a  p  orbital  behaves 
in  such  a  way  that  it  is  most  likely  found  in  either 
of  two  regions  located  on  opposite  sides  of  the 
nucleus.  The  motion  of  a  p  electron  creates  an 
electron  distribution  that  is  shaped  somewhat 
like  a  dumbbell.  We  can  place  the  axis  of  the 
dumbbell  along  one  of  the  three  perpendicular 
cartesian  coordinate  axes.  Just  as  there  are  three 
distinct  coordinate  axes,  there  are  three  distinct 
p  orbitals,  each  with  its  axis  perpendicular  to  the 
other  two.  They  are  sometimes  referred  to  as  the 
pz,  py,  and  pz  orbitals  to  emphasize  their  direc- 
tional character.  The  p£  orbital  is  concentrated 
in  the  x  direction — a  px  electron  is  more  apt  to 
be  found  near  the  x  axis  than  anywhere  else.  The 
pu  orbital,  on  the  other  hand,  is  concentrated 
along  the  y  axis.  These  directional  characteristics 
are  useful  in  explaining  the  geometrical  proper- 
ties of  molecules.  Figure  1 5-9  shows  the  electron 
distribution  of  the  2p  orbitals.* 

Every  energy  level  with  n  above  1  has  three  p 
orbitals.  As  n  increases,  the  np  orbitals  place  the 
electron,  on  the  average,  farther  and  farther  from 
the  nucleus,  but  always  with  the  axial  directional 
properties  shown  in  Figure  1 5-9. 

d   AND  /  ORBITALS 

At  this  point  we  might  recast  the  hydrogen  atom 
energy  level  diagram  to  express  what  we  know 

*  The  three  p  orbitals  can  be  considered  to  extend  along 
the  three  perpendicular  axes,  x,  y,  and  r.  Hence,  p  might 
be  thought  of  as  an  abbreviation:  p  =  /perpendicular. 


SEC.     15-1    I    THE    HYDROGEN    ATOM 


263 


2p  x 
orbital 

Fig.  15-9.  Atomic  orbitals:  the  2p  orbitals. 


about  orbitals.  Figure  15-10  shows  each  orbital 
as  a  pigeon  hole.  For  a  given  value  of  n,  there 
are  n2  total  orbitals.  For  n  =  3,  there  are  32  =  9 
orbitals,  five  more  than  accounted  for  by  one  3s 
orbital  and  the  three  3p  orbitals.  These  five  or- 
bitals are  called  d  orbitals  and  they  have  more 
complicated  spatial  distribution  than  do  the  p 
orbitals. 


EXERCISE  15-5 

From  the  information  that  the  numbers  of  s,  p, 
and  d  orbitals  are  1,  3,  and  5,  how  many  of  the 
next  higher  (/)  orbitals  would  you  expect? 
Verify  your  answer  by  calculating  n2  for  n  =  4 
and  comparing  to  your  sum  of  the  numbers  of 
s,  p,  d,  and /orbitals. 


15-1.6    The  Hydrogen  Atom  and 
the  Periodic  Table 

At  last  we  are  ready  to  return  to  the  periodic 
table.  At  last  we  are  able  to  begin  answering 
those  who  are  "wondering  why"  about  the  spe- 
cial properties  of  the  electron  populations  in 
Table  15-1.  Let  us  reproduce  Table  15-1  together 
with  the  numbers  of  orbitals  of  the  hydrogen 
atom.  The  suggestion  of  a  connection  is  irresist- 
ible, as  seen  in  Table  15-11. 

The  hydrogen  atom  orbitals  give  us  the  num- 
bers 2,  8,  18,  and  32 — the  numbers  we  find 
separating  the  specially  stable  electron  popula- 
tions of  the  inert  gases.  It  was  necessary  to 
multiply  n2  by  two — an  important  factor  that 
could  not  have  been  anticipated.  Furthermore, 
it  will  be  necessary  to  find  an  explanation  for  the 
occurrence  of  eight-electron  differences  both  at 
neon  and  at  argon  and  eighteen-electron  differ- 
ences both  at  krypton  and  at  xenon. 

Nevertheless,  we  seem  to  have  a  significant 
start  toward  explaining  the  periodic  table.  We 


Table  15-II.     stable  electron  populations  and  the  hydrogen  atom 

THE  INERT  GASES  THE  HYDROGEN  ATOM 


number 

of 

number  of 

element 

electrons 

differences 

n 

orbitals 

2X«' 

helium 

2 

2 

n  =  1 

n2  =     1 

2X1=2 

neon 

10 

10  -    2  =    8 

n  =  2 

W2  -     4 

2X4=8 

argon 

18 

18  -  10  =    8 

n  =  3 

„2  =    9 

2  X    9=18 

krypton 

36 

36  -  18  =  18 

n  =  4 

rt2  =   16 

2  X  16  =  32 

xenon 

54 

54  -  36  =  18 

radon 

86 

86  -  54  =  32 

264 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


-100 


Energy, 
heal/mole 


©•  •©©©  ■•@®@®@ 


-t 


-200 


-300 


Us) 


Fig.  15-10.  The  energy  level  scheme  of  the  hydrogen 
atom. 


can  understand  the  chemical  trends  within  the 
first  ten  elements  in  terms  of  the  hydrogen  atom 
orbitals  with  the  following  two  assumptions: 

(1)  that  atoms  of  these  elements  have  orbitals 
and  energy  levels  that  are  qualitatively  like 
those  of  the  hydrogen  atom; 

(2)  that  a  single  orbital  (of  any  element)  can 
accommodate,  at  most,  two  electrons. 


The  first  assumption  springs  from  the  similarities 
noticed  in  Table  15-11  and  is  well  substantiated 
by  a  wealth  of  spectral  study  of  these  atoms.  The 
second  assumption  is  stimulated  by  the  factor  of 
two  needed  in  the  last  column  of  Table  15-11. 
With  these  two  assumptions,  we  can  propose 
the  electronic  arrangement  of  lowest  energy  for 
each  atom.  We  do  so  by  mentally  placing  elec- 
trons successively  in  the  empty  orbitals  of  lowest 
energy.  The  electron  orbital  of  lowest  energy  is 
the  Is  orbital.  The  single  electron  of  the  hydro- 
gen atom  can  occupy  this  orbital.  In  the  helium 


SEC.     15-2    |    MANY-ELECTRON    ATOMS 


265 


atom,  there  are  two  electrons  and  the  nuclear 
charge  is  two.  Since  each  orbital  can  accommo- 
date two  electrons,  both  of  the  electrons  can  go 
into  the  Is  orbital.  The  resulting  electronic  ar- 
rangement is  described  by  the  notation  Is2,  which 
means  that  there  are  two  electrons  in  the  Is 
orbital.  The  notation  Is2  is  called  the  electron 
configuration. 

Now  let  us  consider  what  will  happen  when 
there  are  three  electrons  near  a  triply  charged 
nucleus,  as  in  the  lithium  atom.  The  first  two  of 
the  three  electrons  go  into  the  lowest  energy 
orbital,  the  Is  orbital.  When  this  orbital  has  two 
electrons  in  it,  it  is  completely  filled,  according 
to  our  second  assumption.  Any  additional  elec- 
trons must  be  placed  in  orbitals  of  higher  energy. 
Therefore,  the  third  electron  in  lithium  goes  into 
the  2s  orbital,  and  we  write  the  electron  configu- 
ration as  ls22s'.  Despite  the  nuclear  charge  of 
three  in  the  lithium  atom,  this  last  electron  is 
rather  weakly  bound  because  the  2s  electron  in 
lithium  spends  most  of  its  time  farther  away  from 
the  nucleus  than  do  the  Is  electrons.  This  elec- 
tron should  be  easily  removed  to  give  Li+,  which 
is,  indeed,  the  characteristic  behavior  of  an 
alkali  element. 

The  beryllium  atom  has  one  more  electron 
than  does  the  lithium  atom.  The  fourth  electron 
that  enters  the  beryllium  atom  can  occupy  the  2s 
orbital  to  give  a  configuration  of  ls22s2.  The  two 
2s  electrons  will  be  most  easily  removed,  tending 


to  form  the  Be+-  ion. 

There  is  no  more  room  in  the  2s  orbital  for  a 
fifth  electron,  which  appears  when  we  move  on 
to  the  boron  atom.  However,  another  orbital 
with  principal  quantum  number  2  is  available. 
A  2p  orbital  accepts  the  fifth  electron,  giving  the 
configuration  \s22s22pl.  Continuing  this  process, 
we  obtain  the  following  configurations: 


carbon  atom 

Is2 

2s22p\2p\ 

nitrogen  atom 

Is2 

2s22p\2p\2p\ 

oxygen  atom 

Is2 

2s22p\2p\2p\ 

fluorine  atom 

Is2 

2s22p]2pl2p\ 

neon  atom 

Is2 

2s22plx2p\2p\ 

If  we  proceed  to  the  next  element,  sodium  atom, 
we  are  again  forced  to  use  an  orbital  with  the 
next  higher  quantum  number: 

sodium  atom  Is2      2s22p\2p\l2p\       3s1 

Again  there  is  one  electron  which  spends  most 
of  its  time  farther  away  from  the  nucleus  than 
any  of  the  others.  This  one  electron  could  be 
easily  removed  to  give  Na+,  and  we  return  to  the 
type  of  chemistry  shown  by  lithium. 

The  hydrogen  atom  energy  levels,  together 
with  our  two  assumptions,  have  provided  a  good 
explanation  of  some  of  the  properties  of  the  first 
eleven  elements.  We  shall  see  that  they  explain 
the  entire  periodic  table. 


15-2     MANY-ELECTRON  ATOMS 


All  atoms  display  line  spectra.  In  general  these 
spectra  are  much  more  complicated  than  the 
atomic  hydrogen  spectrum  shown  in  Figure  15-3. 
Nevertheless,  these  spectra  can  be  interpreted  in 
terms  of  the  concepts  we  have  developed  for  the 
hydrogen  atom. 


15-2.1    Energy  Levels  of  Many-Electron  Atoms 

Analysis  of  the  spectra  of  many-electron  atoms 
shows  the  following  similarities  to  the  hydrogen 
atom  case. 


(1)  All  atoms  have  "stationary  states"  and  can 
hold  only  particular  values  of  energy. 

(2)  The  atomic  spectra  can  be  understood  in 
terms  of  transitions  between  energy  levels 
corresponding  to  these  particular  values  of 
energy. 

(3)  The  energy  level  diagrams  resemble  the  hy- 
drogen atom  level  diagram  except  that  the  n2 
levels  with  the  same  value  of  n  no  longer  all 
have  the  same  energy. 

Figure  15-11  shows  a  schematic  energy  level 
diagram  of  a  many-electron  atom.  Blue  patterns 


Arbitrary 

energy 

Scale 


65  ooo 

o 

5s 

o 


4s 

o 


3s 


2s 


o 


Is 


5P 


4p 


3P 


2p 


5d 


4d 


3d 


+f 


32 


18 


18 


Fig.  15-11.  A   schematic  energy  level   diagram   of  a   many-electron  atom. 


SEC.    15-3    |    IONIZATION    ENERGY    AND    THE    PERIODIC    TABLE 


267 


indicate  levels  with  the  same  hydrogen  atom 
principal  quantum  number.  The  effect  of  placing 
many  electrons  into  the  energy  levels  of  the  hy- 
drogen atom  is  a  "tilting"  of  the  diagram.  The/? 
orbitals  are  slightly  higher  in  energy  than  are  the 
s  orbitals  of  the  same  value  of  n.  The  d  orbitals 
and  /  orbitals  of  this  same  value  of  n  are  suc- 
cessively even  higher.  The  result  is  that  the  3d 
orbitals  are  raised  approximately  to  the  energy 
of  the  45  and  4p  orbitals.  The  35  and  3p  orbitals 
are  left  more  or  less  isolated  in  energy.  They  are 
much  higher  in  energy  than  are  the  25  and  2p 
orbitals  but  much  lower  than  is  the  cluster  of 
45,  4/7,  and  3d  orbitals. 


15-2.2    The  Periodic  Table 

Now  we  can  see  the  development  of  the  entire 
periodic  table.  The  special  stabilities  of  the  inert 
gases  are  fixed  by  the  large  energy  gaps  in  the 
energy  level  diagram,  Figure  15-11.  The  number 
of  orbitals  in  a  cluster,  multiplied  by  two  because 
of  our  double  occupancy  assumption,  fixes  the 
number  of  electrons  needed  to  reach  the  inert 
gas  electron  population.  The  numbers  at  the 


right  of  Figure  15-11  are  exactly  the  numbers  we 
found  as  differences  between  inert  gas  electron 
populations  (see  Table  15-1). 

It  is  important  to  recognize  what  we  have 
done.  We  have  shown  how  the  energy  level  dia- 
gram of  hydrogen  was  deduced  from  the  atomic 
hydrogen  spectra.  We  have  seen  that  the  energy 
level  diagram  of  a  many-electron  atom  can  be 
regarded  as  a  tilted  hydrogen  atom  diagram. 
Such  a  diagram  contains  clusters  of  energy  levels 
with  large  energy  gaps  between.  These  clusters  of 
energy  levels  provide  a  basis  for  explaining  the 
electron  populations  of  the  inert  gases,  provided 
we  assume  two  electrons  can  occupy  each  orbital. 

We  have  not  explained  why  two,  but  no  more 
than  two,  electrons  can  occupy  each  orbital.  This 
is  not  known  and  is  accepted  because  the  facts 
of  nature  require  it.  This  assumption  is  called 
the  Pauli  Principle. 

We  have  given  much  explanation  of  why  the 
clusters  and  energy  gaps  of  the  energy  level  dia- 
gram should  give  rise  to  a  periodicity  of  chemical 
properties  as  we  move  across  the  periodic  table. 
We  shall  seek  such  an  explanation  in  terms  of 
the  energy  necessary  to  remove  an  electron  from 
an  atom,  the  ionization  energy. 


15-3     IONIZATION  ENERGY  AND  THE  PERIODIC  TABLE 


The  amount  of  energy  required  to  remove  the  most 
loosely  bound  electron  from  a  gaseous  atom  is 
called  the  ionization  energy.  We  can  repre- 
sent this  process  by  the  equation 


gaseous  atom  +  energy 


gaseous  ion 

+  gaseous  electron 


In  terms  of  our  energy  level  diagram  it  is  the 
energy  necessary  to  lift  an  electron  from  the 
highest  occupied  orbital  the  rest  of  the  way  up  to 
the  limit  corresponding  ton  =  <».  Figure  15-12 
shows  the  ionization  process  for  lithium  atom  in 
terms  of  an  energy  level  diagram.  Each  orbital 
is  shown  as  a  pigeon  hole,  designated  O.  If  this 
orbital  is  occupied  by  one  electron,  it  can  be 
indicated  by  a  diagonal  line  across  the  pigeon 
hole:  0.  If  two  electrons  occupy  the  orbital, 
crossed  diagonal  lines  are  shown,  one  for  each 


electron:  <g>.  Of  course  we  assume  the  Pauli 
Principle  that  only  two  electrons  can  occupy  a 
given  orbital.  Hence,  an  orbital  shown  as  <g>  is 


Fig.  15-12 
o 


\ 


-500 


^     -1,000 

I 

,*    -isoo 


The  ionization  of  a  lithium  atom 

0tooo     /p 


J24  kcal 


Is 


OOO       2P 
2s 


Li  ($)  +  energy     — »■     Li*(g)  ■/■  e~(g) 


268 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.    15 


filled;  an  orbital  shown  as  Q  or  0  is  half- 
occupied  ;  an  orbital  shown  as  O  is  empty. 

The  term  ionization  energy  is  also  applied  to 
the  removal  of  the  most  loosely  bound  electron 
from  an  atom  that  has  already  lost  one  or  more 
electrons  (that  is,  an  ion).  Hence  the  "ionization 
energy"  of,  say,  Mg+  is  the  energy  of  the  process 

Mg+  (g)  -f  energy  -*  Mg+2  (g)  +  e~(g) 

Since  the  process  removes  the  second  electron 
from  a  magnesium  atom,  the  ionization  energy 
of  Nig"1"  is  called  the  "second  ionization  energy" 
of  magnesium. 

15-3.1    Measurement  of  Ionization  Energy 

The  ionization  energy  provides  a  basis  for  under- 
standing the  periodicity  of  the  chemistry  of  the 
elements.  Owing  to  the  stimulation  of  Bohr's 
ideas,  many  systematic  determinations  of  ioniza- 
tion energies  were  carried  out  between  1914  and 
1920.  The  first  determinations  were  done  by 
bombarding  an  atomic  vapor  with  electrons 
whose  kinetic  energy  was  known  accurately. 
When  the  kinetic  energy  of  the  bombarding  elec- 
trons is  increased  to  a  certain  critical  value, 
singly  charged  positive  ions  can  be  detected  elec- 
trically. These  ions  result  from  collisions  between 
atoms  and  the  bombarding  electrons  that  have 
been  given  just  enough  kinetic  energy  to  cause 
the  most  weakly  bound  electron  to  be  ejected 
from  the  atom.  This  critical  value  was  found  to 
be  characteristic  of  the  substance  being  investi- 
gated. Table  15-111  shows  some  of  the  measured 
ionization  energies  for  some  of  the  lighter  ele- 
ments.* 


Table  75-/// 

IONIZATION    ENERGIES 
OF    THE    ELEMENTS 


ATOMIC 
NUMBER 


ELEMENT 


IONIZATION  ENERGY 

(kcal/mole) 


1 

H 

313.6 

2 

He 

566.7 

3 

Li 

124.3 

4 

Be 

214.9 

5 

B 

191.2 

6 

C 

259.5 

7 

N 

335 

8 

O 

313.8 

9 

F 

401.5 

10 

Ne 

497.0 

11 

Na 

118.4 

12 

Mg 

175.2 

13 

Al 

137.9 

14 

Si 

187.9 

15 

P 

241.7 

16 

S 

238.8 

17 

CI 

300 

18 

Ar 

363.2 

19 

K 

100.0 

following  it.  This  is  followed  by  a  slow  rise  in  the 
ionization  energy  (another  regularity)  as  we 
proceed  across  a  row  of  the  periodic  table.  The 
best  way  to  see  the  trend  is  in  a  graphical  pre- 
sentation of  the  data  of  Table  15-111,  as  shown 
in  Figure  15-13.  We  see  that  the  ionization  en- 


Fig.  15-13.  Ionization   energy   as   a   function   of   the 
atomic  number. 


15-3.2    Trends  in  Ionization  Energies 

Examine  the  figures  in  the  last  column  of  Table 
15-111  and  search  for  regularities.  The  most  ob- 
vious one  is  the  dramatic  change  in  ionization 
energy  between  each  inert  gas  and  the  element 

*  The  ionization  energy  is  given  in  this  book  in  units 
of  kilocalories  per  mole,  the  energy  that  would  be  re- 
quired to  remove  an  electron  from  each  one  of  a  mole  of 
atoms.  These  units  allow  an  easy  comparison  between 
ionization  energies  and  the  energy  changes  that  occur  in 
ordinary  chemical  reactions. 


5 

3 


600 


4-00 


ZOO  - 


2     4 


6     8    10   1Z    14   16   16  20  ZZ 
Atomic   number 


SEC.     15-3    |    IONIZATION    ENERGY    AND    THE    PERIODIC    TABLE 


269 


ergy  increases  more  or  less  regularly  across  a 
row  of  the  periodic  table,  reaching  a  maximum 
at  the  inert  gas.  As  soon  as  we  encounter  an 
alkali  metal,  we  notice  that  the  ionization  energy 
is  quite  small,  and  in  subsequent  elements  the 
general  upward  trend  repeats  itself.  There  is  a 
startling  similarity  between  the  regularities  we 
have  found  in  the  ionization  energies  and  the 
periodicity  of  chemical  properties.  We  shall  see 
that  this  is  not  an  accident:  the  trend  in  chemical 
behavior  as  we  move  across  the  periodic  table 
can  be  explained  in  terms  of  the  trends  in  the 
ionization  energies. 

Let  us  begin  by  contrasting  the  ionization 
properties  of  sodium  and  chlorine: 

Nate;  — *•  Na+teJ  +  e~(g) 

AH  =  +118.4  kcal/mole 
Cite;   —*■  C\+(g)  +  e-(g) 

AH  =  +  300  kcal/mole 

Since  sodium  has  a  low  ionization  energy,  1 18.4 
kcal/mole,  a  relatively  small  amount  of  energy 
is  required  to  remove  an  electron.  This  is  con- 
sistent with  the  chemical  evidence  that  sodium 
tends  to  form  compounds  involving  the  ion,  Na+. 
The  ease  of  forming  Na+  ions  can  be  explained 
in  terms  of  the  low  ionization  energy  of  the 
sodium  atom.  In  contrast,  it  requires  a  large 
amount  of  energy,  300  kcal/mole,  to  remove  an 
electron  from  a  chlorine  atom.  It  is  not  surpris- 
ing, then,  that  this  element  shows  little  tendency 
to  lose  electrons  in  chemical  reactions.  Instead, 
chlorine  commonly  acquires  electrons  to  form 
negative  ions,  CI-. 

Between  sodium  and  chlorine,  there  is  a  slow 
rise  in  ionization  energy.  For  magnesium  and 
aluminum  the  ionization  energy  is  still  rather 
low.  Hence  electrons  are  readily  lost  and  positive 
ions  can  be  expected  to  be  important  in  the 


chemistry  of  these  elements.  As  the  ionization 
energy  rises,  the  chemistries  of  silicon,  phos- 
phorus, and  sulfur  show  a  trend  toward  electron 
sharing.  For  these  elements,  an  inert  gas  electron 
population  cannot  be  reached  by  losing  electrons 
because  the  ionization  energy  required  is  too 
high.  They  seek  the  inert  gas  stability  by  sharing 
electrons  or,  for  sulfur  and  chlorine,  by  acquiring 
electrons  to  form  negative  ions. 

These  correlations  between  ionization  energy 
and  chemical  properties  confirm  the  idea  that 
the  electronic  structure  of  an  element  determines 
its  chemical  behavior.  In  particular,  the  most 
weakly  bound  electrons  are  of  greatest  im- 
portance in  this  respect.  We  shall  call  the 
electrons  that  are  most  loosely  bound,  the  valence 
electrons. 

15-3.3    Ionization  Energies  and 
Valence  Electrons 

It  is  possible  to  remove  two  or  more  electrons 
from  a  many-electron  atom.  Of  course  it  is 
always  harder  to  remove  the  second  electron 
than  the  first  because  the  second  electron  to  come 
off  leaves  an  ion  with  a  double  positive  charge 
instead  of  a  single  positive  charge.  This  gives  an 
additional  electrical  attraction.  Even  so,  the 
values  of  successive  ionization  energies  have 
great  interest  to  the  chemist. 

Consider  the  three  elements,  sodium,  magne- 
sium, and  aluminum.  For  each  of  these  elements 
we  know  several  ionization  energies,  correspond- 
ing to  processes  such  as  the  following: 


NateJ 
Na+teJ 


Na+teJ  +  e-(g) 

Ei  =  first  ionization  energy 


(5) 


Na+Ygj  +  e-(g) 

E2  =  second  ionization  energy    (6) 


Table  15-IV.     successive  ionization  energies  of  Na,  ms.  and  ai 

(KILOCALORIES  PER  MOLE) 


ELEMENT 


ELECTRON  CONFIGURATION, 
NEUTRAL  ATOM 


Et 


Et 


sodium 

Is* 

2s>2? 

3sl 

118 

1091 

1653 

— 

magnesium 

Is1 

2y*2/7« 

3j» 

175 

345 

1838 

2526 

aluminum 

Is* 

2$lp 

3^3/7' 

138 

434 

656 

2767 

270 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


Na+3(gJ 


Na+3fg;  +  e-(g) 

E3  =  third  ionization  energy 


(7) 


Na+"(gj  +  e-(g) 

Ei  =  fourth  ionization  energy      8) 


The  experimental  values  of  these  energies  are 
shown  in  Table  15-IV.  Let  us  begin  by  comparing 
sodium  and  magnesium.  For  each,  the  first 
ionization  process  removes  a  3s  electron,  the 
most  weakly  bound.  Nevertheless,  the  ionization 
energies  are  somewhat  different: 

Na(g J  +H8  kcal  — +  Na+(g)  +  er(g)     (9) 
Mg(g)  +  175  kcal  — *-  Mg+(g)  +  e~(g)    (10) 

The  difference  is  caused  by  the  higher  nuclear 
charge  of  magnesium.  Magnesium  is  element  12, 
hence  it  has  twelve  protons  in  the  nucleus,  com- 


pared to  eleven  protons  in  the  nucleus  of  the 
sodium  atom.  Of  course  the  valence  electrons  are 
more  strongly  attracted  to  the  +12  nucleus  of 
Mg  than  the  +11  nucleus  of  Na. 

The  second  ionization  energy,  however,  re- 
verses the  situation: 


Na+fgj  +  1091  kcal 
Mg+(g)  +    345  kcal 


Na+Ygj  +  e~(g) 
Mg-W  +  e-(g) 


an 

(12) 


For  sodium,  it  takes  three  times  as  much  energy 
to  remove  the  second  electron  as  it  does  for 
magnesium.  We  can  understand  the  energies  in 


Fig.  15-14.  Energy    level   diagram   representation   of 

the  ionization  of  magnesium  and  sodium 
atoms. 


-1,000 


-2,000±r 


t  OOO         3* 


175  kcal 


2j» 

2s 


OOO 


345 kcal 


<> 


O 


OOO 


1838  kcal 


o     OOO 


T® 


s 


T®  /*  T® 

Mg(9)  +  175 kcal— Mg+(9)  +  e~(g)  MS+2(g)  +  1838  kcal~Mo+3(9)  + e~(S) 

Aig+(g)  +  345  kcal 2*fg+2(9)  +  e~(9) 


lOOO     33p 


118  kcal 
m      ®®®       ±Zf 


OOO 


1091  kcal 


®®0 


^   -1,000 

\ 


-2,000}? 

J®  Is  T< 

Na(g)  +  118  kcal—Na+(g)  +   e~ (g) 

Na+ (g)  +  1091  kcal -Na.+Z(g)  +    e~  (g) 


OOO 


SEC.     15-3    |    IONIZATION    ENERGY    AND    THE    PERIODIC    TABLE 


271 


(77)  and  (72)  in  terms  of  two  principles.  First, 
for  magnesium  the  second  ionization  energy  (345 
kcal)  exceeds  the  first  (175  kcal)  because  reaction 
(72)  takes  a  3s  electron  away  from  the  positively 
charged  ion,  Mg+,  whereas  reaction  (70)  removes 
a  3s  electron  from  an  initially  neutral  atom.  This 
same  factor  is  operative  in  sodium,  of  course, 
but  in  addition,  the  second  ionization  energy  of 
sodium  must  remove  an  electron  from  a  2p 
orbital  instead  of  the  35  orbital.  This  2p  orbital 
is  one  of  a  cluster  much  lower  in  energy  than 
the  35-3/j  cluster.  Figure  15-14  shows  that  this 
behavior  is  readily  understood  with  the  aid  of 
the  energy  level  diagram.  We  conclude  that  it  is 
difficult  to  remove  two  electrons  from  sodium 
but  it  is  relatively  easy  to  remove  two  electrons 
from  magnesium.  This  is  why  we  say  sodium  has 
one  valence  electron  and  magnesium  has  two. 
Removing  more  than  one  electron  from  sodium 
or  more  than  two  from  magnesium  is  very  diffi- 
cult because  orbitals  much  lower  in  the  energy 
diagram  are  involved. 

Continuing  to  aluminum,  we  see  that  its  first 
ionization  energy  is  below  the  first  ionization 
energy  of  magnesium — despite  the  fact  that  alu- 
minum has  the  higher  nuclear  charge.  We  find 
the  explanation,  however,  in  the  second  column 
of  Table  15-1V.  For  aluminum  we  remove  a  3p 
electron  to  produce  Al+  whereas  a  35  electron 
is  removed  from  Mg  to  form  Mg+.  A  glance  at 
the  energy  level  diagram,  Figure  15-11,  shows 
that  the  3>p  level  is  somewhat  above  the  35  level, 
hence  it  should  be  easier  to  remove  the  3p 
electron. 

If  we  continue  to  remove  electrons  from  alu- 
minum, we  discover  a  very  large  increase  in 
ionization  energy  when  the  fourth  electron  is  re- 
moved. Again  this  is  because  the  fourth  electron 
must  be  withdrawn  from  a  2p  orbital,  an  orbital 
much  lower  on  the  energy  level  diagram.  We 
conclude  that  three  electrons,  the  two  35  and  the 
one  3p,  are  more  easily  removed  than  the  others. 
Since  aluminum  has  three  easily  removed  elec- 
trons, aluminum  is  said  to  have  three  valence 
electrons. 

Remembering  how  we  placed  electrons  in  the 
lowest  empty  orbitals,  two  per  orbital,  we  can 
now  generalize  concerning  the  number  of  valence 


electrons  a  given  atom  possesses.  We  count  the 
electrons  placed  in  the  orbitals  that  form  the 
highest  partially  filled  cluster  of  energy  levels. 
These  electrons  are  most  easily  removed  and 
they  determine  the  chemistry  of  the  atom. 


EXERCISE  15-6 

Explain  why  chemists  say  that  boron  has  three 
valence  electrons  and  that  chlorine  has  seven. 
How  many  valence  electrons  has  fluorine?  Oxy- 
gen? Nitrogen? 


15-3.4    The  Fourth  Row  of  the  Periodic  Table 

Turn  back  to  Figure  15-1 1,  the  energy  level  dia- 
gram of  a  many-electron  atom,  and  consider  the 
occupied  orbitals  of  the  element  potassium.  With 
19  electrons  placed,  two  at  a  time,  in  the  orbitals 
of  lowest  energy,  the  electron  configuration  is 

potassium  atom         Is2      2s*2p6      3s*3p6      4s1 

Potassium  has  one  valence  electron.  It  is  the  first 
member  of  the  fourth  row,  the  row  based  on  the 
cluster  of  orbitals  with  about  the  same  energy  as 
the  45  orbital.  There  are  nine  such  orbitals,  the 
45  orbital,  the  three  Ap  orbitals,  and  the  five  3d 
orbitals.  Hence  the  fourth  row  of  the  periodic 
table  will  differ  from  the  second  and  third  rows. 
The  fourth  row,  as  seen  in  the  periodic  table, 
consists  of  eighteen  elements. 

Calcium  is  the  second  element  of  the  fourth 
row.  It  has  two  electrons  more  than  the  argon 
inert  gas  population  and  these  two  electrons  both 
occupy  the  45  orbital: 

calcium  atom  Is2       2522p6       3523p6       4s2 

When  we  add  the  next  electron  to  form  the 
element  scandium,  the  orbital  of  lowest  energy 
that  is  available  is  one  of  the  3d  orbitals  (since 
the  3d  orbitals  are  slightly  lower  in  energy  than 
the  4p  orbitals).  As  succeeding  electrons  are 
added  to  form  other  elements,  they  enter  the  3d 
orbitals  until  the  ten  available  spaces  in  these 
orbitals  are  filled. 

The  elements  that  are  formed  when  the  3d  elec- 
trons are  added  are  called  the  transition  metals 


272 


ELECTRONS    AND    THE    PERIODIC    TABLE    I    CHAP.     15 


or  the  transition  elements.  Since  their  chemi- 
cal properties  and  electronic  structures  are  unlike 
those  of  any  of  the  lighter  elements  we  have 
discussed,  it  is  reasonable  that  these  transition 
elements  should  head  a  new  set  of  ten  columns 
of  the  periodic  table.  The  next  orbitals  in  line 
for  occupancy  are  the  4/7,  and  it  is  not  surprising 
that  the  chemical  properties  of  the  element  gal- 
lium, which  has  one  4/?  electron,  resemble  those 
of  aluminum,  which  has  one  3p  electron.  This 
row  of  the  periodic  table  is  completed  when  the 
4/7  orbitals  are  entirely  filled.  We  see  that  the 
reason  the  fourth  row  of  the  periodic  table  con- 
tains eighteen  elements  is  that  the  five  3d  orbitals 
have  energies  that  are  approximately  the  same 
as  those  of  the  45  and  4/7  orbitals.  The  ten  extra 
spaces  for  electrons  provided  by  the  3d  orbitals 
increase  the  length  of  the  period  from  eight  to 
eighteen. 

Just  as  the  long  fourth  row  of  the  periodic  table  arises 
from  filling  the  4s,  3d,  and  4p  orbitals,  the  fifth  row,  which 
also  consists  of  eighteen  elements,  comes  from  filling  the 
5s,  4d,  and  5p  orbitals.  In  the  sixth  row,  something  new 
happens.  After  the  6s  and  the  first  one  of  the  5d  electrons 
have  entered,  subsequent  electrons  go  into  the  4)  orbitals. 
The  fact  that  there  are  seven  4/  orbitals  means  that 
fourteen  electrons  can  be  accommodated  in  this  manner. 
Filling  the  4/orbitals  gives  rise  to  a  series  of  elements  with 
almost  identical  chemical  properties  called  the  rare  earth 


elements.  Once  the  4/orbitals  are  full,  electrons  enter  the 
5dand  6p  orbitals  until  the  sixth  period  is  completed  with 
the  inert  gas  radon. 

We  see  that  the  rows  of  the  periodic  table  arise 
from  filling  orbitals  of  approximately  the  same 
energy.  When  all  orbitals  of  similar  energy  are 
full  (two  electrons  per  orbital),  the  next  electron 
must  be  placed  in  an  s  orbital  that  has  a  higher 
principal  quantum  number,  and  a  new  period  of 
the  table  starts.  We  can  summarize  the  relation 
between  the  number  of  elements  in  each  row  of 
the  periodic  table  and  the  available  orbitals  of 
approximately  equal  energy  in  Table  15-V. 


Table  15-V 

THE    NUMBER    OF    ELEMENTS    IN    EACH 
ROW    OF    THE    PERIODIC    TABLE 


ROW  OF 

NO.  OF 

LOWEST  ENERGY  ORBITALS 

TABLE 

ELEMENTS 

AVAILABLE  TO  BE  FILLED 

1 

2 

Is 

2 

8 

2s,  2p 

3 

8 

3s,  3p 

4 

18 

4s,  3d,  4p 

5 

18 

5s,  4d,  5/7 

6 

32 

6s,  4/,  5d,  6/7 

7 

Is,  5f,  6d,  Ip 

QUESTIONS  AND  PROBLEMS 


1.  Which  of  the  following  statements  concerning 
light  is  FALSE? 

(a)  It  is  a  form  of  energy. 

(b)  All  photons  possess  the  same  amount  of 
energy. 

(c)  It  cannot  be  bent  by  a  magnet. 

(d)  It  includes  the  part  of  the  spectrum  called 
X-rays. 

2.  Use  the  energy  level  diacram  in  Figure  15-7  to 
calculate  the  energy  required  to  raise  the  electron 
in  a  hydrogen  atom  from  level  #1  to  level  #2; 
from  level  #1  to  level  #3;  from  level  #1  to  level 
#4.  Compare  these  energies  with  the  spectral 
lines  shown  in  Figure  15-3,  p.  255. 


Your  plot  in  Exercise  15-2  suggested  that  the 
energy  levels  given  in  Figure  15-7  are  systemati- 
cally related.  To  explore  this  relationship  further, 
divide  the  energy  of  each  level  by  that  of  the 
first  level  (using  the  right-hand  scale).  How  are 
the  fractions  so  obtained  related  to  the  numbers 
of  the  energy  levels? 

Calculate,  using  frequency  units  (cycles  per  sec- 
ond) and  Figure  15-3,  the  lines  predicted  by  the 
notched  beam  due  to  changes  beginning  in  notch 
§7>.  Use  the  complete  light  spectrum  shown  in 
Figure  14-14  (p.  247)  to  decide  in  what  spectral 
region  these  additional  lines  were  found.  (It  was 
this  sort  of  prediction  that  actually  led  to  the 
discovery  of  this  set  of  lines.) 


QUESTIONS    AND    PROBLEMS 


273 


5.  According  to  the  quantum  mechanical  descrip- 
tion of  the  \s  orbital  of  the  hydrogen  atom,  what 
relation  exists  between  the  surface  of  a  sphere 
centered  about  the  nucleus  and  the  location  of 
an  electron  ? 

6.  What  must  be  done  to  a  25  electron  to  make  it  a 
3s  electron?  What  happens  when  a  35  electron 
becomes  a  25  electron? 

7.  If  the  energy  difference  between  two  electronic 
states  is  46.12  kcal/mole,  what  will  be  the  fre- 
quency of  light  emitted  when  the  electron  drops 
from  the  higher  to  the  lower  state? 

Planck's  constant 

=  9.52  X  10~14  (kcal  sec)/mole 

8.  Determine  the  value  of  En  for  n  —  1,  2,  3, 
4,  for  a  hydrogen  atom  using  the  relation 
En  =  —  313.6/h2.  For  each  En,  indicate  how 
many  orbitals  have  this  energy. 

9.  The  quantum  mechanical  description  of  the  Is 
orbital  is  similar  in  many  respects  to  a  descrip- 
tion of  the  holes  in  a  much  used  dartboard.  For 
example,  the  "density"  of  dart  holes  is  constant 
anywhere  on  a  circle  centered  about  the  bullseye, 
and  the  "density"  of  dartholes  reaches  zero  only 
at  a  very  long  distance  from  the  bullseye  (effec- 
tively, at  infinity).  What  are  the  corresponding 
properties  of  a  15  orbital? 

In  the  light  of  your  answer,  point  out  errone- 
ous features  of  the  following  models  of  a  hydro- 
gen atom  (both  of  which  were  used  before 
quantum  mechanics  demonstrated  their  inade- 
quacies). 

(a)  A  ball  of  uniform  density. 

(b)  A  "solar  system"  atom  with  the  electron 
circling  the  nucleus  at  a  fixed  distance. 

10.  Name  the  elements  that  correspond  to  each  of 
the  following  electron  configurations 

l52 

l52  251 

l52  2522p1 

I52        2s22p3 

\s2        2522p6        3523p6        4a1 


1 1 .  Make  a  table  listing  the  principal  quantum  num- 
bers (through  three),  the  types  of  orbitals,  and 
the  number  of  orbitals  of  each  type. 

12.  The  electron  configuration  for  lithium  is  l5225' 
and  for  beryllium  it  is  l52252.  Estimate  the  ap- 
proximate ionization  energies  to  remove  first 
one,  then  a  second,  electron.  Explain  your  esti- 
mates. 

13.  What  trend  is  observed  in  the  first  ionization 
energy  as  you  move  from  lithium  down  the  col- 
umn I  metals?  On  this  basis,  can  you  suggest  a 
reason  why  potassium  or  cesium  might  be  used 
in  preference  to  sodium  or  lithium  in  photo- 
electric cells? 

14.  Consider  these  two  electron  populations  for 
neutral  atoms: 


A.  I52      2522/?6 

B.  I52      2522p6 


351; 
6sl. 


Which  of  the  following  is  FALSE? 

(a)  Energy  is  required  to  change  A  to  B. 

(b)  A  represents  a  sodium  atom. 

(c)  A  and  B  represent  different  elements. 

(d)  Less  energy  is  required  to  remove  one  elec- 
tron from  B  than  from  A. 

15.  How  many  valence  electrons  has  carbon?  Sili- 
con? Phosphorus?  Hydrogen?  Write  the  elec- 
tron configurations  for  neutral  atoms  of  each 
element. 

16.  The  first  four  ionization  energies  of  boron  atoms 
are  as  follows: 

Ex  =  191  kcal/mole 
£2  =  578 
£3  =  872 
£4  -  5962 

Explain  the  magnitudes  in  terms  of  the  electron 
configurations  of  boron  and  deduce  the  number 
of  valence  electrons  of  boron. 


CHAPTER 


16 


Molecules  in  the 
Gas  Phase 


For  the  nature  of  the  chemical  bond  is  the  problem  at  the  heart  of  all 
chemistry. 

BRYCE     CRAWFORD,     JR.,     1953 


A  molecule  is  a  cluster  of  atoms  that  persists 
long  enough  to  have  characteristic  properties 
which  identify  it.  The  questions  we  would  like 
to  answer  are  "Why  does  the  cluster  of  atoms 
persist?"  and  "Why  does  the  clustering  result  in 
characteristic  properties?"  In  this  chapter  we 


will  restrict  our  attention  to  molecules  as  they 
exist  in  the  gas  phase.  Then,  in  Chapter  17,  we 
shall  consider  what  additional  ideas  we  need  in 
order  to  understand  the  forces  which  cause  the 
formation  of  liquids  and  solids. 


16-1    THE  COVALENT  BOND 

When  two  atoms  become  fixed  together  they  are 
said  to  form  a  molecule.  In  "explaining"  the 
properties  of  a  molecule  we  use  models  such  as 
two  styrofoam  spheres  glued  together,  or  two 
wooden  spheres  held  together  by  a  stick,  or 
perhaps  by  a  spring.  In  each  model  it  is  necessary 
to  provide  a  connection — glue,  stick,  or  spring. 
It  is  natural  to  assume  that  there  is  a  connection 
between  the  atoms  in  a  molecule.  This  connec- 
tion is  called  the  chemical  bond. 

16-1.1    The  Hydrogen  Molecule 

Under  normal  conditions  of  temperature  and 
pressure,  hydrogen  is  a  gas.  By  weighing  a  meas- 
274 


ured  volume  of  hydrogen  gas  and  applying 
Avogadro's  Hypothesis,  we  discover  that  a  mole- 
cule of  hydrogen  contains  two  hydrogen  atoms. 
Only  if  the  temperature  is  raised  to  several  thou- 
sand degrees  are  the  collisions  with  other  mole- 
cules sufficiently  energetic  to  knock  a  hydrogen 
molecule  apart: 

H2(g)  +±:  U(g)  +  U(g)       AH  =  103.4  kcal    (/) 

Since  energy  is  absorbed  in  reaction  (7),  the 
molecule  H2  is  more  stable  (has  a  lower  energy) 
than  two  separated  atoms.  This  chemical  bond 
(and  every  chemical  bond)  forms  because  the 
energy  is  lower  when  the  atoms  are  near  each 
other. 


SEC.     16-1    I    THE    COVALENT    BOND 


275 


e 


© 


+ 


e 


<& 


e 


'lA 


® 


ti 


+ 


d5 


Fig.  16-1.  The  formation  of  a  molecule  of  hydrogen, 
Hs. 


THE   ORIGIN    OF   THE   STABILITY 
OF   THE    CHEMICAL   BOND 

To  see  why  the  energy  is  lower  when  the  atoms 
are  near  each  other,  we  must  examine  interac- 
tions among  the  electric  charges  of  the  atoms. 
Figure  16-1  shows  the  reverse  of  reaction  (7)  in 
a  schematic  way.  Quantum  mechanics  tells  us 
that  the  Is  orbital  of  each  hydrogen  atom  has 
spherical  symmetry  before  reaction.  This  is  sug- 
gested by  the  shading  in  Figure  16-1.  Yet,  at  any 
instant,  we  picture  the  electron  at  some  particu- 
lar point,  as  shown  by  the  negative  charge  of 
electron  1  located  a  distance  riA  from  nucleus  A. 
The  energy  of  hydrogen  atom  A  can  be  explained 


1A\ 


in  terms  of  the  average  attraction  between  elec- 
tron 1  and  nucleus  A.  This  is  fixed  by  the  average 
of  the  distance  between  the  two,  riA.  The  same 
is  true  of  hydrogen  atom  B — electron  2  and 
nucleus  B  attract  each  other.  Now  consider  the 
new  electrical  interactions  present  after  the  two 
atoms  have  moved  close  together.  Now  electron 
1  feels  the  attraction  of  both  protons.  Electron  2, 
as  well,  feels  the  attraction  of  both  protons.  This 
is  the  "glue"  that  holds  the  two  atoms  together. 
The  chemical  bond  in  H2  forms  because  each  oj 
the  two  electrons  is  attracted  to  two  protons 
simultaneously.  This  arrangement  is  energetically 
more  stable  than  the  separated  atoms  in  which 


Fig.  16-2.  Attractive  and  repulsive  forces  in  the  hy- 
drogen molecule. 


Electron- proton,  distances: 
attractive  forces 


R 


Electron.- electron   and 

proton  -proton   distances: 

repulsive  forces 


b 


■■J 


> 


■^ 


-o- 


New 


®- 


5 


276 


MOLECULES    IN    THE    GAS    PHASE    I    CHAP.     16 


each  electron  is  attracted  to  only  one  proton. 
Figure  16-2 A  shows  a  possible  set  of  the 
electron-proton  distances  as  they  might  be  seen 
if  it  were  possible  to  make  an  instantaneous 
photograph.  Such  distances  fix  the  attractions 
that  cause  the  chemical  bond.  But  it  is  well  to 
remember  there  are  also  repulsions  caused  by 
the  approach  of  the  two  atoms,  as  shown  in 
Figure  16-2B.  The  two  electrons  repel  each  other 
and  the  two  protons  do  the  same.  These  repul- 
sions tend  to  push  the  two  atoms  apart.  Which 
are  more  important,  the  two  new  attraction 


terms  or  the  two  new  repulsion  terms  illustrated 
in  Figure  16-2?  Experiment  shows  that  the  new 
attraction  terms  dominate — a  stable  chemical 
bond  is  formed.  That  is  not  to  say  the  repulsions 
are  not  felt.  In  fact,  the  proton-proton  repulsions 
prevent  the  two  hydrogen  atoms  from  approach- 
ing even  closer.  The  stable  bond  length  in  the 
hydrogen  molecule  is  fixed  by  a  balance  between 


Fig.  16-3.  Schematic  representation  of  the  interaction 
between  two  atoms. 


y\..     A  simplified    representation   of  a    Is   orbi-ta.1: 


wilt    be    shown     as 


© 


with    the   addition 

of  shading    to 
indicate    occupancy 


B.      A   simplified    representation    of    orbifral    occu.pa.ncy. 


0 


Is    orbital 
emp<ty 


Is    orbital 

containing 

one    electron  ■ 

"half  filled  " 


Is    orbital 
con  -ta  in  ing 
■two    ele  c  tro  n  s  ■ 
"filled  " 


C.       Overlap   and    bonding    of  the    hydrogen     molecule : 


+ 


> 


H 


+ 


H 


H, 


D.      Absence  of  overlap  for    two    helium    atoms . 


+ 


<=] 


He 


He 


2He 


SEC.     16-1    I    THE    COVALENT    BOND 


277 


Electron. -pro-ton  distances •' 
attractive  forces 


rV"'S  ^4* 


the  forces  of  attraction  (Figure  16-2 A)  and  the 
forces  of  repulsion  (Figure  16-2B). 

We  see  in  Figure  16-2  that  bringing  two  hydrogen 
atoms  together  produces  two  new  repulsions  and  two  new 
attractions.  Experiment  shows  that  a  chemical  bond  is 
formed ;  the  energy  of  attraction  predominates  over  the 
energy  of  repulsion.  Why  is  this  so?  An  explanation  is 
found  in  the  mobility  of  the  electrons.  The  electrons  do 
not  occupy  fixed  positions  but  move  about  through  the 
molecule.  They  take  advantage  of  this  mobility  to  remain 
as  far  from  each  other  as  possible.  They  preferentially 
occupy  positions  like  those  shown  in  Figure  16-1  in  which 
each  electron  is  closer  to  both  nuclei  than  to  the  other 
electron.  The  two  electrons  preferentially  move  away 
from  positions  in  which  they  would  be  near  each  other. 
Thus  they  are  said  to  "correlate"  their  motion  so  as  to 
remain  apart,  reducing  the  electron-electron  repulsion. 

OVERLAP    AND   THE   CHEMICAL   BOND 

We  can  simplify  our  discussion  of  the  chemical 
bonding  in  the  hydrogen  molecule  with  the  aid 
of  Figure  16-3.  First,  in  Figure  16-3  A,  we  picture 
the  electron  distribution  in  cross-section.  The 
electron  distribution  extends  outward  far  from 
the  nucleus  and  uniformly  in  all  directions.  But 
the  distribution  is  concentrated  near  the  nucleus 
so  we  ought  to  focus  our  attention  on  the  center 
region  of  the  Is  orbital.  We  do  so  by  representing 
the  Is  orbital  by  a  circle  with  radius  large  enough 
to  contain  most  of  the  electron  distribution. 

An  orbital  can  accommodate  either  one  or  two 
electrons  but  no  more.  Figure  16-3B  shows  a  way 
of  differentiating  an  empty  \s  orbital,  a  Is  or- 
bital containing  one  electron,  and  a  Is  orbital 
containing  two  electrons. 

Now,  in  Figure  16-3C,  consider  the  interaction 


Electron.-  electron   and 
proton-proton    distances ■' 
repulsive    forces 


4r. 


,-■4- 


\ 

New    / 


!      ^    ''     ' 


jV«w 


Fig.  16-4.  Attractive  and  repulsive  forces  when  helium 
atoms  approach. 


of  two  hydrogen  atoms.  Each  atom  has  a  single 
electron  in  a  b  orbital.  As  the  two  hydrogen 
atoms  approach  each  other,  the  circles  overlap 
each  other.  In  this  region  of  overlap  the  two 
electrons  are  shared  by  the  two  protons  (as 
shown  by  the  crosshatched  and  shaded  area). 
This  sharing,  which  permits  the  two  electrons  to 
be  near  both  protons  a  good  part  of  the  time, 
causes  the  chemical  bond.  When  a  bond  arises 
from  equal  sharing,  it  is  called  a  covalent  bond.* 

16-1.2    Interaction  Between  Helium  Atoms 

A  measurement  of  the  density  of  helium  gas 
shows  that  it  is  a  monatomic  gas.  Molecules  of 
He2  do  not  form.  What  difference  between  hy- 
drogen atoms  and  helium  atoms  accounts  for  the 
absence  of  bonding  for  helium?  The  answer  to 
this  question  also  must  lie  in  the  attractive  and 
repulsive  electrical  interactions  between  two  he- 
lium atoms  when  they  approach  each  other. 
Figure  16-4 A  shows  the  attractive  forces  in  one 
of  our  hypothetical  instantaneous  snapshots. 
There  are,  of  course,  four  electrons  and  each  is 
attracted  to  each  nucleus.  In  Figure  16-4B  we 
see  the  repulsive  forces.  Taking  score,  we  find 
in  Figure  16-4A  eight  attractive  interactions,  four 


*  The  prefix  "co"  in  "covalent"  conveys  the  notion  of 
"sharing"  as  it  does  in  the  words  "coworker,"  "co- 
author," etc.  The  stem  of  the  word,  "-valent"  refers  to 
"combination." 


278 


MOLECULES    IN    THE    GAS    PHASE   I    CHAP.     16 


more  than  in  the  separated  atoms.  In  addition, 
we  count  seven  repulsive  interactions,  five  more 
than  in  the  separated  atoms.  Again  we  appeal  to 
experiment  and  we  learn  that  the  four  new  at- 
tractive terms  are  not  sufficient  to  outweigh  the 
five  new  repulsions.  A  chemical  bond  does  not 
form. 

Thus  we  find  that  an  explanation  of  the  bond- 
ing in  H2  and  the  absence  of  bonding  for  He2  lies 
in  the  relative  magnitudes  of  attractive  and  re- 
pulsive terms.  Quantum  mechanics  can  be  put  to 
work  with  the  aid  of  advanced  and  difficult 
mathematics  to  calculate  these  quantities,  to  tell 
us  which  is  more  important.  Unfortunately, 
solving  the  mathematics  presents  such  an  ob- 
stacle that  only  a  handful  of  the  very  simplest 
molecules  have  been  treated  with  high  accuracy. 
Nevertheless,  for  some  time  now  chemists  have 
been  able  to  decide  whether  chemical  bonds  can 
form  without  appealing  to  a  digital  computer. 

Figure  16-3D  shows  the  simplified  representa- 
tion of  the  interaction  of  two  helium  atoms.  This 
time  each  helium  atom  is  crosshatched  before  the 
two  atoms  approach.  This  is  to  indicate  there  are 
already  two  electrons  in  the  Is  orbital.  Our  rule 
of  orbital  occupancy  tells  us  that  the  Is  orbital 
can  contain  only  two  electrons.  Consequently, 
when  the  second  helium  atom  approaches,  its 
valence  orbitals  cannot  overlap  significantly.  The 
helium  atom  valence  electrons  fill  its  valence 
orbitals,  preventing  it  from  approaching  a  sec- 
ond atom  close  enough  to  share  electrons.  The 
helium  atom  forms  no  chemical  bonds.* 

16-1.3    Representations  of  Chemical  Bonding 

We  propose,  then,  that  chemical  bonds  can  form 
if  valence  electrons  can  be  shared  by  two  atoms 
using  partially  filled  orbitals.  We  need  a  short- 
hand notation  which  aids  in  the  use  of  this  rule. 
Such  a  shorthand  notation  is  called  a  represen- 
tation of  the  bonding. 

*  Each  helium  atom  does  have,  of  course,  vacant  2s 
and  2p  orbitals  which  extend  farther  out  than  the  filled 
Is  orbital.  The  electrons  of  the  second  helium  atom  can 
"overlap"  with  these  vacant  orbitals.  Since  this  overlap 
is  at  great  distance,  the  resulting  attractions  are  extremely 
small.  This  type  of  interaction  presumably  accounts  for 
the  attractions  that  cause  helium  to  condense  at  very  low 
temperatures. 


ORBITAL    REPRESENTATION 
OF   CHEMICAL    BONDING 

Our  rule  about  covalent  bond  formation  can  be 
applied  quite  simply  through  an  orbital  repre- 
sentation: 


2s 


2p 


H 


o  ooo 


"■  \a       O    OOO 

Bond    can  form.  (2) 


He 


He 


o    OOO 
O    OOO 


Bond   c an.no f  form. 

In  this  representation  there  is  no  need  to  con- 
sider the  next  higher  energy  level  cluster — the 
2s,  2/7  orbitals.  For  hydrogen  and  helium  these 
are  much  higher  in  energy  and  can  give  rise  only 
to  extremely  weak  attractions. 

ELECTRON   DOT   REPRESENTATION 
OF   CHEMICAL   BONDING 

The  sharing  of  electrons  can  be  shown  by  rep- 
resenting valence  electrons  as  dots  placed  be- 
tween the  atoms: 


H   +H 


H:H 


(i) 


We  shall  use  both  orbital  and  electron  dot  rep- 
resentations to  show  chemical  bonding. 

16-1.4    The  Bonding  of  Fluorine 

Our  explanation  of  chemical  bonding  is  of  value 
only  if  it  has  wide  applicability.  Let  us  examine 
its  usefulness  in  considering  the  compounds  of 
the  second-row  elements,  beginning  with  fluo- 
rine. 

Under  normal  conditions  of  temperature  and 
pressure,  fluorine  is  a  gas.  From  gas  density 
experiments  we  discover  that  a  molecule  of  fluo- 
rine contains  two  atoms.  There  is  a  chemical 
bond  between  the  two  fluorine  atoms.  Let  us  see 
if  our  expectations  agree  with  these  experimental 
facts. 


SEC.    16-1    I   THE    COVALENT   BOND 


279 


ORBITAL   REPRESENTATION    OF   THB 
BONDING    OF   FLUORINB 

A  fluorine  atom  has  the  orbital  occupancy  shown 
below: 


pacity  and  a  stable  compound  is  formed,  HF. 

In  each  of  these  cases,  F2  and  HF,  we  find  that 

the  fluorine  atom  forms  one  bond — in  F2  it  is  to 

a  second  fluorine  atom,  in  HF  it  is  to  a  hydrogen 


is 


2s 


zp 


3s 


3P 


*  ®       ®    <g®0     o  ooo 


w 


We  see  that  the  neutral  fluorine  atom  has  seven 
valence  electrons;  that  is,  seven  electrons  occupy 
the  outermost  partially  filled  cluster  of  energy 
levels.  This  cluster  of  energy  levels,  the  valence 
orbitals,  contains  one  electron  less  than  its  ca- 
pacity permits.  Fluorine,  then,  has  the  capacity 
for  sharing  one  electron  with  some  other  atom 
which  has  similar  capacity.  If,  for  example,  an- 
other fluorine  atom  approaches,  they  might  share 
a  pair  of  electrons  and  form  a  covalent  bond: 

Is  Zs  2p 


atom.  We  describe  this  single  bonding  capacity 
by  saying,  fluorine  is  univalent. 

ELECTRON    DOT   REPRESENTATION 
OF   THE    BONDING    OF   FLUORINE 

In  the  electron  dot  method  of  showing  chemical 
bonds  it  is  necessary  to  show  only  the  valence 
electrons.  In  fluorine  there  are  seven — the  pair 
of  electrons  in  the  Is  orbital  is  so  tightly  bound 


3s 


3P 


*  <g>        <g>    <8®0     O  OOO 

*  <8>        ®    ®®Q     O  OOO 


(5) 


After  the  second  atom  approaches,  sharing  its 
electron,  each  fluorine  atom  now  has  "filled"  all 
of  its  valence  orbitals.  No  additional  bonding 
capacity  remains.  Hence  F2  does  not  add  a  third 
or  fourth  atom  to  form  F3,  F4,  etc. 

Now  consider  the  possibility  of  the  bonding 
that  might  occur  if  a  fluorine  atom  encounters 
a  hydrogen  atom.  Again  fluorine  has  an  oppor- 
tunity to  share  electrons  with  a  second  atom 
having  a  partially  filled  valence  orbital: 

Is  2s  2p 


that  it  plays  little  role  in  the  chemistry  of  fluo- 
rine. In  this  representation,  then,  we  show  the 
reaction  between  two  fluorine  atoms  as  follows: 


:F-  +  -F: 


:F:F: 


(7) 


From  (7)  we  conclude  that  a  covalent  bond  can 
form  between  two  fluorine  atoms.  Furthermore, 
a  census  of  the  number  of  electrons  owned  or 


3s 


3p 


*  ®        <g>    ®80 


O  OOO 
O    OOO 


«5) 


Is 

Fluorine  now  has,  by  sharing  a  pair  of  electrons 
with  hydrogen,  "filled"  its  valence  orbitals.  It 
has  no  further  bonding  capacity.  The  same  is 
true  of  the  hydrogen  atom,  though  for  this  atom 
there  is  but  one  valence  orbital,  the  Is  orbital. 
Since  no  partially  filled  valence  orbitals  remain 
for  either  atom,  there  is  no  more  bonding  ca- 


2s 


2r 


shared  by  either  of  the  fluorine  atoms  shows  that 
the  valence  orbitals  are  filled.  For  example,  the 
fluorine  atom  on  the  left  feels  the  electrical  at- 
tractions of  eight  electrons  near  at  hand,*  as 

*  Remember,  we  are  omitting  from  the  discussion  the 
tightly  bound  h  electrons  of  each  fluorine  atom. 


280 


MOLECULES    IN    THE    GAS    PHASE    I    CHAP.     16 


shown  in  Figure  16-5.  Since  eight  electrons  is  just 
the  capacity  of  the  2s,  2p  valence  orbitals,  each 
fluorine  atom  has  reached  the  energetically  stable 
arrangement  of  an  inert  gas. 


The  fluorine    atom  The  fluorine  atom. 

on  the   left  on  the  right 

feels   eight-  electrons  feels  eight  electrons 


:    F :    -)F    : 


:  F  I:  F  : 


v.        ••     / 


Fig.  16-5.  The  electrons  near  each  fluorine  atom  in 
F:. 


Hydrogen  fluoride  can  be  represented  by  the 
electron  dot  picture 


:F    +    H 


F:H 


(5) 


Again,  a  census  of  the  number  of  electrons  near 
each  of  the  atoms  shows  that  this  is  a  stable 
arrangement.  True,  the  hydrogen  atom  has,  close 
at  hand,  only  two  electrons  whereas  fluorine  has 
eight.  This  is  energetically  desirable,  however, 
because  hydrogen  has  only  one  valence  orbital, 
the  Is  orbital.  Two  electrons  just  fill  this  orbital. 


The  fluorine  atom  The  hydrogen    atom 

feels   eight  electrons        feels   two   electrons 


[:   F    Ah 


F   \\    H     ) 


Fiq.   16-6.  The  electrons  near  each  atom  in  HF. 


We  see  that  the  bonding  of  a  fluorine  atom  to 
another  fluorine  atom  or  to  a  hydrogen  atom  can 
be  explained  in  terms  of  sharing  electrons  so  as 
to  fill  the  partially  filled  valence  orbitals.  This 


Is 


Zs 


2p 


3s 


sharing  makes  the  molecule  F2  (or  HF)  energeti- 
cally more  stable  than  the  separated  atoms  would 
be.  The  energy  stability  results  from  the  shared 
electrons  being  attracted  simultaneously  to  both 
positive  nuclei.  A  chemical  bond  results. 

ELECTRON    AFFINITY    OF    THE    FLUORINE    ATOM 

Experiment  shows  that  a  gaseous  fluorine  atom  can 
acquire  an  electron  to  form  a  stable  ion,  F~(g).  We  can 
discuss  the  energy  of  formation  of  this  ion  in  the  same 
way  that  we  treated  ionization  energies.  The  first  ioniza- 
tion energy  of  fluorine  atom  is  the  energy  required  to 
remove  an  electron  from  a  neutral  atom  in  the  gas  phase. 
We  shall  call  this  energy  E\.  Then  the  heat  of  reaction 
can  be  written  in  terms  of  Ex : 

*(g)  — *■  F+(gj  +  trig)       AH  =  E,  (9) 

The  second  ionization  energy,  E2,  refers  to  reaction  (10): 

F+(g)  — >-  F**(e)  +  e-(g)        AH  =  E2        (10) 

Now  we  can  add  a  new  process  with  an  energy  which 
might  logically  be  called  E0: 


?-(g) 


F(g)  +  e-ig)        AH  =  Eo 


UD 


Comparing  equations  (9),  (10),  and  (//),  we  see  that  £o 
is  just  the  ionization  energy  of  F~(g).  By  usual  practice, 
however,  the  reverse  of  reaction  (11)  is  usually  considered. 
Of  course  the  heat  of  reaction  (12)  is  just  the  negative  of 
that  of  reaction  (//): 


F(g)  +  e'ig) 


F-(g) 


AH  =    -En 


(12) 


The  energy  change  of  reaction  (12)  is  called  the  electron 
affinity  of  the  fluorine  atom.  It  is  symbolized  by  E  and, 
as  defined  here,  is  a  negative  quantity  if  heat  is  released 
when  the  ion  is  formed: 


E=  -Eo 


U3) 


Electron  affinities  are  difficult  to  measure  and  are 
known  reliably  for  only  a  small  fraction  of  the  hundred 
or  so  elements.  The  electron  affinity  of  fluorine  is  one 


that  is  known: 

E  =  —  83  kcal,  mole 

(14) 

or 

Eo  =  +83  kcal  mole 

(15) 

The  experimental  quantities  shown  in  (14)  and  (15)  indi- 
cate that  the  F~  ion  is  more  stable  than  a  fluorine  atom 
and  an  electron.  Energetically,  a  fluorine  atom  "wants" 
another  electron.  It  is  profitable  to  express  reaction  (12) 
in  terms  of  orbital  occupancy: 


Is 


Zs 


3s 


eg)       (g)    ®g0     O  *  •"<*"»  — 


0<75> 


T" 


SEC.    16-2    I    BONDING    CAPACITY    OF    THE    SECOND-ROW    ELEMENTS 


281 


The  neutral  fluorine  atom  has  seven  valence  electrons; 
that  is,  seven  electrons  occupy  the  highest  partially  filled 
cluster  of  energy  levels.  This  cluster  of  energy  levels  thus 
contains  one  fewer  electron  than  its  capacity  permits.  The 
electron  affinity  of  fluorine  shows  that  the  addition  of  this 
last  electron  is  energetically  favored.  This  is  in  accord 
with  much  other  experience  which  shows  that  there  is  a 
special  stability  to  the  inert  gas  electron  population. 

In  view  of  the  electron  affinity  of  a  fluorine  atom,  we 
can  speculate  on  what  would  be  the  result  of  a  collision 
between  two  fluorine  atoms.  Will  a  reaction  occur?  The 
energy  is  one  of  the  factors  which  determines  the  answer. 
First  let  us  consider  a  reaction  that  does  not  occur  spon- 
taneously. 


F(g)  +  F(g) 


F+(g)  +  F-(g) 


(17) 


Reaction  (17)  can  be  rewritten  in  two  steps: 

F(g) 
F(g)  +  e-(g) 


fluorine  atom  number  1  shares  one  valence  electron  of 
fluorine  atom  number  2.  Thus  a  part  of  the  electron  affin- 
ity of  fluorine  atom  number  1  is  "satisfied"  though  it  was 
not  necessary  to  take  the  electron  away  from  atom  num- 
ber 2.  Meanwhile,  fluorine  atom  number  2  is  getting  the 
same  sort  of  energy  benefit  from  the  valence  electron  of 
atom  number  I.  Each  fluorine  atom  has  acquired  another 
electron  at  least  a  part  of  the  time.  We  have  gained  a  part 
of  the  stability  of  reaction  (19)  without  paying  for  reac- 
tion (18).  The  most  energy  that  could  be  released  by  such 
an  electron  sharing  would  be  double  the  electron  affinity 
of  fluorine,  2  X  83  =  166  kcal.  This  takes  no  account  of 
the  work  done  in  bringing  the  two  positive  nuclei  near 
each  other.  Nor  can  we  expect  to  gain  the  whole  electron 
affinity  under  conditions  of  electron  sharing.  Yet  the 
energy  released  when  two  fluorine  atoms  form  a  bond  is 
36.6  kcal/mole,  a  reasonable  fraction  of  the  maximum 
possible. 


F+(g)  +  e-(g) 
F~(g) 


&H  =      Ex  =    401.5  kcal/mole 
AH  -  -Eo  =  -83 


Net        F(g)  +  F(g) 


F+(g)  +  F~(g) 


AH  = 


318.5 


(18) 
U9) 

(17) 


We  see  that  reaction  (77)  is  energetically  unfavorable. 
The  stability  of  F~  is  more  than  outweighed  by  the  diffi- 
culty of  removing  an  electron  from  another  fluorine  atom. 
There  is  another  possible  consequence  of  a  collision 
between  two  fluorine  atoms.  The  two  atoms  can  remain 
together  to  form  a  molecule.  Each  atom  has  a  valence 
electron  in  a  "half-filled"  orbital.  We  can  imagine  these 
two  atoms  orienting  so  that  these  "half-filled"  orbitals 
overlap  in  space.  Then  the  "half-filled"  valence  orbital  of 


Now  we  can  say  why  the  chemical  bond  forms  between 
two  fluorine  atoms.  First,  the  electron  affinity  of  a  fluorine 
atom  makes  it  energetically  favorable  to  acquire  one  more 
electron.  Two  fluorine  atoms  can  realize  a  part  of  this 
energy  stability  by  sharing  electrons.  All  chemical  bonds 
form  because  one  or  more  electrons  are  placed  so  as  to  feel 
electrostatic  attraction  to  two  or  more  positive  nuclei 
simultaneously. 


16-2     BONDING  CAPACITY  OF  THE  SECOND-ROW  ELEMENTS 


In  Chapter  6  we  saw  that  the  chemical  com- 
pounds of  the  third-row  elements  display  a 
remarkable  regularity.  Return  to  Chapter  6  and 
reread  Section  6-6.2.  The  same  simple  trend  in 
chemical  formulas  is  discovered  in  the  second 
row  of  the  periodic  table.  Now  we  have  a  basis 
for  explaining  why  these  trends  are  found. 


16-2.1    The  Bonding  Capacity  of  Oxygen  Atoms 

The  neutral  oxygen  atom  has  eight  electrons.  Six 
of  these  occupy  the  2s,  2p  orbitals  and  are  much 
more  easily  removed  than  the  two  in  the  Is 
orbital.  Therefore  oxygen  has  six  valence  elec- 
trons. The  25,  2p  orbitals  are  the  valence  orbitals. 
They  can  accommodate  the  valence  electrons  in 
two  ways,  as  follows: 


Is 


2s 


3s 


3p 


®so    o  ooo 


(20) 


or 


p    ®         <g>     &O0      O  COO 


(21) 


282 


MOLECULES    IN    THE    GAS    PHASE   I    CHAP.    16 


Remember  the  spatial  arrangement  of  the  p  or- 
bitals?  Each  one  protrudes  along  one  of  the 
three  cartesian  axes  (as  shown  in  Figure  15-9). 
If  the  electrons  have  the  orbital  occupancy  of 
(20),  then  two  electrons  occupy  the  p  orbital 


atom  has  partially  filled  valence  orbitals.  Elec- 
tron sharing  can  occur,  placing  electrons  close 
to  two  nuclei  simultaneously.  Hence  a  stable 
bond  can  occur.  This  is  shown  in  representations 
(22)  and  (23). 


Is 


2s 


3s  3P 

O  OOO 

o  ooo 

2s  2p 


(22) 


or 


:0. 


'H 


:0'.K 


(23) 


protruding  along  the  x  axis  (px)  and  two  elec- 
trons occupy  the/?  orbital  protruding  along  the>> 
axis  (pv).  In  the  occupancy  of  (21),  two  electrons 
occupy  px,  only  one  electron  occupies  pu,  and  the 
last  is  in  />,.  Thus  (21)  differs  from  (20)  by  the 
movement  of  one  of  the  electrons  from  pv  to  a 
different  region  of  space,  pz.  Since  electrons  repel 
each  other,  we  can  expect  that  the  configuration 
which  keeps  the  electrons  farther  apart,  (21),  is 
the  lower  in  energy.  Experiment  shows  that  it  is 
and  we  shall  base  our  discussion  of  the  bonding 
of  oxygen  on  orbital  occupancy  (21).  However, 
the  occupancy  represented  by  (20)  also  contrib- 
utes to  the  chemistry  of  oxygen  atoms. 

Suppose  a  hydrogen  atom  approaches  an  oxy- 
gen atom  in  its  most  stable  state,  (27).  Each 


In  either  representation,  (22)  or  (23),  we  see 
that  there  is  residual  bonding  capacity  remaining 
in  the  species  OH.  In  (22)  the  third  2p  orbital  has 
a  single  electron  but  a  capacity  for  two.  This 
means  more  bonding  can  occur.  In  (23)  a  census 
of  the  electrons  near  the  oxygen  atom  indicates 
there  are  only  seven.  The  oxygen  atom  would  be 
more  stable  if  it  could  add  one  more  electron. 
With  either  representation,  we  conclude  that  OH 
should  be  able  to  react  with  another  hydrogen 
atom.  See  representations  (24),  (25). 

Now  we  have  the  compound  H20.  By  either 
representation,  the  bonding  capacity  of  oxygen 
is  expended  when  two  bonds  are  formed.  Oxygen 
is  said  to  be  divalent,  and  the  compound  H20  is 
extremely  stable.  Each  of  the  atoms  in  H20  has 
filled  its  valence  orbitals  by  electron  sharing. 


is 


2s 


2s 


O    (XT) 
O  CCO 

o  ooo 


(24) 


2S 


OiH 


K 


•O'.K 
H 


(25) 


SEC.    16-2   I   BONDING   CAPACITY    OF   THE   SECOND-ROW    ELEMENTS 


283 


reaction  between  two  OH  molecules     The  electron  dot  representation  is 


Though  OH  is  reactive,  it  is  a  cluster  of  atoms 
with  sufficient  stability  to  be  identified  as  a  mole- 
cule. It  is  present  in  a  number  of  high  tempera- 
ture flames,  for  example.  Its  chemistry  might  be 
expected  to  be  like  that  of  fluorine  atoms.  Com- 
pare the  electron  dot  formulas 


:F 


0:H 


Since  two  fluorine  atoms  react,  forming  a  cova- 
lent  bond,  we  can  expect  two  OH  molecules  to 
do  the  same  sort  of  thing 

H  H 

6:   — *-  :6  6: 


o.+ 

H 


(26) 


H 


Reaction  (26)  yields  the  compound  H202.  This 
is  the  formula  of  the  well-known  substance, 
hydrogen  peroxide.  By  these  considerations  of 
chemical  bonding,  we  see  that  the  structure  of 
HjOj  must  involve  an  oxygen-oxygen  bond: 


H 

I 
O— O 

I 
H 


(27) 


EXERCISE  16-1 


Predict  the  structure  of  the  compound  S2CI2  from 
the  electron  dot  representation  of  the  atoms. 
After  you  have  predicted  it,  turn  back  to  Figure 
6-12,  p.  103,  and  check  your  expectation. 


:OF: 

:F:'" 

Again  we  find  that  oxygen  is  divalent. 


(29) 


EXERCISE  16-2 


Draw  orbital  and  electron  dot  representations  of 
each  of  the  following  molecules:  OF,  F202,  HOF, 
HF02.  Which  of  these  is  apt  to  be  the  most 
reactive? 


16-2.2    The  Bonding  Capacity 
of  Nitrogen  Atoms 

For  the  same  reason  we  discussed  for  oxygen 
atoms,  the  nitrogen  atom  is  most  stable  when  it 
has  the  maximum  number  of  partially  filled 
valence  orbitals.  This  keeps  the  electrons  as  far 
apart  as  possible.  The  most  stable  state  of  the 
nitrogen  atom  is  as  follows: 


rJV 


is 


0 


2s 


0     000  v» 


It  is  now  straightforward  to  predict  that  nitrogen 
will  form  a  stable  hydrogen  compound  with  for- 
mula NH3.  Nitrogen  is  trivalent.  A  similar  com- 
pound, NF3,  will  be  formed  with  fluorine.  The 
electron  dot  formulas  are 


OXYGEN-FLUORINE   COMPOUNDS 

It  is  a  simple  matter  to  predict  that  oxygen  will 
form  a  stable  compound  with  two  fluorine 
atoms,  FsO.  The  orbital  representation  is* 

Is  2s  2x> 


/ 


0 
0 
0 


0    00Q 
0    0&0 


(28) 


*  Henceforth  we  will  omit  empty  orbitals  much  higher 
in  energy  than  the  valence  orbitals. 


H 

:F: 

(31) 

:N:H 

and         :  N :  F : 

H 

ammonia 

:F: 

nitrogen  trifluoride 

(32) 

EXERCISE  16-3 

The  molecule  NH2  has  residual,  unused  bonding 
capacity  and  is  extremely  reactive.  The  molecule 
N2Hi  (hydrazine)  is  much  more  stable.  Draw  an 
electron  dot  representation  of  the  bonding  of 
hydrazine.  Draw  its  structural  formula  (show 
which  atoms  are  bonded  to  each  other). 


284 


MOLECULES    IN    THE    GAS    PHASE   I    CHAP.     16 


16-2.3    The  Bonding  Capacity  off  Carbon  Atoms 

There  are  a  number  of  orbital  occupancies  that 
we  might  consider  for  the  carbon  atom: 


Is 


2S 


2P 


>C 


® 


<s>   ®oo  <«> 


<g>    00O  <*> 


First  let  us  compare  (33)  and  (34).  By  our  con- 
ventional argument  that  electrons  repel  each 
other,  configuration  (34)  should  be  more  stable 
than  (33).  The  second,  (34),  places  one  electron 
in  each  of  the  px  and  pv  orbitals  whereas  (33) 
places  two  electrons  in  the  same  orbital,  px.  It  is 
an  experimental  fact  that  (34)  is  more  stable 
than  (33). 

Now  we  can  predict  the  chemistry  of  the  car- 
bon atom  in  this  state.  It  should  be  divalent, 
forming  compounds  CH2  and  CF2.  Let  us  con- 
sider one  of  these,  say  CH2. 


1* 


H 


Is 


2s 


\ 


C'.H 
H 


(36) 


Here  is  a  situation  we  haven't  met  before. 
After  using  the  two  available  partially  filled 
orbitals  to  form  covalent  bonds  with  hydrogen 
atoms,  there  remains  a  vacant  valence  orbital. 
In  the  electron  dot  formulation  (36)  we  see  that 
the  carbon  atom  finds  itself  near  only  six  elec- 
trons in  CH2.  The  valence  orbitals  will  accom- 
modate eight  electrons.  Because  one  valence  or- 


bital is  completely  vacant,  we  can  expect  C7/2  to 
be  reactive. 

In  fact,  both  CH2  and  CF2  are  considered  to 
be  stable*  but  extremely  reactive  molecules. 
Though  there  is  reaction  mechanism  evidence 
verifying  the  existence  of  each  species,  it  is  not 
possible  to  prepare  either  substance  pure.  This 
great  reactivity  shows  that  energy  considerations 
favor  the  use  of  all  four  of  the  valence  orbitals  if 
possible.  This  argument  leads  us  to  consider  a 
third  orbital  occupancy: 


lr 


2s 


2P 


0     020    <"> 


With  orbital  occupancy  (37),  a  carbon  atom 
has  four  half-filled  valence  orbitals.  True,  (37)  is 
somewhat  less  stable  than  (34)  because  an  elec- 
tron was  raised  from  the  25  energy  level  to  the 
slightly  higher  2p  energy  level.  This  process  is 
called  "promoting"  the  electron.  On  the  other 
hand,  the  promotional  energy  is  not  very  large, 
and  in  return  for  it  the  carbon  atom  acquires  the 
capacity  to  form  four  covalent  bonds.  Each  co- 
valent bond  increases  stability,  more  than  com- 
pensating for  the  energy  investment  in  promoting 
one  of  the  2s  electrons.  With  orbital  occupancy 
(37),  carbon  can  share  pairs  of  electrons  with, 
for  example,  four  hydrogen  atoms  or  four  fluo- 
rine atoms.  Hence,  carbon  is  tetravalent: 


H 

H:C:H 

H 

methane 


:F: 

:  F :  C :  F : 

":F:" 

carbon  tetrafluoride 


(38) 


(39) 


*  The  molecule  CH2  is  stable  in  the  sense  that  it  does 
not  spontaneously  break  into  smaller  fragments.  It  is 
reactive  because  other  molecules  formed  from  this  group 
of  atoms  have  much  lower  energy. 


SEC.    16-2   |   BONDING    CAPACITY    OF   THE   SECOND-ROW    ELEMENTS 


285 


EXERCISE  16-4 

Draw  electron  dot  formulas  for  the  molecules 
CH3,  CF3,  CHF3,  CH2F2,  CH3F.  Which  will  be 
extremely  reactive? 

EXERCISE  16-5 

Draw  an  electron  dot  and  a  structural  formula 
for  the  molecule  C2H6  (ethane)  which  forms  if 
two  CH3  molecules  are  brought  together.  Explain 
why  C2H6  is  much  less  reactive  than  CH3. 


is 
B        <g) 


2  s  2P 

0"0(20 


(44) 


16-2.4    The  Bonding  Capacity  of  Boron  Atoms 

The  boron  atom  presents  the  same  sort  of  option 
in  orbital  occupancy  as  does  carbon: 

Is  2s  2p 


H     H     H 


sB       (g) 


**    ®        0    00O  w 

The  electron  configuration  (41)  is  somewhat 
higher  in  energy  than  (40).  It  is  necessary  to 
promote  a  2s  electron  to  the  2p  state  to  obtain 
(41).  In  return,  however,  the  boron  atom  gains 
bonding  capacity.  Whereas  a  boron  atom  can 
form  only  one  covalent  bond  in  configuration 
(40),  it  can  form  three  in  configuration  (41). 
Since  each  bond  lowers  the  energy,  the  chemistry 
of  boron  is  fixed  by  the  electron  configuration 
(41). 

Now  we  can  expect  that  boron  will  be  tri- 
valent.  We  predict  that  there  should  be  molecules 
such  as  BH3  and  BF3 


H 

B:H 
H 

F: 

B:F: 

F: 


We  cannot  help  noticing,  however,  that  there 
remains  a  completely  vacant  valence  orbital.  For 
example,  for  BH3  the  orbital  representation 
would  be  as  shown  in  (44). 

The  last  vacant  2p  orbital  is  reminiscent  of  the 
configuration  found  for  CH2  [see  (35)].  Since 
CHj  is  very  reactive,  presumably  BH3  will  be  the 


same.  Such  is  the  case.  There  is  only  indirect 
evidence  establishing  the  existence  of  BH3.  In- 
stead boron  forms  a  series  of  unusual  compounds 
with  hydrogen,  the  simplest  of  which  is  called 
0OO     W     diborane,  B2H6. 

You  might  wish  to  predict  the  structure  of  diborane 
(which  is  now  known)  but  do  not  be  discouraged  if  you 
are  not  able  to.  Its  structure,  once  elucidated,  came  as 
quite  a  surprise  to  even  the  most  sophisticated  chemists. 
The  explanation  of  the  structure  is,  even  today,  composed 
of  a  large  proportion  of  words  and  a  small  proportion  of 
understanding. 

Boron  is  an  obliging  element.  On  the  one  hand 
it  conforms  to  our  expectation  that  the  unused 
valence  orbital  will  affect  the  bonding  capacity 
of  boron,  as  shown  by  the  reactivity  of  BH3.  On 
the  other  hand  boron  conforms  to  our  expecta- 
tion that  electron  configuration  (41)  will  make 
boron  trivalent.  The  compound  BF3  is  a  stable, 
gaseous  compound  and,  in  contrast  to  BH3,  is 
readily  prepared  in  a  pure  form.  The  explanation 
of  how  the  fluorine  atoms  are  able  to  satisfy  in 
part  the  bonding  capacity  of  the  vacant  2p  or- 
bital (though  hydrogen  atoms  cannot)  must  wait 
until  a  later  chemistry  course.  For  our  interest 
here,  BF3  is  the  most  stable  boron-fluorine  com- 
pound and  it  demonstrates  the  trivalent  bonding 
capacity  of  boron. 

16-2.5    The  Bonding  Capacity 
of  Beryllium  Atoms 

The  beryllium  atom,  like  boron  and  carbon,  can 
promote  an  electron  in  order  to  form  more 
chemical  bonds: 


(42) 


(43) 


,Be 


Is 

0 


2s 


0     000  <«> 


286 


MOLECULES    IN    THE    GAS    PHASE    I    CHAP.     16 


Therefore  we  should  expect  in  the  gaseous  state 
to  find  molecules  such  as  BeH2  and  BeF2.  These 
molecules  have  been  detected.  On  the  other 
hand,  beryllium  has  the  trouble  boron  has,  only 
in  a  double  dose.  It  has  two  vacant  valence  or- 
bitals.  As  a  result,  BeH2  and  BeF2  molecules,  as 
such,  are  obtained  only  at  extremely  high  tem- 
peratures (say,  above  1000°K).  At  lower  temper- 
atures these  vacant  valence  orbitals  cause  a 
condensation  to  a  solid  in  which  these  orbitals 
can  participate  in  bonding.  We  shall  discuss  these 
solids  in  the  next  chapter. 

16-2.6    The  Bonding  Capacity  of  Lithium  Atoms 

There  is  little  new  to  be  said  about  the  bonding 
capacity  of  a  lithium  atom.  With  just  one  valence 
electron,  it  should  form  gaseous  molecules  LiH 
and  LiF.  Because  of  the  vacant  valence  orbitals, 
these  substances  will  be  expected  only  at  ex- 
tremely high  temperatures.  These  expectations 
are  in  accord  with  the  facts,  as  shown  in  Table 
16-1,  which  summarizes  the  formulas  and  the 
melting  and  boiling  points  of  the  stable  fluorides 
of  the  second-row  elements.  In  each  case,  the 
formula  given  in  the  table  is  the  actual  molecular 
formula  of  the  species  found  in  the  gas  phase. 


16-2.7    Valence 

Very  often  the  word  valence  is  used  in  discussins 
the  nature  of  chemical  bonding.  Unfortunately 
this  word  has  been  used  as  a  noun  to  mean  a 
number  of  different  things.  Sometimes  valence 
has  been  used  to  mean  the  charge  on  an  ion, 
sometimes  it  has  meant  the  total  number  of 
atoms  to  which  a  particular  atom  will  bond,  and 
at  other  times,  the  word  valence  has  been  used 
to  mean  oxidation  number.  Perhaps  the  most 
widely  accepted  definition  of  the  word  is  that  it 
is  the  number  of  hydrogen  atoms  with  which  an 
atom  can  combine,  or  release,  in  a  chemical  re- 
action. It  is  clear  that  a  word  with  so  many  mean- 
ings might  confuse  a  discussion  of  chemical 
bonding.  For  this  reason  we  have  avoided  and 
will  continue  to  avoid  using  the  word  as  a  noun 
in  this  book. 

We  have,  however,  made  a  careful  definition 
of  the  term  "valence  electrons"  ("the  electrons 
that  are  most  loosely  bound";  see  p.  269). 
We  have  also  used  carefully  the  term  "valence 
orbitals"  to  mean  the  entire  cluster  of  orbitals 
of  about  the  same  energy  as  those  which  are 
occupied  by  the  valence  electrons.  In  both  of 
these  uses,  the  word  valence  is  used  as  an 
adjective. 


Table  16-1.     the  fluorides  of  the  elements  in  the  second  row 

OF    THE    PERIODIC    TABLE 

Li  Be  B  C  N 


Formula 

LiF 

BeFj 

BF, 

CF4 

NF, 

F,0 

F, 

Melting  point  (°K) 

1143 

1073 

146 

89 

56 

49 

50 

Boiling  point  (°K) 

1949 

— 

172 

145 

153 

128 

85 

16-3    TREND  IN  BOND  TYPE  AMONG  THE  SECOND-ROW  FLUORIDES 


We  have  termed  the  chemical  bond  in  the  hydro- 
gen molecule,  H2,  a  covalent  bond.  This  indicates 
that  electrons  are  shared  so  that  they  are  simul- 
taneously and,  on  the  average,  equally  near  two 
nuclei.  This  makes  the  system  more  stable  and  a 
chemical  bond  results. 


All  chemical  bonds  occur  because  electrons 
can  be  placed  simultaneously  near  two  nuclei. 
Yet  it  is  often  true  that  the  electron-sharing 
which  permits  this  is  not  exactly  equal  sharing. 
Sometimes  the  electrons,  though  close  to  both 
nuclei,  tend  to  distribute  nearer  to  one  nucleus 


SEC.    16-3    I    TREND    IN    BOND    TYPE    AMONG    THE    FIRST-ROW    FLUORIDES 


287 


than  to  the  other.  We  can  see  why  by  contrasting 
the  chemical  bonding  in  gaseous  fluorine,  F2,  and 
in  gaseous  lithium  fluoride,  LiF. 

16-3.1    The  Bonding  in  Gaseous 
Lithium  Fluoride 

We  have  already  treated  the  bonding  in  an  F2 
molecule.  Since  neither  fluorine  atom  can  pull  an 
electron  entirely  away  from  the  other,  they  com- 
promise by  sharing  a  pair  of  electrons  equally. 
How  does  the  chemical  bonding  in  the  lithium 
fluoride  molecule  compare? 

As  we  have  mentioned  earlier,  lithium  has  one 
valence  electron,  hence  can  share  a  pair  of  elec- 
trons with  one  fluorine  atom: 


Li 


<8> 


Is 


When  the  bonding  electrons  move  closer  to  one 
of  the  two  atoms,  the  bond  is  said  to  have  ionic 
character. 

In  the  most  extreme  situation,  the  bonding 
electrons  move  so  close  to  one  of  the  atoms  that 
this  atom  has  virtually  the  electron  distribution 
of  the  negative  ion.  This  is  the  case  in  gaseous 
LiF.  In  an  electron  dot  representation,  we  might 
show 


or 


Li   :F: 


Li+   F- 


(50) 


(51) 


When  a  formula  like  (50)  or  (51)  provides  a  use- 
ful basis  for  discussing  the  properties  of  a  mole- 


2s 


Is 
2s  2P 


<s  ooo 


(46) 


or  Ll'.F'. 


(47) 


Thus  we  can  expect  a  stable  molecular  species, 
LiF.  The  term  "stable"  again  means  that  energy 
is  required  to  disrupt  the  molecule.  The  chemical 
bond  lowers  the  energy  because  the  bonding 
electron  pair  feels  simultaneously  both  the  lith- 
ium nucleus  and  the  fluorine  nucleus.  That  is  not 
to  say,  however,  that  the  electrons  are  shared 
equally.  After  all,  the  lithium  and  fluorine  atoms 
attract  the  electrons  differently.  This  is  shown  by 
the  ionization  energies  of  these  two  atoms: 

F(g)  —+  F+(g)  +  e-(g) 

A//  =  401.5  kcal/mole    (48) 

Li(g)  — >-  Li+fgj  +  e-(g) 

AH  =  124.3  kcal/mole    (49) 

Clearly  the  fluorine  atom  holds  electrons  much 
more  strongly  than  does  the  lithium  atom.  As  a 
result,  the  electron  pair  in  the  lithium  fluoride 
bond  is  more  strongly  attracted  to  the  fluorine 
atom  than  to  the  lithium.  The  energy  is  lower  if 
the  electrons  spill  toward  the  fluorine  atom. 


cule,  the  bond  in  that  molecule  is  said  to  be  an 
ionic  bond. 

CONTRAST   OF   COVALENT    AND 
IONIC    BONDS 

The  fluorine  molecule  is  held  together  by  the 
energy  gain  resulting  from  placing  a  bonding 
pair  of  electrons  near  both  fluorine  nuclei  simul- 
taneously. The  electrons  move  about  in  the 
molecule  in  such  a  manner  that,  on  the  average, 
they  are  distributed  symmetrically  between  and 
around  the  identical  fluorine  nuclei.  This  sym- 
metrical distribution  is  reasonable,  since  the  two 
fluorine  nuclei  attract  the  bonding  electrons 
equally.  The  lithium  fluoride  molecule  also  is 
held  together  by  the  energy  gain  resulting  from 
placing  a  bonding  pair  of  electrons  near  the 
lithium  and  fluorine  atoms  simultaneously.  In 
this  case,  however,  the  electrons  move  in  such  a 
way  as  to  remain  closer  to  the  fluorine  than  the 


288 


MOLECULES    IN    THE    GAS    PHASE    !    CHAP.    16 


Covalent  bond 


F-F 


^ 


Covalent  bond 
yvith  partial 
ionic  character 


F-Ct 


Ionic    bond 


F-Li 


<5k 


.jss> 


<=*- 


Fig. 


76-7.  Electron 


distributions     m     various     bond 


lithium  atom.  Fluorine  attracts  the  bonding  elec- 
trons more  strongly  than  does  lithium. 

We  see  again  that  there  is  but  one  principle 
which  causes  a  chemical  bond  between  two 
atoms:  all  chemical  bonds  form  because  electrons 
are  placed  simultaneously  near  two  positive  nuclei. 
The  term  covalent  bond  indicates  that  the  most 
stable  distribution  of  the  electrons  (as  far  as 
energy  is  concerned)  is  symmetrical  between  the 
two  atoms.  When  the  bonding  electrons  are 
somewhat  closer  to  one  of  the  atoms  than  the 
other,  the  bond  is  said  to  have  ionic  character. 
The  term  ionic  bond  indicates  the  electrons  are 
displaced  so  much  toward  one  atom  that  it  is  a 
good  approximation  to  represent  the  bonded 


atoms  as  a  pair  of  ions  near  each  other.  Figure 
16-7  shows  schematically  how  the  electron  dis- 
tributions are  pictured  in  covalent,  partially 
ionic,  and  ionic  bonds.  The  figure  also  shows 
how  the  electrons  might  look  in  an  instantaneous 
snapshot.  In  each  type  of  bond,  the  electron- 
nucleus  attractions  account  for  the  energy  sta- 
bility of  the  molecule. 

THE    ELECTRIC    DIPOLE 
OF    THE    IONIC    BOND 

The  spilling  of  negative  electric  charge  toward 
one  of  the  atoms  in  an  ionic  bond  causes  a 
charge  separation.  This  can  be  represented 
crudely  as  in  the  last  drawing  in  Figure  16-8. 
The  molecule  is  electrically  positive  at  the  lithium 
end  and  electrically  negative  at  the  fluorine  end. 
It  is  said  to  possess  an  electric  dipole.  The  mole- 
cule is  then  called  a  polar  molecule.  The  forces 
between  molecules  possessing  electric  dipoles  are 
much  stronger  than  those  between  nonpolar 
molecules.  We  shall  see  in  Chapter  17  that  these 
forces,  too,  involve  the  same  electrical  interac- 
tions we  have  discussed  here. 

The  last  representation  of  Figure  16-8  is  com- 
monly used  and  it  is  the  simplest  way  of  showing 
a  bond  dipole.  The  arrow  means  that  the  nega- 
tive charge  is  mainly  at  one  end  of  the  bond.  The 
directional  property  of  the  arrow  implies  that 
the  force  this  molecule  exerts  on  another  mole- 
cule depends  upon  the  direction  of  approach  of 
the  second  molecule. 

16-3.2     Ionic  Character  in  Bonds  to  Fluorine 

We  can  expect  the  effects  just  discussed  to  be  at 
work  in  the  bonds  fluorine  forms  with  other  ele- 


Fig. 


Li 


16-8.  Representations   of   the   electric   dipole   of 

gaseous  lithium  fluoride. 


SEC.     16-3    I    TREND    IN    BOND    TYPE    AMONG    THE    FIRST-ROW    FLUORIDES 


289 


ments.  The  ionization  energies  of  the  elements 
give  us  a  rough  clue  to  the  electron-nuclear 
attractions.  Table  16-11  compares  the  ionization 
energies  of  each  element  of  the  second  row  with 
that  of  fluorine.  The  last  column  describes  the 
type  of  chemical  bond. 

The  trend  in  bond  type  shown  in  Table  16-11 
has  important  influence  on  the  trend  in  proper- 
ties of  the  fluorine  compounds.  The  trend  arises 
because  of  the  increasing  difference  between 
ionization  energies  of  the  two  bonded  atoms. 


gen  cannot  be  predicted  from  its  measured  ioniza- 
tion energy. 

Examination  of  the  properties  of  a  number  of 
compounds  involving  hydrogen  indicates  that 
the  ionic  character  of  bonds  to  hydrogen  are 
roughly  like  those  of  an  element  with  ionization 
energy  near  200  kcal/mole.  Thus  the  hydrogen 
fluorine  bond  in  HF  is  ionic  and  chemists  believe 
that  the  electrons  are  spilled  toward  the  fluorine 
atom,  leaving  the  hydrogen  atom  with  a  partial 
positive  charge.  Hydrogen  acts  like  an  element 


Table  16-11.     bond  types  in  some  fluorine  compounds 

IONIZATION  ENERGIES  (kcal/mole) 


COMPOUND 

BOND 

ELEMENT  BONDED  TO  F 

FLUORINE 

BOND  TYPE 

FF 

F— F 

F 

401.5 

401.5 

Covalent 
slightly 

OF2 

O— F 

O 

313.8 

401.5 

ionic 

NF3 

N— F 

N 

335 

401.5 

increasing 
ionic 

CF, 

C— F 

C 

259.5 

401.5 

character         increasing 

covalent 

character 

BF3 

B— F 

B 

191.2 

401.5 

slightly 

BeF2 

Be— F 

Be 

214.9 

401.5 

covalent 

LiF 

Li— F 

Li 

124.3 

401.5 

'        Ionic 

16-3.3     Ionic  Character  in  Bonds  to  Hydrogen 

In  Chapter  6  the  element  hydrogen  was  charac- 
terized as  a  family  by  itself.  Often  its  chemistry 
distinguishes  it  from  the  rest  of  the  periodic 
table.  We  find  this  is  the  case  when  we  attempt 
to  predict  the  ionic  character  of  bonds  to  hy- 
drogen. 

The  ionization  energy  of  the  hydrogen  atom, 
313.6  kcal/mole,  is  quite  close  to  that  of  fluorine, 
so  a  covalent  bond  between  these  two  atoms  in 
HF  is  expected.  Actually  the  properties  of  HF 
show  that  the  molecule  has  a  significant  electric 
dipole,  indicating  ionic  character  in  the  bond. 
The  same  is  true  in  the  O — H  bonds  of  water 
and,  to  a  lesser  extent,  in  the  N — H  bonds  of 
ammonia.  The  ionic  character  of  bonds  to  hydro- 


with  lower  ionization  energy  than  fluorine.  The 
same  is  true  but  in  decreasing  amount  for  hydro- 
gen when  it  is  bonded  to  oxygen  and  to  nitrogen. 
The  carbon-hydrogen  bond  has  only  a  slight 
ionic  character.  At  the  other  end  of  the  periodic 
table,  gaseous  lithium  hydride  is  known  to  have 
a  significant  electric  dipole  but  now  with  the 
electric  dipole  turned  around.  In  LiH  the  elec- 
trons are  spilled  toward  the  hydrogen  atom, 
leaving  the  lithium  atom  with  a  partial  positive 
charge.  This  is  in  accord  with  the  low  ionization 
energy  of  lithium,  124.3  kcal/mole,  well  below 
the  value  of  200  kcal/mole  that  we  have  assigned 
to  hydrogen.  For  our  purposes,  it  suffices  to  dis- 
cuss the  bonding  of  hydrogen  in  terms  of  an 
apparent  ionization  energy  near  200  kcal/mole. 


290 


MOLECULES   IN   THE   GAS   PHASE   I   CHAP.    16 


16-3.4    Bond  Energies  and  Electric  Dipoles 

It  is  found  experimentally  that  a  bond  between  two 
atoms  with  very  different  ionization  energies  tends  to  be 
stronger  than  a  bond  between  atoms  with  similar  ioniza- 
tion energies.  Since  electric  dipoles  are  caused  by  differ- 
ences in  the  ionization  energies  of  bonded  atoms,  we  can 
conclude  that  strong  bonds  are  expected  in  molecules 
with  electric  dipoles. 

For  example,  contrast  the  bond  energies  of  the  gaseous 
molecules  Na2,  Clj,  and  NaCl : 


Na2fs; 

CUfg) 

NaClfgj 


2NafgJ 
2C\(g) 

Nate;  +  cite; 


Aff(Na2)  =  17  kcal 
Atf(Oi)  =  57  kcal 
A//(NaCl)  =  ? 


A  rough  estimate  of  the  bond  energy  of  NaCl  could  be 
based  upon  the  bond  energies  of  Na2  and  Cl2,  17  kcal 
and  57  kcal.  Since  the  17  kcal  bond  energy  of  Na2  is 
derived  from  the  sharing  of  an  electron  pair  between  two 
sodium  atoms,  the  one  sodium  atom  in  NaCl  might  be 
expected  to  contribute  one-half  this  amount  to  the  bond 
energy  of  NaCl,  V  =  8.5  kcal.  In  a  similar  way,  the 
single  chlorine  atom  in  NaCl  might  contribute  one-half 
the  bond  energy  of  Cl2,  Af  =  28.5  kcal.  Thus  we  arrive 
at  an  estimate  of  A#(NaCl): 

^(Nad)  (estimated)  =  W*  +  "*X> 
=  8.5  +  28.5  =  37.0 

Experimentally  we  discover  that  A#(NaCl)  is  much 
larger,  98.0  kcal— a  discrepancy  of  98  -  37  =  61  kcal. 
This  discrepancy  is  explained  in  terms  of  the  large  dif- 
ference in  ionization  energy  of  sodium  and  chlorine 
atoms: 

Cite;  — *•  Cl+te;  +  e~         Ei  =  300  kcal 
Nate;  — *■  Na+te;  +  er        E,  =  118  kcal 


The  large  difference  implies  that  the  energy  is  lowered 
even  more  than  37  kcal  because  the  electron  pair  need  not 
remain  equally  shared  between  the  two  atoms  in  NaCl. 
Instead  they  can  concentrate  nearer  the  atom  that  holds 
electrons  more  tightly  (the  chlorine  atom)  if  a  net  lower- 
ing of  the  energy  results. 

Table  16-1 1 1  collects  some  data.  The  existence  of  a 
correlation  between  the  ionization  energy  difference, 
Ei(X)  —  E\(Y),  and  the  bond  energy  discrepancy, 
AHxy  —  i(A/fxi  +  A//y,),  is  obvious. 

Needless  to  say,  if  ionic  character  affects  the  energy 
stability  of  a  chemical  bond  it  also  affects  the  chemistry 
of  that  bond.  The  tendency  toward  minimum  energy  is 
one  of  the  factors  that  determine  what  chemical  changes 
will  occur.  As  a  bond  becomes  stronger,  more  energy  is 
required  to  break  that  bond  to  form  another  compound. 
Hence  we  see  that  ionic  bonds  are  favored  over  covalent 
bonds  and  that  ionic  character  in  a  bond  affects  its 
chemistry. 


EXERCISE  16-6 

From  the  following  bond  energy  data  and  the  ionization 
energies  given  in  Table  15-111,  calculate  the  entries  in  the 
last  two  columns  of  Table  16-111  for  the  compounds  LiF 
and  LiBr.  The  ionization  energy,  Eu  for  bromine  atom  is 
273  kcal/mole. 


Li2te; 

F2fg; 

Br2te; 

LiFte; 

LiBrte; 


2Lite; 
2Fte; 

2Brte; 

Lite;  +  Fte; 

Lite;  +  Brte; 


AH  =  25  kcal 
AH  =  36 
AH  =  45.5 
AH  =  137 
AH  =  101 


Table  16-111.     ionization  energy  differences  and  bond 

ENERGY    DISCREPANCIES 

MOLECULE 

X  Y  XY  Ei(X)-Ei(Y) 


AHxy  -  \(AHx*  +  AHy,) 


Na 

K 

NaKte; 

18  kcal/mole 

0  kcal/mole 

CI 

Br 

ClBrte; 

27 

+1 

CI 

Li 

LiClte; 

176 

58 

ci 

Na 

NaClte; 

182 

61 

CI 

K 

Kcite; 

200 

66.5 

F 

Li 

LiFte; 

277 

106 

16-4    MOLECULAR  ARCHITECTURE 


The  properties  of  a  molecule  are  primarily  deter- 
mined by  the  bond  types  which  hold  it  together 


and  by  the  molecular  "architecture."  By  archi- 
tecture we  mean  the  structure  of  the  molecule— 


SBC.    16-4   I    MOLBCULAR    ARCHITBCTURB 


291 


the  shape  of  the  molecule.  We  shall  investigate 
what  is  known  about  the  molecular  structures  of 
the  second-row  hydrides  and  fluorides. 

16-4.1    The  Shapes  of  H,0  and  FtO 

The  orbital  representations  of  the  bonding  in 
H20  and  in  F20  suggest  that  two  p  orbitals  of 
oxygen  are  involved  in  the  bonding  [see  repre- 
sentation (24)].  Figure  16-9  shows  the  spatial 
arrangement  we  assign  to  the  p  orbitals  (assum- 
ing they  are  like  hydrogen  atom  orbitals).  If  the 


to  the  ball-and-stick  model  of  the  molecule.  A 
line  is  drawn  between  the  oxygen  atom  and  each 
hydrogen  atom  to  indicate  that  a  chemical  bond 
holds  these  two  atoms  together.  No  line  is  drawn 
between  the  two  hydrogen  atoms  since  we  feel 
they  are  not  directly  bonded  to  each  other.  Now 
we  can  apply  our  discussion  of  the  role  of  elec- 
trons in  bonding  to  add  to  the  meaning  of  this 
line  representation.  A  line  is  drawn  between  two 
atoms  to  indicate  that  a  pair  of  electrons  is 
shared  between  these  two  atoms,  resulting  in  a 
chemical  bond. 


Fig.  16-9.  The  expected  shape  of  the  HJO  molecule: 
p*  bonding. 


spatial  arrangement  persists  after  the  bonds 
form,  the  molecular  shape  would  be  fixed,  as 
shown.  The  molecule  would  be  bent,  with  an 
angle  near  90°.  The  same  would  be  true  for  F20. 
The  measured  bond  angles  are  as  follows: 


F    ^^    F 


ZH-0-H  =  1O4.S0 
(52a) 

ZF-O-F  m  202  • 

(52b) 

It  is  generally  true  that  a  divalent  atom  with 
twop  orbitals  as  valence  orbitals  forms  an  angular 
molecule.  Since  this  prediction  is  reliable,  the 
bonding  is  usually  characterized  by  identifying 
the  valence  orbitals.  Oxygen  is  said  to  use  p2 
(read,  "p  two")  bonding  in  water  and  FjO. 

Notice  that  the  structural  formulas  (52a)  and 
(52b)  use  another  representation  of  the  bonding. 
It  is  quite  familiar,  of  course,  since  it  corresponds 


16-4.2    The  Shapes  of  NH,  and  NF, 

In  NH3  and  NF3,  three  p  orbitals  are  involved  in 
the  bonding  [see  representation  (30)].  Figure 
16-10  shows  the  spatial  arrangement  implied  by 
assuming  persistence  of  the  hydrogen  atom  or- 
bitals after  bonding.  We  expect,  then,  that  am- 


Fig.  16-10.  The  expected  shape  of  the  NH,  molecule: 
p*  bonding. 


monia  has  a  pyramidal  shape  (a  pyramid  with 
a  three-sided  base).  The  bond  angles  should  be 
near  90°.  Both  NH3  and  NF3  do  have  pyramidal 
shapes.  The  measured  bond  angles  are  as  fol- 
lows: 

Hz:—f-.._7^a  ZH-K-H-  107° 

(53a) 


F-"" 


ZF-N-F  =  102* 
(53b) 


292 


MOLECULES    IN    THE    GAS    PHASE   I    CHAP.    16 


Again,  experiments  show  that  it  is  generally 
true  that  a  trivalent  atom  with  three  p  orbitals  as 
valence  orbitals  forms  a  pyramidal  molecule.  The 
bonding  is  called  pz  bonding  (read,  "/?  three"). 

16-4.3    The  Shapes  of  CH<  and  CF4 

The  bonding  of  methane,  CH4,  and  that  of  car- 
bon tetrafluoride,  CF4,  involve  four  valence  or- 
bitals, the  2s  orbital  and  the  three  2p  orbitals. 
Four  bonds  are  formed  by  carbon  and,  as  before, 
we  characterize  the  bonding  by  naming  the  va- 
lence orbitals:  sp%  (read,  "sp  three").  This  time, 
however,  the  assumption  of  the  persistence  of 
the  spatial  distributions  of  the  hydrogen  atom 
orbitals  does  not  indicate  directly  what  bond 
angles  to  expect.  Experiment  shows,  however, 
that  spz  bonding  always  gives  bond  angles  which 
are  exactly  or  very  close  to  tetrahedral  angles. 
This  means  that  the  angle  between  any  two 
carbon-hydrogen  bonds  is  109°28'.  The  structure 
is  called  tetrahedral  because  the  four  hydrogen 
atoms  occupy  the  positions  of  the  corners  of  a 
regular  tetrahedron  (a  four-sided  figure  with 
equal  edges).  The  structure  is  shown  in  Figure 
16-11. 

16-4.4    The  Shape  of  BF, 

The  boron  atom  in  BF3  uses  the  2s  and  two  2p 
orbitals  in  bonding.  Therefore  the  bonding  is 


Fig.  16-11.  The   tetrahedral   bonding   of  carbon:  sp* 
bonding. 


Fig.  16-12.  The  structure  of  BF3:  sp-  bonding. 


called  sp"1.  Again  we  must  let  experiment  tell  us 
the  bond  angles  which  are  found  with  sp2  bond- 
ing. The  structure  of  BF3  is  that  of  an  equilateral 
triangle.  The  structure,  shown  in  Figure  16-12, 
is  planar  and  each  of  the  three  fluorine  atoms  is 
the  same  distance  from  the  boron  atom  as  are 
the  others. 


16-4.5    The  Shape  of  BeF; 

Beryllium  atom  in  gaseous  BeF2  uses  the  2s  and 
only  one  2p  orbital  in  bonding.  The  bonding  is 
called  sp.  Experiment  shows  that  the  molecule  is 
linear  and  symmetrical,  as  shown  in  Figure 
16-13.  The  structure  of  gaseous  BeH2  is  un- 
doubtedly linear  and  symmetric  as  well,  by 
analogy  to  BeF2. 


Fig.  16-13.  The  structure  of  BeF2:  sp  bonding. 


-h£>- 


SEC.     16-4    I    MOLECULAR    ARCHITECTURE 


293 


16-4.6    Summary  of  Bonding  Orbitals 
and  Molecular  Shape 

From  the  data  presented  here,  the  orbitals  in- 
volved in  bonding  correlate  with  the  molecular 
architecture.  The  relationships  are  summarized 
in  Table  16-IV. 


THE    MOLECULAR    D1POLE   OF    LlF 

The  lithium  fluoride  bond  is  highly  ionic  in  char- 
acter because  of  the  large  difference  in  ionization 
energies  of  lithium  and  fluorine.  Consequently, 
gaseous  lithium  fluoride  has  an  unusually  high 
electric  dipole. 


Table  16-IV.     bonding  orbitals,  bonding  capacity,  and  molecular   shape 


ELEMENT 


BONDING 
ORBITALS 


BONDING 
CAPACITY 


MOLECULAR  SHAPE 
OF  THE  FLUORIDE 


EXAMPLE 


He 

none 

0 

monatomic 

He 

Li 

s 

1 

linear,  diatomic  molecule 

LiF 

Be 

sp 

2 

linear 

BeF, 

B 

sp2 

3 

planar,  triangular 

BF, 

C 

sp3 

4 

tetrahedral 

CF« 

N 

p* 

3 

pyramidal 

NF, 

O 

P2 

2 

bent 

OF, 

F 

P 

1 

linear,  diatomic  molecule 

F2 

Ne 

none 

0 

monatomic 

Ne 

16-4.7     Molecular  Shape  and  Electric  Dipoles 

Consider  the  fluorides  of  the  second-row  ele- 
ments. There  is  a  continuous  change  in  ionic 
character  of  the  bonds  fluorine  forms  with  the 
elements  F,  O,  N,  C,  B,  Be,  and  Li.  The  ionic 
character  increases  as  the  difference  in  ionization 
energies  increases  (see  Table  16-11).  This  ionic 
character  results  in  an  electric  dipole  in  each 
bond.  The  molecular  dipole  will  be  determined 
by  the  sum  of  the  dipoles  of  all  of  the  bonds, 
taking  into  account  the  geometry  of  the  mole- 
cule. Since  the  properties  of  the  molecule  are 
strongly  influenced  by  the  molecular  dipole,  we 
shall  investigate  how  it  is  determined  by  the 
molecular  architecture  and  the  ionic  character 
of  the  individual  bonds.  For  this  study  we  shall 
begin  at  the  left  side  of  the  periodic  table. 


THE    MOLECULAR    DIPOLE    OF    BeF2 

The  beryllium-fluorine  bond  is  also  highly  ionic 
in  character.  However,  there  are  two  such  Be-F 
bonds  and  the  electrical  properties  of  the  entire 
molecule  depend  upon  how  these  two  bonds  are 
oriented  relative  to  each  other.  We  must  find 
the  "geometrical  sum"  of  these  two  bond  dipoles. 
The  geometrical  sum  of  two  arrows  can  be 
understood  simply  with  the  aid  of  Figure  16-14. 
Figure  16-14A  shows  how  two  arrows  pointing 
in  the  same  direction  combine  to  give  a  longer 
arrow.  Figure  16-14B  shows  how  two  arrows 
oppositely  directed  combine  to  give  a  shorter 


Fig. 


16-14.  The    geometrical    sum    of    dipoles:    both 

length  and  direction  are  important. 


+ 


combine    to   fftve 


+ 


combine    -to     give 


+ 


\ 


combine     -to     give 


\ 


294 


MOLECULES    IN    THE    GAS    PHASE    I    CHAP.     16 


Fig.  16-15.  The   absence   of   a   molecular   dipole    in 
BeFt. 


arrow.  Figure  16-14C  shows  how  two  arrows 
that  are  not  parallel  add  to  give  an  arrow  in  a 
new  direction. 

Now  we  can  apply  the  process  of  combination 
shown  in  Figure  16-14  to  BeF2.  In  the  linear, 
symmetric  BeF2  molecule,  the  two  bond  dipoles 
point  in  opposite  directions.  Since  the  two  bonds 
are  equivalent,  there  is  a  complete  cancellation, 
as  shown  in  Figure  16-15.  Hence  the  molecule 
has  no  net  dipole;  the  molecular  dipole  is  zero. 

THE    MOLECULAR    DIPOLES 
OF    BF3    AND    CF4 

Both  of  these  molecules  are  thought  to  have 


=      O 

zero.  Careful  consideration  of  the  geometry 
shows  that  there  is  a  complete  cancellation  by 
the  bond  dipoles  in  each  molecule.  This  cancel- 
lation is  shown  in  Figure  16-16  for  BF3.  The 
molecular  dipole  is  zero. 


THE    MOLECULAR    DIPOLE    IN    F20 

Since  F20,  with  p-  bonding,  is  a  bent  molecule, 
the  two  bond  dipoles  do  not  cancel  each  other 
as  they  do  in  BeF2.  On  the  other  hand,  the  ioni- 
zation energies  of  oxygen  and  fluorine  are  not 
very  different,  so  the  electric  dipole  of  each  bond 
is  small  in  magnitude.  These  add  together,  ac- 
cording to  the  geometry,  to  give  a  polar  mole- 
cule, as  shown  in  Figure  16-17. 


moderate  amounts  of  ionic  character  in  each     Fig.  16-16.  The  absence  of  a  molecular  dipole  in 

bond.  Yet  the  molecular  dipoles  are  each  exactly  fiF,. 


^^    J^ 


+  /*< 


—  bd  =  o 


SEC.    16-5    I    DOUBLE    BONDS 


295 


•V 


•s  =  ss> 


I 


16-5    DOUBLE  BONDS 

In  deciding  the  bonding  capacity  of  a  given  atom 
from  the  second  row,  we  have  counted  the  num- 
ber of  hydrogen  atoms  or  fluorine  atoms  with 
which  it  would  combine.  Thus  oxygen  combines 
with  two  hydrogen  atoms  to  form  water,  H20. 
Oxygen  is  said  to  be  divalent.  Oxygen  shares  two 
pairs  of  electrons,  one  pair  with  each  hydrogen 
atom.  Each  of  these  shared  pairs  forms  a  single 
bond. 

16-5.1    Bonding  in  the  Oxygen  Molecule 

Now  let  us  investigate  the  oxygen  molecule, 
which  experiment  tells  us  has  the  molecular  for- 
mula 02.  We  might  begin  by  considering  forma- 
tion of  a  single  bond  between  two  oxygen  atoms, 
as  represented  by  the  orbital  representation 


Is 


2s 


2P 

<8>    <8ES> 


*   <8>        ® 


(54) 


We  see  that  each  oxygen  atom  has  residual  bond- 
ing capacity.  Each  atom  could,  for  example, 
react  with  a  hydrogen  atom  to  form  hydrogen 
peroxide,  as  shown  in  electron  dot  representa- 
tion (26).  Each  oxygen  atom  could  react  with  a 
fluorine  atom  to  form  F202.  In  short,  each  oxy- 
gen atom  is  in  need  of  another  atom  with  an 
electron  in  a  half-filled  valence  orbital  so  that  it 
can  act  as  a  divalent  atom. 


Fig.   16-17.  The  molecular  dipole  of  FaO. 


But  suppose  oxygen  can  find  no  hydrosen 
atoms  or  fluorine  atoms.  Then,  it  does  the  lazy 
thing:  the  two  atoms,  already  bound  by  one 
bond,  form  a  second  bond  with  each  other.  The 
result  might  be  shown  in  the  orbital  representa- 
tion (55). 

**  2s  2P 


0    <8>        (8)    ®SQ 
0    <8>        <8>    <800 


(55) 


There  is  strong  evidence  supporting  this  pro- 
posal. The  bond  in  the  oxygen  molecule  is 
stronger  than  the  oxygen-oxygen  bond  in  hy- 
drogen peroxide  (more  energy  is  required  to 
break  it).  The  vibrational  frequency  of  the  oxy- 
gen molecule  is  higher  than  that  of  a  normal 
single  bond,  showing  that  there  is  additional 
bonding  (see  Section  14-3.4).  The  bond  length 
in  the  02  molecule  is  1.21  A.  In  the  gaseous 
hydrogen  peroxide  molecule,  the  oxygen-oxygen 
distance  is  1.48A.  The  short  bond  length  in  the 
02  molecule  shows  that  the  two  oxygen  atoms 
are  drawn  together  more  effectively  than  in 
HOOH,  suggesting  there  are  extra  bonding  elec- 
trons in  02. 

Because  all  of  the  evidence  we  have  examined 
is  consistent  with  the  orbital  representation  (55), 


296 


MOLECULES    IN    THE    GAS    PHASE    I    CHAP.     16 


the  bond  in  02  is  called  a  double  bond.  An  elec- 
tron dot  representation  can  be  written  as  follows 


:o:  :o: 


(56) 


The  representation  (56)  shows  two  pairs  of  elec- 
trons shared.  Each  oxygen  atom  finds  itself  near 
eight  electrons.  There  is,  on  the  one  hand,  a 
stable  molecule,  because  all  of  the  bonding 
capacity  of  each  oxygen  atom  is  in  use.  On  the 
other  hand,  this  special  aspect  of  the  bonding  of 
oxygen  undoubtedly  contributes  to  the  reactivity 
of  oxygen. 

16-5.2    Ethylene:  A  Carbon-Carbon 
Double  Bond 

Ethylene  is  a  simple  compound  of  carbon  and 
hydrogen  with  the  formula  C2H4.  Thus  it  has  two 
less  hydrogen  atoms  than  does  ethane,  C2H6. 
This  means  that  to  write  a  structure  of  ethylene 
we  must  take  account  of  two  electrons  that  are 
not  used  in  C — H  bond  formation.  Suppose  we 
write  an  electron  dot  representation  involving 
only  single  bonds 

H  H 
H:C:C:H  (57) 

This  formula  has  two  unpaired  electrons,  repre- 
senting unused  bonding  capacity.  This  objection- 
able situation  can  easily  be  rectified  by  allowing 
the  two  unpaired  electrons  to  pair,  and  thus  form 
an  additional  two-electron  bond.  Now  the  car- 
bon atoms  are  joined  by  a  double  bond,  just  as 
the  oxygen  atoms  in  02  are  double  bonded  to 
each  other 


H  .  .  H 

:c: :c: 

H  H 


(58) 


CHEMICAL   REACTIVITY   OF   ETHYLENE 

In  ethane,  C2H6,  all  of  the  bonds  are  normal 
single  bonds.  Experiment  shows  that  ethane  is  a 
fairly  unreactive  substance.  It  reacts  only  when 
treated  with  quite  reactive  species  (such  as  free 
chlorine  atoms),  or  when  it  is  raised  to  excited 
energy  states  by  heat  (as  in  combustion). 
Ethylene,  on  the  other  hand,  reacts  readily 


with  many  chemical  reagents.  Having  four  elec- 
trons forming  the  carbon-carbon  bond,  the  elec- 
trons of  the  double  bond  seem  to  be  accessible 
to  attack.  We  find  that  the  typical  reactions  of 
ethylene  are  those  with  reagents  that  seek 
electrons.  For  example,  oxidizing  agents  are 
electron-seeking,  and  we  would  expect  the  dou- 
ble bond  to  be  readily  oxidized.  This  is  indeed 
the  case.  Ethylene  will  reduce  (that  is,  be  oxidized 
by)  such  oxidizing  agents  as  potassium  perman- 
ganate or  potassium  dichromate  at  ordinary 
temperatures.  Under  these  same  mild  conditions 
ethane  is  completely  unreactive  to  the  same  re- 
agents. 

GEOMETRICAL    FEATURES    OF    ETHYLENE 

The  shape  of  the  ethylene  molecule  has  been  learned  by  a 
variety  of  types  of  experiments.  Ethylene  is  a  planar 
molecule — the  four  hydrogen  and  the  two  carbon  atoms 
all  lie  in  one  plane.  The  implication  of  this  experimental 
fact  is  that  there  is  a  rigidity  of  the  double  bond  which 
prevents  a  twisting  movement  of  one  of  the  CH:  groups 
relative  to  the  other.  Rotation  of  one  CH2  group  relative 
to  the  other — with  the  C — C  bond  as  an  axis — must  be 
energetically  restricted  or  the  molecule  would  not  retain 
this  flat  form. 

CIS-TRANS    ISOMERISM    OF    ETHYLENE 
DERIVATIVES 

It  is  possible  to  replace  hydrogen  atoms  of  ethylene  by 
halogen  atoms.  For  example,  one  such  compound  has  the 
formula  C2H2C12.  In  preparing  such  a  compound,  chem- 
ists discovered  long  ago  that  they  could  obtain  three  dif- 
ferent pure  substances  with  this  same  formula,  QH2Q2. 
Different  compounds  with  the  same  molecular  formulas  are 
called  isomers.  The  existence  of  three  separate  QH2CI2 
isomers  is  readily  explained  in  terms  of  the  molecular 
geometry.  The  three  structures  possible  with  this  formula 
are  shown  in  Figure  16-18.  They  are  all  called  dichloro- 
ethylene. 

Two  types  of  isomerism  are  involved.  Formula  (59) 
differs  from  (60)  and  (61).  In  formula  (59)  both  chlorine 
atoms  are  attached  to  the  same  carbon  atom.  In  both 

(60)  and  (61)  there  is  one  chlorine  atom  attached  to  each 
carbon  atom.  The  difference  between  (59)  and  the  other 
pair,  (60)  and  (61),  is  indicated  by  calling  these  molecules 
structural  isomers. 

The  pair  of  isomers  (60)  and  (61)  differ  in  another  way. 
Though  each  has  one  chlorine  atom  attached  to  each 
carbon  atom,  in  (60)  they  are  on  the  "same  side"  of  the 
double  bond.  This  relationship  is  called  the  cis  form.  In 
(67)  the  chlorine  atoms  are  across  from  each  other.  This 
relationship  is  called  the  trans  form.  Formulas  (60)  and 

(61)  identify  cis  and  trans  isomers  of  dichioroethylene. 


QUBSTIONS    AND    PROBLEMS 


297 


Fig.  16-18.  The  isomers  of  dichloroethylene. 


Experiment  shows  that  it  is  exceedingly  difficult  to 
convert  (59)  into  (60)  or  (61).  To  make  such  a  conversion, 
bonds  must  be  broken  and  reformed.  Such  reactions  are 
almost  always  quite  slow  because  the  activation  energies 


must  be  almost  as  large  as  the  energies  of  the  bonds  being 
broken.  In  contrast,  the  conversion  of  (60)  into  (61)  (or 
the  reverse)  can  be  accomplished  merely  by  heating  the 
substance.  No  bonds  must  be  broken  completely — only 
a  rotation  around  the  carbon-carbon  double  bond  is 
necessary.  This  process  has  a  much  lower  activation 
energy  and  the  reaction  occurs  at  moderate  temperatures. 


QUESTIONS  AND  PROBLEMS 


1.  Which  one  of  the  following  statements  is  FALSE 
as  applied  to  this  equation  ? 


Wg) 


H(g)  +  H(g)         AH  =  103.4  kcal 


(a)  The  positive  AH  means  the  reaction  is  endo- 
thermic. 

(b)  Two  grams  of  H(g)  contain  more  energy 
than  2  grams  of  H2(g). 

(c)  Weight  for  weight,  H(g)  would  be  a  better 
fuel  than  H2(g). 


(d)  The  spectrum  of  H2(g)  is  the  same  as  the 
spectrum  of  H(g). 

2.  What  are  the  molecular  species  present  in  gase- 
ous neon,  argon,  krypton,  and  xenon?  Explain. 

3.  Determine  the  number  of  attractive  forces  and 
the  number  of  repulsive  forces  in  LiH. 

4.  What  energy  condition  must  exist  if  a  chemical 
bond  is  to  form  between  two  approaching 
atoms  ? 


298 


MOLECULES    IN    THE    GAS    PHASE   I    CHAP.     16 


5.  What  valence  orbital  and  valence  electron  condi- 
tions must  exist  if  a  chemical  bond  is  to  form 
between  two  approaching  atoms  ? 

6.  Give  the  orbital  and  also  the  electron  dot  repre- 
sentations for  the  bonding  in  these  molecules: 
Cl2,  HG1,  C120. 

7.  Using  the  electron  dot  representation,  show  a 
neutral,  a  negatively  charged,  and  a  positively 
charged  OH  group. 

8.  Draw  the  orbital  representation  of  the  molecule 
N2H4,  hydrazine. 

9.  Knowing  the  orbitals  carbon  uses  for  bonding, 
use  the  periodic  table  to  predict  the  formula  of 
the  chloride  of  silicon.  What  orbitals  does  silicon 
use  for  bonding? 

10.  Draw  the  orbital  representations  of 

(a)  sodium  fluoride, 

(b)  beryllium  fluoride,  BeF2. 

11.  In  general,  what  conditions  cause  two  atoms  to 
combine  to  form: 

(a)  a  bond  that  is  mainly  covalent; 

(b)  a  bond  that  is  mainly  ionic; 

(c)  a  polar  molecule? 

12.  What  type  of  bonding  would  you  expect  to  find 
in  MgO?  Explain. 


13.  Considering  comparable  oxygen  compounds, 
predict  the  shape  of  H2S  and  H2S2  molecules. 
What  bonding  orbitals  are  used? 

14.  Predict  the  formula  and  molecular  shape  of  a 
hydride  of  phosphorus. 

15.  Draw  an  electron  dot  representation  for  the 
NH4"  ion.  What  shape  do  you  predict  this  ion 
will  have? 

16.  Predict  the  type  of  bonding  and  the  shape  of  the 
ion  BF^~ . 

17.  Consider  the  two  compounds  CH3CH3  (ethane) 
and  CH3NH2  (methylamine).  Why  does  CH3NH2 
have  an  electric  dipole  while  CH3CH3  does  not  ? 

18.  Consider  the  following  series:  CH4,  CH3C1, 
CH2C12,  CHCla,  CCL.  In  which  case(s)  will  the 
molecules  have  electric  dipoles?  Support  your 
answer  by  considering  the  bonding  orbitals  of 
carbon,  the  molecular  shape  of  the  molecules, 
and  the  resulting  symmetry. 

19.  Predict  the  structure  of  the  compound  N2Fj  from 
the  electron  dot  representation  of  the  atoms  and 
the  molecule. 

20.  Which  of  the  isomers  of  dichloroethylene  shown 
in  Figure  16-18  will  be  polar  molecules? 

21.  Draw  structural  formulas  for  all  the  isomers  of 
ethylene  (C2H4)  in  which  two  of  the  hydrogen 
atoms  have  been  replaced  by  deuterium  atoms. 
Label  the  cis  and  the  trans  isomers. 


LINUS    C.    PAULING,    19  0  1 


No  other  living  chemist  has  contributed  more  to  our  under- 
standing of  chemical  bonding  than  Linus  C.  Pauling.  His 
ideas  pervade  every  aspect  of  chemistry.  These  ideas  have 
won  him  some  seventeen  medals  and  high  awards,  including 
the  1954  Nobel  Prize  in  Chemistry.  His  international  re- 
nown is  bespoken  by  his  election  to  honorary  membership 
in  sixteen  scientific  societies  in  ten  different  countries. 

Linus  Pauling  was  born  in  Portland,  Oregon,  and  his 
hobbies  as  a  boy  were  largely  scientific.  At  11  he  began  an 
insect  collection,  which  led  to  reading  books  on  entomology. 
At  13  Linus  found  a  chemistry  book  in  the  family  library 
and  founded  a  laboratory  in  the  family  basement.  By  the 
time  he  entered  Oregon  State  College  he  was  determined 
to  become  a  chemical  engineer.  In  1922  Pauling  received 
the  B.S.  degree,  and  this  was  followed  by  the  Ph.D.  at 
California  Institute  of  Technology.  By  now  his  interest  had 
turned  to  the  fundamental  aspects  of  chemistry  and  after  a 
year  of  post-doctoral  s.udy  in  Europe  he  returned  to  the 
faculty  at  California  Institute  of  Technology.  There  he 
established  and  pursued  his  illustrious  career. 

Linus  Pauling's  prodigious  scientific  productivity  has 
broadly  influenced  the  face  of  chemistry.  His  interest  fo- 
cused on  the  chemical  bond  and  he  was  one  of  the  earliest 
of  chemists  to  recognize  the  importance  of  the  quantum 
mechanical  point  of  view.  He  gave  quantitative  meaning  to 
the  electronegativity  concept.  He  discussed  the  mixing  of 
the  ionic  and  covalent  character  of  chemical  bonds  and 
introduced  the  term  "resonance" — a  concept  that  manv 


chemists  criticize  but  that  most  chemists  use  regularly. 
Pauling  gave  detailed  consideration  to  the  effective  sizes 
of  atoms  in  molecules  and  crystals.  He  became  an  authority 
on  hydrogen  bonding  and  he  proposed  a  theory  of  metallic 
bonding.  He  advocated  and,  with  his  colleagues,  established 
the  existence  of  helical  structures  in  proteins.  And,  while 
publishing  over  300  papers,  he  authored  several  books  that 
won  wide  acceptance:  Introduction  to  Quantum  Me- 
chanics (with  E.  B.  Wilson,  Jr.),  The  Nature  of  the 
Chemical  Bond,  General  Chemistry,  and  College  Chem- 
istry. 

Pauling  has  a  deep  sensitivity  to  the  welfare  of  mankind 
and  he  has  worked  energetically  to  awaken  the  conscience 
of  society  to  its  new  responsibilities  in  a  nuclear  age.  At- 
tempting to  inform  the  public  of  the  pressing  need  for 
lasting  peace,  he  has  written  a  book  entitled  No  More  War. 
He  has  engaged  in  many  public  forums  and  debates,  even 
though  these  activities  sometimes  attracted  ridicule  in  times 
when  fear  made  his  cause  an  unpopular  one.  In  recognition 
of  these  activities  he  was  awarded  the  1962  Nobel  Peace 
Prize,  and  thus  became  the  second  person  in  history  to 
receive  the  Nobel  Prize  twice. 

There  is  no  chemistry  course  given  today  that  is  not  in- 
fluenced by  the  ideas  of  Linus  C.  Pauling.  He  is  a  man  of 
broad  imagination,  dramatic  personality,  and  boundless 
inspiration.  Mankind  will  long  benefit  because  he  chose  to 
explore  the  frontiers  of  science. 


CHAPTER 


17 


The  Bonding  in 
Solids  and  Liquids 


//  is  •  '  '  possible  to  discuss  the  structure  of  any  substance  •  •  •  by 
describing  the  types  •  •  •  of  its  bonds  and  in  this  way  to  account  for  its 
characteristic  properties. 

LINUS     PAULING,     1939 


Any  pure  gas,  when  cooled  sufficiently,  will  con- 
dense to  a  liquid  and  then,  at  a  lower  tempera- 
ture, will  form  a  solid.  There  is  great  variance 
in  the  temperature  at  which  this  condensation 
occurs.  Apparently  there  is  a  corresponding  vari- 
ance of  the  forces  in  liquids  and  solids.  For 
example,  lithium  fluoride  gas  at  one  atmosphere 
pressure  condenses  when  cooled  below  1949°K. 
When  the  temperature  is  lowered  to  1 143°K,  the 
liquid  forms  a  clear  crystal.  In  contrast,  lithium 
gas  at  this  pressure  must  be  cooled  to  1599°K 
before  it  forms  a  liquid  and  this  liquid  does  not 
solidify  until  the  temperature  reaches  453°K.  The 
solid  is  a  white,  soft  metal,  not  resembling 
crystalline  lithium  fluoride  at  all.  Fluorine  gas  is 
equally  distinctive.  At  one  atmosphere  pressure 
it  must  be  cooled  far  below  room  temperature 
before  condensation  occurs,  at  85°K.  Then  the 
liquid  solidifies  to  a  crystal  at  50°K.  Why  do 
these  three  materials  behave  so  differently?  Can 
we  understand  this  great  variation?  Let  us  begin 
by  finding  a  common  point  of  departure. 

Two  or  more  atoms  remain  near  each  other 
in  a  particular  arrangement  because  the  energy 
300 


favors  that  arrangement.  This  is  true  whether 
the  cluster  of  atoms  is  strongly  or  weakly  bound, 
whether  it  contains  a  few  atoms  or  1023  atoms, 
whether  the  arrangement  is  regular  (as  in  a 
crystal)  or  irregular  (as  in  a  liquid).  The  cluster 
of  atoms  is  stable  if  and  only  if  the  energy  is 
lower  when  the  atoms  are  together  than  when 
they  are  apart. 

Furthermore,  there  is  but  one  reason  that  two 
or  more  atoms  have  lower  energy  when  they  are 
in  proximity.  In  this  way  electrons  can  be  close 
to  two  or  more  positive  nuclei  simultaneously. 
However,  the  magnitude  of  the  attractive  forces 
varies  greatly,  depending  on  how  close  the  elec- 
trons are  able  to  approach  these  positive  nuclei. 
This  approach  distance  is  fixed  by  the  electron 
occupancy  of  the  valence  orbitals. 

Thus  the  occupancy  of  the  valence  orbitals  is 
the  clue  we  shall  follow  in  our  attempt  to  predict 
when  to  expect  a  substance  to  be  a  high-melting, 
salt-like  crystal,  when  to  expect  a  metal,  when 
to  expect  a  low-melting,  molecular  crystal.  This 
is  an  ambitious  program.  Let's  see  how  far  we 
can  go,  beginning  with  the  pure  elements. 


SEC.    17-1    I    THE    ELEMENTS 


301 


17-1    THE  ELEMENTS 


The  examples  just  mentioned  include  two  ele- 
ments, fluorine  and  lithium.  Fluorine  forms  a 
weakly  bound  molecular  solid.  Lithium  forms  a 
metallic  solid.  Let  us  see  how  we  can  account  for 
this  extreme  difference,  applying  the  principles 
of  bonding  treated  in  Chapter  16. 

17-1.1    van  der  Waals  Forces 

The  diatomic  molecule  of  fluorine  does  not  form 
higher  compounds  (such  as  F3,  F4,  •  •  •)  because 
each  fluorine  atom  has  only  one  partially  filled 
valence  orbital.  Each  nucleus  in  F2  is  close  to  a 
number  of  electrons  sufficient  to  fill  the  valence 
orbitals.  Under  these  circumstances,  the  diatomic 
molecule  behaves  like  an  inert  gas  atom  toward 
other  such  molecules.  The  forces  that  cause 
molecular  fluorine  to  condense  at  85°K  are, 
then,  the  same  as  those  that  cause  the  inert  gases 
to  condense.  These  forces  are  named  van  der 
Waals  forces,  after  the  Dutch  scientist  who 
studied  them. 

When  the  outer  orbitals  of  all  of  the  atoms  in 
the  molecules  are  filled — giving  inert  gas  con- 
figurations— then  the  electrons  of  another  mole- 
cule cannot  approach  the  nuclei  closely.  When 
molecules  of  this  sort  approach  each  other,  the 
energy  is  lowered  only  a  few  tenths  of  a  kilo- 
calorie  per  mole.  This  weak  interaction  is  typical 
of  van  der  Waals  forces. 

We  have,  now,  a  simple  rule  for  predicting 


when  a  weakly  bound  molecular  liquid  and  a 
low-melting  crystal  will  be  formed  by  a  given 
element.  If  the  element  forms  a  molecule  that 
gives  each  atom  the  orbital  occupancy  of  an  inert 
gas,  then  only  van  der  Waals  interactions  among 
such  molecules  remain. 

We  shall  take  up  later  in  this  chapter  the  fac- 
tors that  determine  the  magnitude  of  van  der 
Waals  forces.  For  the  moment,  we  will  merely 
observe  that  the  elements  forming  van  der  Waals 
liquids  and  solids  are  concentrated  in  the  upper 
right-hand  corner  of  the  periodic  table  (see  Fig- 
ure 17-1).  These  are  the  elements  able  to  form 
stable  molecules  that  satisfy  completely  the 
bonding  capacity  of  each  atom. 


EXERCISE  17-1 

Gaseous  phosphorus  is  made  up  of  P4  molecules 
with  four  phosphorus  atoms  arranged  at  the 
corners  of  a  regular  tetrahedron.  In  such  a 
geometry,  each  phosphorus  atom  is  bound  to 
three  other  phosphorus  atoms.  Would  you  ex- 
pect this  gas  to  condense  to  a  solid  with  a  low 
or  high  melting  point?  After  making  a  prediction 
on  the  basis  of  the  valence  orbital  occupancy, 
check  the  melting  point  of  phosphorus  in  Table 
6-V1II,  p.  101. 


Fig.  17-1.  Elements    that    form    molecular    crystals 

bound  by  van  der  Waals  forces. 


I« 

He 

*z 

o2 

Px 

Ne 

p+ 

Ss 

Clz 

Ar 

Brz 

Kr 

I* 

Xe 

Air, 

Rn 

1            1            I            1 

1                1                1                I 

.                    1 

302 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS   I    CHAP.    17 


Fig.  17-2.  Carbon    forms   network   solids:    diamond 
and  graphite. 


17-1.2    Covalent  Bonds  and  Network  Solids 

Fluorine,  F2,  oxygen,  02,  and  nitrogen,  N2,  all 
form  molecular  crystals  but  the  next  member  of 
this  row  of  the  periodic  table,  carbon,  presents 
another  situation.  There  does  not  seem  to  be  a 
small  molecule  of  pure  carbon  that  consumes 
completely  the  bonding  capacity  of  each  atom. 
As  a  result,  it  is  bound  in  its  crystal  by  a  network 
of  interlocking  chemical  bonds. 

With  one  2s  and  three  2p  orbitals  available  for 
the  bonding  of  carbon,  we  can  expect  it  to  form 
a  lattice  in  which  each  atom  forms  four  bonds. 


h, 


Furthermore,  sp*  bonding  is  connected  with 
tetrahedral  bond  angles  (as  in  Figure  16-11). 
These  expectations  are  consistent  with  the  ex- 
perimentally determined  structure  of  diamond, 
shown  in  Figure  17-2. 

Diamond  is  a  naturally  occurring  form  of 
pure,  crystalline  carbon.  Each  carbon  atom  is 
surrounded  by  four  others  arranged  tetrahe- 
drally.  The  result  is  a  compact  structural  network 
bound  by  normal  chemical  bonds.  This  descrip- 
tion offers  a  ready  explanation  for  the  extreme 
hardness  and  the  great  stability  of  carbon  in  this 
form. 


Fig.  17-3.  Elements  that  form  solids  involving  cova- 
lent bonding. 


B 


Si 


(re 


N, 


As 


Sh 


Bi 


Se 


Te 


r<> 


CI, 


Br, 


Atz 


He 


Ne 


At 


Kr 


Xe 


Rn 


SEC.    17-1    I    THE    ELEMENTS 


303 


Graphite  is  another  solid  form  of  carbon.  In 
contrast  to  the  three-dimensional  lattice  struc- 
ture of  diamond,  graphite  has  a  layered  structure. 
Each  layer  is  strongly  bound  together  but  only 
weak  forces  exist  between  adjacent  layers.  These 
weak  forces  make  the  graphite  crystal  easy  to 
cleave,  and  explain  its  softness  and  lubricating 
qualities. 

The  elements  that  form  network  solids  lie  on 
the  right  side  of  the  periodic  table,  bordering 
the  elements  that  form  molecular  crystals  on 
one  side  and  those  that  form  metals  on  the  other. 
Thus  they  are  intermediate  between  the  metals 
and  the  nonmetals.  In  this  borderline  region 
classifications  are  sometimes  difficult.  Whereas 
one  property  may  suggest  one  classification,  an- 
other property  may  lead  to  a  different  conclu- 
sion. Figure  17-3  shows  some  elements  that  form 
solids  that  are  neither  wholly  metallic  nor  wholly 
molecular  crystals. 


17-1.3    Metallic  Bonding 

We  have  considered  solid  forms  of  the  elements 
fluorine,  oxygen,  nitrogen,  and  carbon.  In  each 
case,  a  solid  is  formed  in  which  the  bonding 
capacity  is  completely  satisfied.  The  remaining 
elements  of  the  second  row.  that  is,  beryllium, 


H, 


Li 


Na 


Be 


My 


and  lithium,  are  metallic.  These  elements  do  not 
have  enough  electrons  to  permit  the  complete 
use  of  the  valence  orbitals  in  covalent  bonding. 
Furthermore,  the  ionization  energies  of  these 
elements  are  quite  low.  We  find  there  are  two 
conditions  necessary  for  metallic  bonding: 
vacant  valence  orbitals  and  low  ionization  en- 
ergies. 

CHARACTERISTIC    PROPERTIES 
OF   METALS 

Perhaps  the  most  obvious  metallic  property  is 
reflectivity  or  luster.  With  few  exceptions  (gold, 
copper,  bismuth,  manganese)  all  metals  have  a 
silvery  white  color  which  results  from  reflecting 
all  frequencies  of  light.  We  have  said  previously 
that  the  electron  configuration  of  a  substance 
determines  the  way  in  which  it  interacts  with 
light.  Apparently  the  characteristic  reflectivity  of 
metals  indicates  that  all  metals  have  a  special 
type  of  electron  configuration  in  common. 

A  second  characteristic  property  of  metals  is 
high  electrical  conductivity.  The  conductivity  is 
so  much  higher  than  that  of  aqueous  electrolyte 
solutions  that  the  charge  movement  cannot  in- 
volve the  same  mechanism.  Again  we  find  a 

Fig.  17-4.  The  metallic  elements. 


At 


Si 


AT, 


S* 


CI, 


He 


Ne 


Ar 


K 


Co. 


Sc 


Ti 


V 


Cr 


Mn 


Fe 


Co 


/Si 


Cu 


Zn 


Go. 


&e 


As 


Se 


Br2 


Kr 


Rb 


Sr 


Zr 


Nb 


Mo 


Tc 


Ru 


Rh 


Pd 


A9 


Cd 


In 


Sn 


Sb 


Te 


I, 


Xe 


Cs 


Ba 


La- 
Lu 


«f 


W 


Re 


Os 


Pt 


Au 


H9 


Tt 


Pb 


Bi 


Po 


At, 


Rn 


Fr 


Ra 


Ac- 

Lw 


La. 

Ce 

Pr 

Nd 

Pm 

Sm 

Eu 

Od 

Tb 

Dy 

Ho 

Er- 

Tm 

Yb 

Lu 

Ac 

Th 

Pa 

U 

Np 

Pu 

Ant 

Ctn 

Bk 

Cf 

Es 

Fm 

Md 

L*v 

304 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS    I    CHAP.    17 


metallic  behavior  that  suggests  there  is  a  special 
electron  configuration. 

Metals  also  possess  unusually  high  thermal 
conductivity,  as  anyone  who  has  drunk  hot 
coffee  from  a  tin  cup  can  testify.  It  is  noteworthy 
that  among  metals  the  best  electrical  conductors 
are  also  the  best  thermal  conductors.  This  is  a 
clue  that  these  two  properties  are  somehow 
related  and,  again,  the  electron  configuration 
proves  to  be  responsible. 

Though  the  mechanical  properties  of  the  vari- 
ous metals  differ,  all  metals  can  be  drawn  into 
wires  and  hammered  into  sheets  without  shatter- 
ing. Here  we  find  a  fourth  characteristic  property 
of  metals:  they  are  malleable  or  workable. 

LOCATION    OF    METALS    IN    THE 
PERIODIC    TABLE 

The  location  of  the  metals  in  the  periodic  table 
is  shown  in  Figure  17-4.  We  see  that  the  metals 
are  located  on  the  left  side  of  the  table,  while  the 
nonmetals  are  exclusively  in  the  upper  right 
corner.  Furthermore,  the  elements  on  the  left 
side  of  the  table  have  relatively  low  ionization 
energies.  We  shall  see  that  the  low  ionization 
energies  of  the  metallic  elements  aid  in  explain- 
ing many  of  the  features  of  metallic  behavior. 

ELECTRON    BEHAVIOR    IN    METALS 

What  is  the  nature  of  the  metallic  bond?  This 
bond,  like  all  others,  forms  because  the  electrons 
can  move  in  such  a  way  that  they  are  simultane- 
ously near  two  or  more  positive  nuclei.  Our 
problem  is  to  obtain  some  insight  into  the  special 
way  in  which  electrons  in  metals  do  this. 

Consider  a  crystal  of  metallic  lithium.  In  its 
crystal  lattice,  each  lithium  atom  finds  around 
itself  eight  nearest  neighbors.  Yet  this  atom  has 
only  one  valence  electron,  so  it  isn't  possible  for 
it  to  form  ordinary  electron  pair  bonds  to  all  of 
these  nearby  atoms.  However,  it  does  have  four 
valence  orbitals  available  so  its  electron  and  the 
valence  electrons  of  its  neighbors  can  approach 
quite  close  to  its  nucleus.  Thus  each  lithium 
atom  has  an  abundance  of  valence  orbitals  but 
a  shortage  of  bonding  electrons. 

Consider  the  dilemma  of  the  valence  electron 
of  a  particular  lithium  atom.  It  finds  eight  neigh- 


bor nuclei  nearby  and  complete  freedom  of 
movement  in  the  empty  valence  orbitals  around 
its  parent  nucleus.  Everywhere  the  electron 
moves  it  finds  itself  between  two  positive  nuclei. 
All  of  the  space  around  a  central  atom  is  a 
region  of  almost  uniformly  low  potential  energy. 
Under  these  circumstances,  it  is  not  surprising 
that  an  electron  can  move  easily  from  place  to 
place.  Each  valence  electron  is  virtually  free  to 
make  its  way  throughout  the  crystal. 

This  type  of  argument  leads  us  to  picture  a 
metal  as  an  array  of  positive  ions  located  at  the 
crystal  lattice  sites,  immersed  in  a  "sea"  of 
mobile  electrons.  The  idea  of  a  more  or  less 
uniform  electron  "sea"  emphasizes  an  important 
difference  between  metallic  bonding  and  ordi- 
nary covalent  bonding.  In  molecular  covalent 
bonds  the  electrons  are  localized  in  a  way  that 
fixes  the  positions  of  the  atoms  quite  rigidly.  We 
say  that  the  bonds  have  directional  character — 
the  electrons  tend  to  remain  concentrated  in 
certain  regions  of  space.  In  contrast,  the  valence 
electrons  in  a  metal  are  spread  almost  uniformly 
throughout  the  crystal,  so  the  metallic  bond  does 
not  exert  the  directional  influence  of  the  ordinary 
covalent  bond. 

We  can  obtain  some  idea  of  the  effectiveness 
of  this  electron  "sea"  in  binding  the  atoms  to- 
gether if  we  compare  the  energy  necessary  to 
vaporize  one  mole  of  a  metal  to  the  free  atoms 
with  the  energy  required  to  break  one  mole  of 
ordinary  covalent  bonds.  We  find  that  the  energy 
necessary  to  vaporize  a  mole  of  one  of  the  alkali 
metals  is  only  one-fourth  to  one-third  of  the 
energy  needed  to  break  a  mole  of  ordinary  co- 
valent bonds.  This  is  not  too  surprising.  The 
ionization  energy  of  a  free  alkali  metal  atom  is 
small ;  this  means  that  the  valence  electron  in  the 
free  atom  does  not  experience  a  strong  attraction 
to  the  nucleus.  Since  the  electron  is  not  strongly 
attracted  by  one  alkali  metal  atom,  it  is  not 
strongly  attracted  by  two  or  three  such  atoms 
in  the  metallic  crystal.  Thus,  the  binding  energy 
between  electrons  and  nuclei  in  the  alkali  metal 
crystals  is  rather  small,  and  the  resulting  metallic 
bonds  are  rather  weak.  We  might  expect,  how- 
ever, that  the  metallic  bond  would  become 
stronger  in  those  elements  which  have  a  greater 


SEC.     17-1    I    THE    ELEMENTS 


305 


number  of  valence  electrons  and  a  greater  nu- 
clear charge.  In  these  cases  there  are  more  elec- 
trons in  the  "sea,"  and  each  electron  is  more 
strongly  bound,  owing  to  the  increased  nuclear 
charge.  This  argument  is  in  accord  with  the  ex- 
perimental heats  of  vaporization  shown  in  Table 
17-1. 


Table  17-1 

HEATS    OF    VAPORIZATION    OF    METALS 
(kcal/mole) 


Second  Row 


Third  row 


Fourth  Row 


Fifth  Row 


Sixth  Row 


Li 

32.2 

Be 
53.5 

B 

129 

Na 
23.1 

Mg 
31.5 

Al 
67.9 

K 
18.9 

Ca 

36.6 

Sc 

73 

Rb 
18.1 

Sr 
33.6 

Y 

94 

Cs 

16.3 

Ba 

35.7 

La 

96 

To  pick  a  specific  case,  let  us  compare  the 
heats  of  vaporization  of  magnesium  and  alumi- 
num. The  higher  value  for  aluminum  shows  that 
the  metallic  bond  is  indeed  stronger  when  the 
number  of  valence  electrons  and  the  charge  on 
the  nucleus  increase.  Thus  the  strength  of  the 
metallic  bond  tends  to  increase  as  we  go  from 
left  to  right  along  a  row  in  the  periodic  table. 
The  transition  metal  elements  are  harder  and 
melt  and  boil  at  higher  temperatures  than  the 
alkali  or  alkaline  earth  metals. 

EXPLANATION    OF    THE    PROPERTIES 
OF    METALS 

The  nonlocalized  or  mobile  electrons  account  for 
the  many  unique  features  of  metals.  Since  metal- 
lic bonds  do  not  have  strong  directional  charac- 
ter, it  is  not  surprising  that  many  metals  can  be 
easily  deformed  without  shattering  their  crystal 
structure.  Under  the  influence  of  a  stress,  one 


plane  of  atoms  may  slip  by  another,  but  as  they 
do  so,  the  electrons  are  able  to  maintain  some 
degree  of  bonding  between  the  two  planes. 
Metals  can  be  hardened  by  alloying  them  with 
elements  which  do  have  the  property  of  forming 
directed  covalent  bonds.  Often  just  a  trace  of 
carbon,  phosphorus,  or  sulfur  will  turn  a  rela- 
tively soft  and  workable  metal  into  a  very  brittle 
solid. 


*^W£% 


Fig.  17-5.  Slippage  of  planes  of  metal  atoms. 


Metals  conduct  electricity  because  some  va- 
lence electrons  are  free  to  move  throughout  the 
solid.  At  the  same  time,  these  mobile  electrons 
are  effective  in  holding  the  crystal  together  be- 
cause wherever  they  move,  they  are  simultane- 
ously close  to  two  or  more  nuclei.  In  covalently 
bonded  solids  the  electrons  are  strongly  localized 
in  the  space  between  a  particular  pair  of  atoms. 
In  order  for  these  substances  to  conduct  elec- 
tricity, a  great  deal  of  energy  must  be  supplied 
to  remove  the  electrons  from  this  region  between 
the  atoms.  This  energy  is  not  available  in  normal 
electric  fields,  so  covalent  substances  do  not 
normally  conduct  electricity. 

The  excellent  heat  conductivity  of  metals  is 
also  due  to  the  mobile  electrons.  Electrons  which 


306 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS   |    CHAP.    17 


are  in  regions  of  high  temperature  can  acquire 
large  amounts  of  kinetic  energy.  These  electrons 
move  through  the  metal  very  rapidly  and  give 
up  their  kinetic  energy  to  heat  the  crystal  lattice 
in  the  cooler  regions.  In  substances  where  the 
electrons  are  highly  localized,  heat  is  conducted 
as  small  amounts  of  energy  are  transferred  from 
one  atom  to  its  immediate  neighbor;  this  is  a 
slower  process  than  electron  energy  conduction. 

To  complete  our  discussion  of  metallic  bond- 
ing we  must  explain  why  metallic  properties 
eventually  disappear  as  we  proceed  from  left  to 
right  along  a  row  in  the  periodic  table. 

We  have  seen  that  the  reasons  for  the  mobility 
of  electrons  in  metals  are  that  they  are  readily 
removed  from  the  atom  (the  ionization  energy  is 


low)  and  that  they  can  be  close  to  two  or  more 
positive  nuclei  just  about  anywhere  in  the  crystal 
(there  are  numerous  vacant  valence  orbitals).  As 
the  nuclear  charge  on  atoms  increases  and  the 
vacant  orbitals  become  filled,  the  regions  imme- 
diately between  two  nuclei  become  relatively 
more  attractive  to  the  electron,  compared  with 
all  other  regions.  Electrons  tend  to  be  more  and 
more  localized  in  these  regions,  and  normal 
covalent  bonds  with  their  directional  character 
appear. 

In  summary  we  can  say  that  the  metallic  bond 
is  a  sort  of  nondirectional  covalent  bond.  It 
occurs  when  atoms  have  few  valence  electrons 
compared  with  vacant  valence  orbitals  and  when 
these  valence  electrons  are  not  held  strongly. 


17-2    COMPOUNDS 


We  have  seen  that  the  pure  elements  may  solidify 
in  the  form  of  molecular  solids,  network  solids, 
or  metals.  Compounds  also  may  condense  to 
molecular  solids,  network  solids,  or  metallic 
solids.  In  addition,  there  is  a  new  effect  that  does 
not  occur  with  the  pure  elements.  In  a  pure  ele- 
ment the  ionization  energies  of  all  atoms  are 
identical  and  electrons  are  shared  equally.  In 
compounds,  where  the  most  stable  electron  dis- 
tribution need  not  involve  equal  sharing,  electric 
dipoles  may  result.  Since  two  bonded  atoms  may 
have  different  ionization  energies,  the  electrons 
may  spend  more  time  near  one  of  the  positive 
nuclei  than  near  the  other.  This  charge  separa- 
tion may  give  rise  to  strong  intermolecular  forces 
of  a  type  not  found  in  the  pure  elements. 

17-2.1    van  der  Waals  Forces  and 
Molecular  Substances 

Though  charge  separations  are  possible  in  com- 
pounds, there  are  many  molecules  that  do  not 
have  appreciable  electric  dipoles.  On  cooling, 
these  molecules  behave  much  like  the  molecules 
of  pure  elements.  If  the  bonding  capacity  of 
each  atom  is  completely  satisfied,  then  only  the 
weak  van  der  Waals  forces  remain  between  mole- 


cules. These  weak  interactions  give  low  melting 
solids  and  low  boiling  liquids  that  retain  many 
of  the  properties  of  the  gaseous  molecules. 

There  are  three  factors  that  seem  to  be  par- 
ticularly important  in  determining  the  magni- 
tudes of  van  der  Waals  forces:  the  number  of 
electrons,  the  molecular  size,  and  the  molecular 
shape.  These  factors  are  effective  both  for  ele- 
ments and  compounds,  though  greater  variety  is 
found  for  compounds. 

VAN    DER    WAALS    FORCES    AND 
NUMBER    OF    ELECTRONS 

We  have  already  observed  in  Chapter  6  that  the 
melting  and  boiling  points  of  the  inert  gases  in- 
crease as  the  number  of  electrons  increases  (see 
Figure  6-3).  Elements  and  compounds  with  co- 
valent bonding  behave  in  the  same  way.  Figure 
17-6  shows  this  in  a  graphical  presentation. 
Figure  17-6 A  shows  the  melting  and  boiling 
point  trends  among  the  inert  gases  and  among 
the  halogens.  The  horizontal  axis  shows  the  row 
number,  which  furnishes  an  index  of  the  total 
number  of  electrons  of  the  respective  elements. 
Figure  17-6B  refers  to  compounds  with  formulas 
CX4.  Again  the  horizontal  axis  shows  the  row 
number  but  now  of  the  outermost  atoms  in  the 


SEC.     17-2    I    COMPOUNDS 


307 


-400 


300 


T,  °K 


200 


100 


400 


300 


200 


100 


*   Sotting  point 
•  Melting  point 


•    Boiling  point 

CH+      CFj         •-  —  — •  Melting  point 


12  3  4  5  6 

Row    number  of  JC  in.    CX 

B 


Fig.  17-6.  The  melting  and  boiling  points  of  some 
molecular  compounds  and  the  halogens. 
A.  The  inert  gases  and  the  halogens.  B. 
The  carbon  compounds  of  formula  CXt. 


molecule  since  these  are  the  atoms  which  "rub 
shoulders"  with  neighboring  molecules.  As  far 
as  van  der  Waals  forces  are  concerned,  it  is  quite 
important  that  CBr4  has  atoms  from  the  fourth 
row  of  the  periodic  table  on  the  "surface"  of  the 
molecule  and  somewhat  less  important  that  the 
central  atom,  carbon,  is  from  the  second  row. 
The  outermost  atoms  are  most  influential  in  fix- 
ing intermolecular  forces. 

VAN    DER    WAALS    FORCES    AND 
MOLECULAR    SIZE 

If  comparisons  are  made  among  similar  mole- 
cules, then  the  larger  the  molecule,  the  higher 
is  its  melting  point.  For  example,  if  we  compare 
methane,  CH4,  and  ethane,  C2H6,  the  exterior 
atoms  are  the  same — hydrogen  atoms.  Still,  the 
boiling  point  of  ethane,  185°K,  is  higher  than 
that  of  methane,  112°K.  This  difference  is  at- 


tributed to  the  fact  that  there  must  be  greater 
contact  surface  between  two  ethane  molecules 
than  between  two  methane  molecules.  The  same 
effect  is  found  for  C2F6  (boiling  point,  195°K) 
and  CF4  (boiling  point,  145°K);  for  C2Br6  (this 
substance  decomposes  at  483°K  before  it  reaches 
its  boiling  point)  and  CBr4  (boiling  point, 
463°K). 

Notice  that  the  two  factors  just  mentioned,  number  of 
electrons  and  molecular  size,  might  lead  to  another  gen- 
eralization— that  the  boiling  point  goes  up  in  proportion 
to  molecular  weight.  The  molecular  weight,  the  molecular 
size,  and  the  number  of  electrons  all  tend  to  increase 
together.  This  molecular  weight-boiling  point  correlation 
has  some  usefulness  among  molecules  of  similar  com- 
position and  general  shape  but  chemists  do  not  feel  that 
there  is  a  direct  causative  relation  between  molecular 
weight  and  boiling  point. 

VAN   DER    WAALS   FORCES   AND 
MOLECULAR    SHAPE 

Substances  whose  structures  have  a  high  degree 
of  symmetry  generally  have  higher  melting  points 
than  closely  related  compounds  that  lack  this 
symmetry.  There  are  striking  examples  of  this 


308 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS   |    CHAP.     17 


Tabic   17-11.       THE    EFFECT    OF    MOLECULAR    SHAPE    ON     MELTING    POINT:    cis- 
A  N  D    trans-   ISOMERS 


m.p.  cis- 


m.p.  trans- 


1,2-dichloroethylene  CICH=CHCI  -80°C 

butenoic  acid  CH3CH=CHCOOH  15° 

fumaric,  maleic  acids  HOOCCH=CHCOOH  130° 


-50°C  (nans-  30°  higher  m.p.) 

72°  (trans-  57°  higher  m.p.) 

290°  (trans-  160°  higher  m.p.) 


among  the  double-bonded  compounds.  The  cis- 
and  trans-  isomers  of  many  such  compounds 
have  melting  point  and  boiling  point  differences 
that  can  be  traced  to  the  differences  in  molecular 
shape.  For  example,  the  higher  melting  point  of 
trans-\,2  dichloroethylene  than  that  of  the  iso- 
meric cis-\,2  dichloroethylene  may  be  partly  ex- 
plained by  arguing  that  the  long  and  symmetrical 
trans-  form  can  pack  into  an  orderly  crystal 
lattice  in  a  neater  and  more  compact  fashion 
than  the  "one-sided"  cis-  form  molecules.  This 
and  two  other  examples  of  this  are  shown  in 
Table  17-11.  (The  dichloroethylene  structures  are 
shown  in  Figure  16-18.) 

Another  example  of  the  influence  of  molecular 
symmetry  on  physical  properties  is  found  in  two 
structural  isomers  of  the  formula  C6Hi2.  These 


are  called  normal  pentane  and  neopentane  and 
their  molecular  shapes  differ  drastically  as 
shown  in  Figure  17-7. 

The  extended  molecule,  H-pentane,  has  a  zig- 
zag shape.  We  see  that  van  der  Waals  forces  act 
between  the  external  envelope  of  hydrogen  atoms 
of  one  molecule  and  those  of  adjacent  molecules. 
This  large  surface  contact  gives  a  relatively  high 
boiling  point.  On  the  other  hand,  this  flexible, 
snake-like  molecule  does  not  pack  readily  in  a 
regular  lattice,  so  its  crystal  has  a  low  melting 
point. 

Contrast  the  highly  compact,  symmetrical  neo- 


Fig.  17-7.  Molecular  shape,  a  factor  that  influences 
melting  and  boiling  points. 


CH* 


CH 


H,C 


Normal  pen-tane 
b.p.  36°C 

TTL.p. 


-130°C 


Neopenrtane 

b.p.  9°C 

m.p.     -200C 


SEC.     17-2    I    COMPOUNDS 


309 


pentane.  This  ball-like  molecule  readily  packs 
in  an  orderly  crystal  lattice  which,  because  of  its 
stability,  has  a  rather  high  melting  point.  Once 
melted,  however,  neopentane  forms  a  liquid  that 
boils  at  a  temperature  below  the  boiling  point  of 
«-pentane.  Neopentane  has  less  surface  contact 
with  its  neighbors  and  hence  is  more  volatile. 
It  is  well  to  add  that  most  of  the  compounds 
of  carbon  condense  to  molecular  liquids  and 
solids.  Their  melting  points  are  generally  low 
(below  about  300°C)  and  many  carbon  com- 
pounds boil  below  100°C.  The  similar  chemistry 
of  the  liquid  and  solid  phases  shows  the  reten- 
tion of  the  molecular  identities. 


17-2.2    Covalent  Bonds  and  Network 
Solid  Compounds 

Compounds  can  form  network  solids  and,  since 
two  or  more  different  atoms  are  involved,  there 
is  much  greater  variety  among  the  network  solid 
compounds  than  among  the  network  solid  ele- 
ments. Silica,  with  empirical  formula  Si02,  is  a 
network  solid.  Silica  and  other  silicon-oxygen 
compounds  make  up  about  87  %  of  the  earth's 
crust.  Almost  all  common  minerals  contain  sub- 
stantial amounts  of  silicates,  the  general  term  for 
silicon-oxygen  solids.  These  are  network  solids 
but  with  interesting  and  important  variations. 
Figures  17-8,  17-9,  and  17-10  show  in  a  schematic 
way  the  three  types  of  network  solids  formed  by 
silicon.  The  silicon  is  always  tetravalent  but  in 
some  of  its  compounds  it  forms  infinite  silicon- 
oxygen-silicon  chains;  in  some  it  forms  infinite 
interlinked  sheets;  and,  in  some,  it  forms  an 
infinite  three-dimensional  network  solid. 

Many  properties  of  silicates  can  be  understood 
in  terms  of  the  type  of  network  lattice  formed. 
In  the  "one-dimensional"  networks,  shown  in 
Figure  17-8,  the  atoms  within  a  given  chain  are 
strongly  linked  by  covalent  bonds  but  the  chains 
interact  with  each  other  through  much  weaker 
forces.  This  is  consistent  with  the  thread-like 
properties  of  many  of  these  silicates.  The  asbes- 
tos minerals  are  of  this  type. 

In  a  similar  way,  the  sheets  of  the  "two- 
dimensional"  network  silicates,  shown  in  Figure 


17-9,  are  held  together  weakly.  Hence  these 
minerals  cleave  readily  into  thin  but  strong 
sheets.  The  micas  have  this  type  of  structure. 
Clays  also  have  this  structure,  and  their  slippery 
"feel"  when  wet  can  be  explained  in  terms  of  the 
hydration  of  the  planes  on  the  outside  of  the 
crystals.  The  three-dimensional  network  shown 
in  Figure  17-10  is  silica  (quartz).  Like  diamond, 
it  is  hard  and  it  has  a  high  melting  point. 
The  various  minerals  that  make  up  granite  are 
of  this  type. 


17-2.3    Metallic  Alloys 

We  have  already  learned  that  metals  may  be  deformed 
easily  and  we  have  explained  this  in  terms  of  the  absence 
of  directional  character  in  metallic  bonding.  In  view  of 
this  principle,  it  is  not  surprising  that  two-element  or 
three-element  metallic  crystals  exist.  In  some  of  these, 
regular  arrangements  of  two  or  more  types  of  atoms  are 
found.  The  composition  then  is  expressed  in  simple  in- 
teger ratios,  so  these  are  called  metallic  compounds.  In 
other  cases,  a  fraction  of  the  atoms  of  the  major  con- 
stituent have  been  replaced  by  atoms  of  one  or  more 
other  elements.  Such  a  substance  is  called  a  solid  solution. 
These  metals  containing  two  or  more  types  of  atoms  are 
called  alloys. 

ELECTRICAL    CONDUCTIVITY 

Electrical  conductivity  in  metals  apparently  depends  upon 
the  smooth  and  uninterrupted  movement  of  electrons 
through  the  lattice.  This  is  suggested  by  the  fact  that  small 
amounts  of  impurities  reduce  the  conductivity  very  much. 
We  shall  see,  in  Chapter  22,  that  copper  is  purified  com- 
mercially to  99.999%  and  the  reason  is  directly  connected 
to  the  consequent  gain  in  electrical  conductivity. 
Table  17-111  shows  some  conductivities  of  copper  with 


Table  17-I/I 

CONDUCTIVITY    OF    COPPER    ALLOYS 
(ALL    AT    20  C    UNLESS    NOTED) 

PERCENT 
PERCENT  ALLOYING  ALLOYING         CONDUCTIVITY 

COPPER  ELEMENT  ELEMENT  (ohm-CmV"1 


100.00 

— 

— 

5.9    X  10s 

99 

Mn  (0°C) 

0.98 

2.1     X  10s 

95.8 

Mn 

4.2 

0.56  X  10s 

97 

Al    (0°C) 

3 

1.2    X  10s 

90 

Al    (0°C) 

10 

0.79  X  10s 

88 

Sn 

12 

0.56  X  10s 

Fig.  17-8.  One-dimensional  network  silicates: 

the  asbestos  minerals. 


Fig.  17-9.  Two-dimensional  network  silicates. 

the  mica  and  clay  minerals. 


**Qk, 


•  Silicon 
O  Oxygen 
X     Other  atoms 


Fig.  17-10.  Three-dimensional  network  silicates: 

granitic  minerals. 


SEC.     17-2    |    COMPOUNDS 


311 


5 

I 

i 

.Vj 


15 

10 

5 

- 

12  3  4 

Percervfr  Jrfti,   by  weight 

Fig.  17-11.  Electrical  resistivity  of  copper  containing 
manganese. 


various  impurities.*  Figure  17-11  shows  the  data  for 
copper-manganese  alloys  graphically.  The  figure  shows 
resistivity,  the  reciprocal  of  conductivity,  plotted  against 
percent  manganese  (by  weight).  The  importance  of  puri- 
fication of  copper  for  electrical  wire  is  evident  in  this 
figure  if  we  remember  that  the  power  lost  in  a  conductor 
is  proportional  to  resistance  (for  a  given  current).  In  a 
conductor  hundreds  of  miles  long,  a  factor  of  two  reduc- 
tion in  resistivity  is  a  lucrative  gain  to  a  company  selling 
electrical  power. 


EXERCISE  17-2 

Use  Figure  17-11  to  estimate  the  resistivities  of  two  metal 
samples,  one  made  of  pure  copper  and  the  other  of  a 
copper-manganese  alloy  containing  one  atom  of  manga- 
nese for  every  one  hundred  copper  atoms.  Calculate  the 
ratio  of  the  cost  due  to  power  loss  from  wire  of  the  impure 
material  to  the  cost  due  to  the  power  loss  from  wire  of 
the  pure  material. 


HARDNESS    AND    STRENGTH 

Alloys  are  harder  and  stronger  than  pure  metals  as  usu- 
ally prepared.  The  most  familiar  example  is  steel  and  pure 
iron.  The  tensile  strength  of  pure  iron  can  be  increased 


*  The  conductivities  are  given  in  units  (ohm-cm)-1.  The 
reciprocal  of  this  number  is  the  resistance  one  would  find 
(in  ohms)  for  a  wire  1  cm  long  and  with  a  cross-section 
of  1  cm*. 


ten-fold  by  the  addition  of  only  a  percent  of  carbon  and 
smaller  amounts  of  nickel  or  manganese.  The  tensile 
strength  of  brass  (65-70%  Cu,  35-30%  Zn)  is  more  than 
twice  that  of  copper  and  four  times  that  of  zinc. 

The  hardness  and  strength  of  alloys  can  be  explained 
in  terms  of  bonding.  The  impurity  atoms  added  may  form 
localized  and  rigid  bonds.  These  tend  to  prevent  the 
slippage  of  atoms  past  each  other,  which  results  in  a  loss 
of  malleability  and  an  increase  in  hardness. 


17-2.4    Ionic  Solids 

Thus  far  we  have  not  considered  the  effects  that 
arise  from  charge  separations.  The  most  extreme 
case  is  represented  by  the  formation  of  ionic 
solids.  Usually,  these  can  be  looked  on  as  arrays 
of  positive  and  negative  ions,  neatly  stacked  so 
that  each  positive  ion  has  only  negative  ion 
neighbors  and  each  negative  ion  has  only  posi- 
tive ion  neighbors.  Figure  5-10  (p.  81)  shows 
such  a  crystal  arrangement,  that  of  sodium 
chloride.  Why  does  such  a  solid  form  and  what 
are  its  properties?  These  are  the  questions  we 
shall  try  to  answer  here. 

THE   STABILITY    OF   IONIC    CRYSTALS 

In  discussing  the  bonding  in  the  gaseous  LiF 
molecule,  the  electric  dipole  of  the  molecule  is 
explained  in  terms  of  the  different  ionization 
energies  of  Li  and  F  atoms.  Though  the  molecule 
holds  together  because  the  bonding  electrons  are 
near  both  nuclei,  the  energy  favors  an  electron 
distribution  concentrated  toward  the  fluorine. 
A  stable  and  polar  molecule  is  formed.  Stable, 
perhaps,  but  in  the  gaseous  state,  reactive!  The 
valence  orbitals  of  the  lithium  atom  are  almost 
vacant.  According  to  our  experience  (for  exam- 
ple, with  CH2,  BH3,  carbon  atoms,  metal  atoms) 
the  presence  of  empty  valence  orbitals  implies 
that  additional  electron  sharing  can  occur.  Lith- 
ium fluoride  molecules  are,  then,  more  stable 
when  they  condense  so  as  to  place  each  lithium 
atom  simultaneously  near  several  fluorine  atoms. 
Just  as  in  metals,  an  atom  with  vacant  orbitals 
is  more  stable  with  several  neighbors.  Then  the 
electrons  held  by  the  neighbor  atoms  can  be  near 
two  or  more  nuclei  at  once.  There  is,  however, 
a  significant  difference  from  metals — in  solid 
lithium  fluoride,  half  of  the  atoms  have  high 


312 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS    I    CHAP.    17 


Fig.  17-12.  Sodium  chloride  crystals. 


ionization  energies.  Fluorine  atoms  hold  their 
electrons  tightly.  Therefore  the  characteristic 
electron  mobility  of  metals  is  not  present  in  the 
ionic  solids.  The  absence  of  mobile  electrons 
implies  that  none  of  the  metallic  properties  is 
expected.  Let  us  see  what  properties  such  a  solid 
does  have. 


PROPERTIES    OF    IONIC    CRYSTALS 

Ionic  solids,  such  as  lithium  fluoride  and  sodium 
chloride,  form  regularly  shaped  crystals  with  well 
defined  crystal  faces.  Pure  samples  of  these  solids 
are  usually  transparent  and  colorless  but  color 
may  be  caused  by  quite  small  impurity  contents 
or  crystal  defects.  Most  ionic  crystals  have  high 
melting  points. 

Molten  lithium  fluoride  and  sodium  chloride 
have  easily  measured  electrical  conductivities. 
Nevertheless,  these  conductivities  are  lower  than 
metallic  conductivities  by  several  factors  of  ten. 
Molten  sodium  chloride  at  750°C  has  a  conduc- 
tivity about  10-5  times  that  of  copper  metal  at 
room  temperature.  It  is  unlikely  that  the  electric 
charge  moves  by  the  same  mechanism  in  molten 
NaCl  as  in  metallic  copper.  Experiments  show 
that  the  charge  is  carried  in  molten  NaCl  by 
Na+  and  Cl~  ions.  This  electrical  conductivity 
of  the  liquid  is  one  of  the  most  characteristic 


properties  of  substances  with  ionic  bonds.  In 
contrast,  molecular  crystals  generally  melt  to 
form  molecular  liquids  that  do  not  conduct  elec- 
tricity appreciably. 


17-2.5    Effects  Due  to  Charge  Separation 

We  have  considered  the  weak  van  der  Waals 
forces  that  cause  the  condensation  of  covalent 
molecules.  The  formation  of  an  ionic  lattice 
results  from  the  stronger  interactions  among 
molecules  with  highly  ionic  bonds.  But  most 
molecules  fall  between  these  two  extremes. 
Most  molecules  are  held  together  by  bonds  that 
are  largely  covalent,  but  with  enough  charge 
separation  to  affect  the  properties  of  the  mole- 
cules. These  are  the  molecules  we  have  called 
polar  molecules. 

Chloroform,  CHC13,  is  an  example  of  a  polar 
molecule.  It  has  the  same  bond  angles  as  meth- 
ane, CH4,  and  carbon  tetrachloride,  CCL».  Car- 
bon, with  spz  bonding,  forms  four  tetrahedrally 
oriented  bonds  (as  in  Figure  16-11).  However, 
the  cancellation  of  the  electric  dipoles  of  the  four 
C — CI  bonds  in  CCI4  does  not  occur  when  one 
of  the  chlorine  atoms  is  replaced  by  a  hydrogen 
atom.  There  is,  then,  a  molecular  dipole  remain- 
ing. The  effects  of  such  electric  dipoles  are 
important  to  chemists  because  they  affect  chemi- 
cal properties.  We  shall  examine  one  of  these, 
solvent  action. 


SEC.    17-2   I    COMPOUNDS 


313 


SOLVENT    PROPERTIES    AND 
MOLECULAR    DIPOLES 

The  forces  between  molecules  are  strongly  af- 
fected by  the  presence  of  molecular  dipoles.  Two 
molecules  that  possess  molecular  dipoles  tend  to 
attract  each  other  more  strongly  than  do  mole- 
cules without  dipoles.  One  of  the  most  important 
results  of  this  is  found  in  solvent  properties. 
Table  17-IV  shows  some  solubility  data  of 

Table  17-IV 

SOLUBILITIES    IN    CARBON    TETRA- 
CHLORIDE,   CCI„    AND    IN    ACETONE, 

CHCOCH     (25°C,   moles   liter) 

SOLVENT 


SOLUTE 

CCh 

CH3COCH3 

SOLUTE 

POLARITY 

{nonpolar) 

(polar) 

CH«,  methane 

nonpolar 

0.029 

0.025 

CjH6,  ethane 

nonpolar 

0.22 

0.13 

CH3C1, 

chloromethane 

polar 

1.7 

2.8 

CH3OCH3, 

methyl  ether 

polar 

1.9 

2.2 

various  solutes  in  the  two  solvents,  carbon  tetra- 
chloride, CCLj  and  acetone,  CH3COCH3.  These 
two  solvents  differ  in  their  polar  properties.  In 
CCU  the  central  carbon  atom  is  surrounded  by 
four  bonds  that  form  a  regular  tetrahedron  like 
that  pictured  in  Figure  16-11.  With  this  molecu- 
lar shape,  CCLj  has  a  zero  molecular  dipole.  In 
contrast,  acetone  has  a  bent  structure  and  the 
oxygen  atom  gives  it  a  significant  electric  dipole. 

Contrast  the  solubilities  in  Table  17-IV.  The 
first  two  substances,  CH4  and  C2H6,  have  zero 
molecular  dipoles.  In  each  case,  the  solubility  in 
CCU  exceeds  the  solubility  in  CH3COCH3.  The 
next  two  substances,  CH3C1  and  CH3OCH3,  have 
nonzero  molecular  dipoles.  In  each  of  these 
cases,  the  solubility  in  acetone  is  the  larger. 

There  is  a  reasonable  explanation  of  the  data 
in  Table  17-IV.  When  a  solute  dissolves,  the 
solute  molecules  must  be  separated  from  each 
other  and  then  surrounded  by  solvent  molecules. 
Furthermore,  the  solvent  molecules  must  be 
pushed  apart  to  make  room  for  the  solute  mole- 


cules. Since  dipoles  interact  strongly  with  each 
other,  a  polar  mblecule  such  as  CH3CI  is  en- 
ergetically more  stable  when  surrounded  by 
solvent  molecules  that  are  also  polar.  Hence, 
CH3CI  has  the  higher  solubility  in  the  polar 
solvent  acetone  than  in  carbon  tetrachloride.  On 
the  other  hand,  a  nonpolar  molecule  such  as 
CH4  will  find  it  difficult  to  wedge  in  between  the 
strongly  interacting  molecules  of  a  polar  solvent 
— more  difficulty  than  it  will  encounter  in  dis- 
solving in  a  nonpolar  solvent.  Hence  CH4  has 
higher  solubility  in  the  nonpolar  solvent,  CCI4. 

SOLUBILITY    OF    ELECTROLYTES 
IN    WATER 

The  dissolving  of  electrolytes  in  water  is  one  of 
the  most  extreme  and  most  important  solvent 
effects  that  can  be  attributed  to  electric  dipoles. 
Crystalline  sodium  chloride  is  quite  stable,  as 
shown  by  its  high  melting  point,  yet  it  dissolves 
readily  in  water.  To  break  up  the  stable  crystal 
arrangement,  there  must  be  a  strong  interaction 
between  water  molecules  and  the  ions  that  are 
formed  in  the  solution.  This  interaction  can  be 
explained  in  terms  of  the  dipolar  properties  of 
water. 

When  an  electric  dipole  is  brought  near  an  ion, 
the  energy  is  lower  if  the  dipole  is  oriented  to 
place  unlike  charges  in  proximity.  Hence  water 
molecules  tend  to  orient  preferentially  around 
ions,  the  positive  end  of  the  water  dipole  pointing 
inward  if  the  ion  carries  negative  charge  and  the 
negative  end  pointing  inward  if  the  ion  carries 
positive  charge.  Figure  17-13  shows  this  process 
schematically:  it  is  called  hydration. 

There  are  two  effects  of  the  orientation  of 
water  dipoles  around  the  ions.  First,  the  energy 
is  lowered  because  the  orientation  serves  to  bring 
unlike  charges  near  each  other.  This  tends  to 
encourage  the  ions  to  leave  the  sodium  chloride 
crystal  and  enter  the  solution.  Also,  however, 
there  is  an  effect  on  randomness  whose  magni- 
tude is  difficult  to  predict.  The  orientation  of  the 
water  molecules  around  the  ion,  fixing  them  rela- 
tive to  the  ion,  constitutes  an  orderly  arrange- 
ment. Since  all  systems  tend  toward  maximum 
randomness,  the  orientation  effect  works  against 
molecules  leaving  the  crystal  to  enter  the  solu- 


314 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS    I    CHAP.     17 


Fig.  17-13.  Hydration  of  ions:  orientation  of  water 
dipoles  around  ions  in  aqueous  solutions. 


tion.  These  two  effects  of  ion-hydration,  lower- 
ing the  energy  of  an  electrolyte  solute  while 
decreasing  the  change  in  randomness  as  it  dis- 
solves, give  water  distinctive  properties  as  an 
electrolyte  solvent.  It  helps  explain,  for  example, 
why  some  salts  absorb  heat  as  they  dissolve  in 
water  (for  example,  NH4C1)  while  some  release 
heat  as  they  dissolve  (for  example,  NaOH).  For 
most  solvents  the  crystal  has  lower  energy  than 
the  solution,  and  heat  is  absorbed  as  solid  dis- 
solves. In  water,  however,  the  hydration  effects 
can  cause  the  solution  to  have  the  lower  energy, 
so  heat  can  be  evolved  during  the  dissolving 
process. 


17-2.6    Hydrogen  Bonds 

In  Figure  17-6 A  we  saw  that  the  boiling  points  of 
symmetrical  molecules  increase  regularly  as  we 
drop  down  in  the  periodic  table.  Figure  17-14 
shows  the  corresponding  plot  for  some  molecules 
possessing  electric  dipoles. 

Consider  first  the  boiling  points  of  HI,  HBr, 
HC1,  and  HF.  The  last,  hydrogen  fluoride,  is  far 
out  of  line,  boiling  at  19.9°C  instead  of  below 
—  95°C  as  would  be  predicted  by  extrapolation 
from  the  other  three.  There  is  an  even  larger 
discordancy  between  the  boiling  point  of  H20 
and  the  value  we  would  predict  from  the  trend 
suggested  by  H2Te,  H2Se,  and  H2S. 

Could  the  extremely  high  boiling  points  of  HF 
and  H20  be  due  to  the  fact  that  these  are  the 
smallest  molecules  of  their  respective  series?  No, 


SEC.     17-2    I    COMPOUNDS 


315 


this  does  not  appear  to  be  the  explanation,  for 
corresponding  discrepancies  are  not  present  in 
the  data  plotted  in  Figure  17-6A.  There  must  be 
some  other  explanation  for  these  exceptional 
boiling  points.  There  must  be  forces  of  some  new 
kind  between  the  molecules  of  H20  and  of  HF 
that  tend  to  keep  them  in  the  liquid  phase. 

These  same  forces  are  recognized  in  solid  com- 
pounds. The  most  familiar  example  is  solid  H20, 
or  ice.  Ice  has  a  crystal  structure  in  which  the 
oxygen  and  hydrogen  atoms  are  distributed  in  a 
regular  hexagonal  crystalline  lattice  that  some- 
what resembles  the  diamond  lattice  (see  Figure 
17-2).  Each  oxygen  atom  is  surrounded  by  four 
other  oxygen  atoms  in  a  tetrahedral  arrange- 
ment. The  hydrogen  atoms  are  found  on  the  lines 
extending  between  the  oxygen  atoms. 

/ 
O— H----0 

/  \ 

The  attractive  force  between  — OH  and  O  must 
be  the  bond  that  joins  the  water  molecules  to- 
gether into  the  crystal  lattice  of  ice.  This  bond 
is  a  hydrogen  bond. 

ENERGY    OF    HYDROGEN    BONDS 

The  hydrogen  bond  is  usually  represented  by 

O — H O  in  which  the  solid  line  represents 

the  original  O — H  bond  in  the  parent  compound 
(as  in  water,  HOH,  or  methyl  alcohol,  CH3OH). 


Fig.  17-14.  The  boiling  points  of  some  hydrides. 
ICO  "'° 


Bo i tiny  poin-t, 


SO  - 


-100 


The  dotted  line  shows  the  second  bond  formed 
by  hydrogen,  the  bond  called  the  hydrogen  bond. 
It  is  usually  dotted  to  indicate  that  it  is  much 
weaker  than  a  normal  covalent  bond.  Considera- 
tion of  the  boiling  points  in  Figure  17-14,  on  the 
other  hand,  shows  that  the  interaction  must  be 
much  stronger  than  van  der  Waals  forces.  Ex- 
periments show  that  most  hydrogen  bonds  re- 
lease between  3  kcal/mole  and  10  kcal/mole 
upon  formation: 


AH  =  -3  to  -7  kcal/mole 


(/) 


The  energy  of  this  bond  places  it  between  van 
der  Waals  and  covalent  bonds.  Roughly  speak- 
ing, the  energies  are  in  the  ratio 


van  der  Waals 

attractions 

1 


hydrogen  bonds  :  covalent  bonds 
10  :  100 


WHERE  HYDROGEN  BONDS  ARE  FOUND 

Hydrogen  bonds  are  found  between  only  a  few 
atoms  of  the  periodic  table.  The  commonest  are 
those  in  which  H  connects  two  atoms  from  the 
group  F,  O,  and  N,  and  less  commonly  CI. 

The  hydrogen  bond  to  fluorine  is  clearly  evi- 
dent in  most  of  the  properties  of  hydrogen  fluo- 
ride. The  high  boiling  point  of  HF,  compared 
with  those  of  the  other  hydrogen  halides,  is  one 
of  several  pieces  of  data  that  show  that  HF  does 
not  exist  in  the  liquid  compound  as  separate  HF 
molecules.  Instead  there  are  aggregates  of  mole- 
cules, which  we  describe  in  general  terms  as 
(HF)*.  Gaseous  hydrogen  fluoride  contains  the 
molecular  species  H2F2,  H3F3,  and  so  on  up  to 
H6F6  as  well  as  some  single  HF  molecules. 

These  species  can  be  represented  in  a  descrip- 
tive formula  such  as  the  following: 

H— F  •  •   H— F        H— F        H— F        (2) 

An  extreme  example  of  the  fluorine-hydrogen 
bond  is  found  in  the  hydrogen  difluoride  ion, 
HF2" .  This  ion  exists  in  acidic  solutions  of  fluo- 
rides, 


H+(aq)  +  ¥~(aq) 
HF(aq)  +  F-(aq) 


HF(aq) 
HF2"  (aq) 


(i) 
00 


Row  number  of  X 
in  HX  and  H2X 


and  in  the  ionic  crystal  lattice  of  salts  such  as 
KHF2.  The  HF2~  ion  may  be  regarded  as  con- 


316 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS   |    CHAP.     17 


sisting  of  two  negatively  charged  fluoride  ions 
held  together  by  a  proton: 


F      ©     F 


It  is  not  typical,  however,  for  few  hydrogen 
bonds  form  with  the  proton  equidistant  from 
the  two  atoms  to  which  it  bonds. 


this  difference  is  that  intramolecular  hydrogen  bonds  can 
exist  between  the  two  — COOH  groups  for  maleic  acid 
but  not  for  fumaric  acid: 


H 


\ 


H 


H 


(5)      maleic  acid      0=C 


OH     -O 


S 


C— O 


(8) 


O 

H  C  H 

\  /   \   / 

fumaric  acid  C  C  O 

/  \ 

0=C  H 


(9) 


INTER-    AND    INTRAMOLECULAR 
HYDROGEN    BONDS 

One  of  the  factors  connected  with  the  formation  of  strong 
hydrogen  bonds  is  the  acidic  character  of  the  hydrogen 
atom  involved.  Thus  the  hydrogen  bond  formed  by  hy- 
drogen fluoride  is  one  of  the  strongest  known.  Acetic  acid, 
CH3COOH,  is  a  representative  of  an  important  class  of 
acidic  hydrogen  bonding  compounds.  All  of  the  members 
of  this  class  possess  the  structural  unit  called  the  car- 
boxylic  acid  group: 


(6) 


O— H 


For  this  type  of  compound,  the  formation  of  hydrogen 
bonds  can  lead  to  the  coupling  of  the  molecules  in  pairs, 
to  form  a  cyclic  structure: 

O...H— O 

2CH3COOH^rCH3— C  C— CH3 

\  S 

O— H        O 

AH  =  -14kcal    (7) 

Here  the  favorable  geometrical  arrangement  with  two 
hydrogen  bonds  contributes  14  kcal  to  the  stability  of  the 
hydrogen  bonded  product,  (7).  These  are  called  inter- 
molecular  hydrogen  bonds  (inter  means  between). 

Hydrogen  bonds  can  also  exist  when  the  O — H  group 
and  the  other  bonding  atom  are  close  together  in  the 
same  molecule  in  such  positions  that  a  ring  can  be  formed 
without  disturbing  the  normal  bond  angles.  These  are 
called  w/ramolecular  hydrogen  bonds  (intra  means 
within). 

An  example  of  intramolecular  hydrogen  bonding 
is  provided  by  the  cis-  and  trans-  forms  of  the  acid 
HOOC— CH=CH— COOH.  The  trans-  form,  fumaric 
acid,  has  a  higher  melting  point  than  the  cis-  form,  maleic 
acid.  In  addition  to  the  general  effect  of  molecular  shape 
(mentioned  earlier  in  this  chapter),  another  reason  for 


This  intramolecular  bonding  in  maleic  acid,  (8),  halves 
its  ability  to  form  intermodular  bonds.  In  fumaric  acid, 
on  the  other  hand,  all  of  the  hydrogen  bonds  form  between 
molecules  (intermolecular  bonds)  to  give  a  stronger,  in- 
terlinked crystal  structure. 

THE  NATURE  OF  THE  HYDROGEN  BOND 

In  the  hydrogen  bond  we  find  the  hydrogen  atom 
attached  to  two  other  atoms.  Yet  our  bonding 
rules  tell  us  that  the  hydrogen  atom,  with  only 
the  \s  orbital  for  bond  formation,  cannot  form 
two  covalent  bonds.  We  must  seek  an  explana- 
tion of  this  second  bond. 

The  simplest  explanation  for  the  hydrogen 
bond  is  based  upon  the  polar  nature  of  F — H, 
O — H,  and  N — H  bonds.  In  a  molecule  such  as 
H20,  the  electron  pair  in  the  O — H  bond  is  dis- 
placed toward  the  oxygen  nucleus  and  away  from 
the  hydrogen  nucleus.  This  partial  ionic  charac- 
ter of  the  O — H  bond  lends  to  the  hydrogen 
atom  some  positive  character,  permitting  elec- 
trons from  another  atom  to  approach  closely  to 
the  proton  even  though  the  proton  is  already 
bonded.  A  second,  weaker  link  is  formed. 

THE    SIGNIFICANCE   OF   THE 
HYDROGEN    BOND 

Hydrogen  bonds  play  an  important  part  in  de- 
termining such  properties  as  solubility,  melting 
points,  and  boiling  points,  and  in  affecting  the 
form  and  stability  of  crystal  structures.  They 
play  a  crucial  role  in  biological  systems.  For  ex- 


QUESTIONS    AND    PROBLEMS 


317 


ample,  water  is  so  common  in  living  matter  that 
it  must  influence  the  chemical  behavior  of  many 
biological  molecules,  most  of  which  can  also 
form  hydrogen  bonds.  Water  can  attach  itself  by 
hydrogen  bonding,  either  by  providing  the  pro- 
ton, as  in 


H 


/ 


O— H 


o=c 


\ 


(10) 


or  by  accepting  the  proton,  as  in 

H 
\  / 


/ 


N— H 


O 


<") 


H 


Furthermore,  intramolecular  hydrogen  bonding 
is  one  of  the  chief  factors  in  determining  the 
structure  of  such  important  biological  sub- 
stances as  proteins,  as  discussed  in  Chapter  24. 


QUESTIONS  AND  PROBLEMS 


1.  Make  a  table  that  contrasts  the  melting  points 
and  boiling  points  of  LiF,  Li,  and  F2,  expressing 
the  temperatures  on  the  Centigrade  scale. 

2.  Without  looking  in  your  textbook,  do  the  fol- 
lowing. 

(a)  Draw  an  outline  of  the  periodic  table,  indi- 
cating the  rows  but  not  the  individual  ele- 
ments. 


(b)  Place  a  number  at  the  left  of  each  row  in- 
dicating the  number  of  elements  in  that  row. 

(c)  Fill  in  the  symbols  for  as  many  of  the  first 
1 8  elements  as  you  can  (leave  blank  any  that 
you  forget). 

(d)  Draw  two  diagonal  lines  across  the  table  to 
separate  it  into  three  regions.  Write  in  each 
region  one  of  the  words  "metals,"  "non- 
metals,"  "covalent  solids." 

(e)  Now  compare  your  diagram  to  Figure  17-4. 


Sulfur  exists  in  a  number  of  forms,  depending  upon  the  temperature  and,  sometimes,  upon  the  past 
history  of  the  sample.  Three  of  the  forms  are  described  below.  A  is  the  room  temperature  form  and 
it  changes  to  B  above  the  melting  point  of  A,  113°C.  B  changes  to  C  on  heating  above  160°C. 


Crystalline  solid 

Yellow  color,  no  metallic 

luster 
m.p.  =  113°C 
Dissolves  in  CSj,  not  in 

water 
Electrical  insulator 


113°C 


B 

Liquid 

Clear,  straw  color 
Viscosity  (fluidity)  about 
the  same  as  water 


Electrical  insulator 


^200°C 


C 

Liquid 

Dark  color 

Very  viscous  (syrupy) 


Electrical  insulator 


Which  of  the  following  structures  would  be  most  likely  to  account  for  the  observed  properties  of  each 
of  the  three  forms  described  above? 


(a)  a  metallic  crystal  of  sulfur  atoms; 

(b)  a  network  solid  of  sulfur  atoms; 

(c)  an  ionic  solid  of  S+  and  S~  ions; 

(d)  a  molecular  crystal  of  Ss  molecules; 

(e)  a  metallic  liquid  like  mercury; 


(0  a  molecular  liquid  of  Ss  molecules; 

(g)  a  molecular  liquid  of  Sn  chains,  with  n  =  a 

very  large  number; 
(h)  an  ionic  liquid  of  S+  and  S"  ions. 


318 


THE    BONDING    IN    SOLIDS    AND    LIQUIDS    I    CHAP.    17 


4.  Contrast  the  bonds  between  atoms  in  metals,  in 
van  der  Waals  solids,  and  in  network  solids  in 
regard  to: 

(a)  bond  strength; 

(b)  orientation  in  space; 

(c)  number  of  orbitals  available  for  bonding. 

5.  Aluminum,  silicon,  and  sulfur  are  close  together 
in  the  same  row  of  the  periodic  table,  yet  their 
electrical  conductivities  are  widely  different. 
Aluminum  is  a  metal;  silicon  has  much  lower 
conductivity  and  is  called  a  semiconductor;  sul- 
fur has  such  low  conductivity  it  is  called  an 
insulator.  Explain  these  differences  in  terms  of 
valence  orbital  occupancy. 

6.  Sulfur  is  made  up  of  Sg  molecules;  each  molecule 
has  a  cyclic  (crown)  structure.  Phosphorus  con- 
tains P4  molecules;  each  molecule  has  a  tetra- 
hedral  structure.  On  the  basis  of  molecular  size 
and  shape,  which  would  you  expect  to  have  the 
higher  melting  point? 

7.  Discuss  the  conduction  of  heat  by  copper  (a 
metal)  and  by  glass  (a  network  solid)  in  terms 
of  the  valence  orbital  occupancy  and  electron 
mobility. 

8.  The  elements  carbon  and  silicon  form  oxides 
with  similar  empirical  formulas :  C02  and  Si02. 
The  former  sublimes  at  —  78.5°C  and  the  latter 
melts  at  about  1700°C  and  boils  at  about  2200°C. 
From  this  large  difference,  propose  the  types  of 
solids  involved.  Draw  an  electron  dot  or  orbital 
representation  of  the  bonding  in  C02  that  is  con- 
sistent with  your  answer. 

9.  How  do  you  account  for  the  following  properties 
in  terms  of  the  structures  of  the  solids? 

(a)  Graphite  and  diamond  both  contain  carbon. 
Both  are  high  melting  yet  the  diamond  is 
very  hard  while  graphite  is  a  soft,  greasy 
solid. 

(b)  When  sodium  chloride  crystals  are  shattered, 
plane  surfaces  are  produced  on  the  frag- 
ments. 


(c)  Silicon  carbide  (carborundum)  is  a  very  high 
melting,  hard  substance,  used  as  an  abrasive. 

10.  If  you  were  given  a  sample  of  a  white  solid, 
describe  some  simple  experiments  that  you 
would  perform  to  help  you  decide  whether  or 
not  the  bonding  involved  primarily  covalent 
bonds,  ionic  bonds,  or  van  der  Waals  forces. 


11. 


12. 


13. 


14. 


If  elements  A,  D,  E,  and  J  have  atomic  numbers, 
respectively,  of  6,  9, 10,  and  11,  write  the  formula 
for  a  substance  you  would  expect  to  form  be- 
tween the  following: 


(a)  D  and  J; 

(b)  A  and  D; 

(c)  £>and  £>; 


(d)  E  and  E; 

(e)  J  and  J. 


In  each  case  describe  the  forces  involved  between 
the  building  blocks  in  the  solid  state. 

Consider  each  of  the  following  in  the  solid  state : 
sodium,  germanium,  methane,  neon,  potassium 
chloride,  water.  Which  would  be  an  example  of 

(a)  a  solid  held  together  by  van  der  Waals  forces 
that  melts  far  below  room  temperature; 

(b)  a  solid  with  a  high  degree  of  electrical  con- 
ductivity that  melts  near  200°C; 

(c)  a  high  melting,  network  solid  involving  co- 
valently  bonded  atoms; 

(d)  a  nonconducting  solid  which  becomes  a  good 
conductor  upon  melting; 

(e)  a  substance  in  which  hydrogen  bonding  is 
pronounced  ? 

Predict  the  order  of  increasing  melting  point  of 
these  substances  containing  chlorine:  HC1,  Cl2, 
NaCl,  CC14.  Explain  the  basis  of  your  prediction. 

Identify  all  the  types  of  bonds  you  would  expect 
to  find  in  each  of  the  following  crystals: 


(a)  argon, 

(b)  water, 

(c)  methane, 

(d)  carbon  monoxide, 

(e)  Si, 


(0  Al, 
(g)  CaCl2, 
(h)  KCIO3, 
(i)  NaCl, 
(j)   HCN. 


15.  Each  of  three  bottles  on  the  chemical  shelf  contains  a  colorless  liquid.  The  labels  have  fallen  off  the 
bottles.  They  read  as  follows. 


Label  No.  1 


Label  No.  2 


//-butanol 
CH3CH2CH2CH2OH 

mol  wt  =  74.12 


w-pentane 

CH3CH2CH2CH2CH| 
mol  wt  =  72.15 


Label  No.  3 

diethyl  ether 
CH3CH2OCH2CH3 

mol  wt  =  74.12 


QUESTIONS    AND    PROBLEMS  319 

The  three  bottles  are  marked  A,  B,  and  C,  and  a  series  of  measurements  were  made  on  the  three  liquids 
to  permit  identification,  as  follows. 

solubility 
m.p.  b.p.  density  AH  vap'n  in  water 


Liquid^  -131. 5°C  36.2°C  0.63  g/cc  85  cal/g  0.036  g/ 100  ml 

Liquid  B  -116  34.6  0.71  89.3  7.5 

Liquid  C  -89.2  117.7  0.81  141  7.9 


Which  liquid  should  be  given  Label  No.  1,  Label  No.  2,  Label  No.  3?  Explain  how  each  type  of  meas- 
urement influenced  your  choices. 

16.  Maleic  and  fumaric  acids  are  cis-  and  trans-  Maleic  acid  gives  up  its  first  proton  more  readily 

isomers  having  two  carboxyl  groups,  than  does  fumaric  acid.  However,  the  opposite 

is  the  case  for  the  second  proton.  Account  for 
HOOC— CH=CH— COOH  this  in  terms  of  structure. 


During  his  brilliant  scientific  career,  Peter  Debye  has 
added  richly  to  our  knowledge  of  the  structure  of  physical 
chemistry.  His  research  contributions  have  won  for  him 
awards  and  honorary  degrees  from  many  countries  and  he 
has  earned  unbounded  respect  wherever  men  seek  a  deeper 
understanding  of  nature. 

Born  in  Maastrecht,  the  Netherlands,  he  graduated  in 
electrical  engineering,  did  his  early  research  in  theoretical 
physics,  and  received  his  doctorate  at  the  University  of 
Munich  in  1908.  Three  years  later,  at  the  age  of  27,  he 
accepted  a  full  professorship  at  the  University  of  Zurich 
where  his  immediate  predecessor  was  Albert  Einstein.  Dur- 
ing this  year  he  developed  two  of  his  most  lasting  and 
fundamental  studies,  establishing  still  accepted  theories  of 
the  specific  heat  of  solids  and  of  the  interactions  among 
polar  molecules.  Shortly  thereafter,  he  returned  to  the 
Netherlands  as  professor  of  theoretical  physics  at  Utrecht. 
As  his  scientific  contributions  multiplied,  he  occupied  pro- 
fessorships successively  at  the  Universities  of  Goettingen 
{Germany),  Zurich  {Switzerland),  Leipzig  {Germany),  and 
Berlin  {Germany).  In  Berlin  he  was  appointed  director  of 
the  Max  Planck  Institute.  During  these  fruitful  years,  his 
research  ranged  through  X-ray  scattering,  interatomic  dis- 
tances, the  theory  of  electrolytes,  magnetic  cooling,  and 
dipole  theory;  this  work  won  for  him  the  Nobel  Prize  in 
Chemistry  for  1936. 


With  the  onset  of  World  War  II,  politics  began  to  inter- 
fere with  his  research.  Debye  was  actually  forbidden  to 
enter  the  Max  Planck  Institute  which  he  directed  because 
he  refused  to  accept  German  citizenship.  Despite  obstruc- 
tion by  the  German  government,  he  left  Germany  by  way 
of  Italy  and  came  to  the  United  States.  In  1940  he  was 
appointed  professor  of  chemistry  and  head  of  the  depart- 
ment of  chemistry  at  Cornell  University.  Six  years  later, 
he  became  an  American  citizen.  During  the  war  years  his 
research  turned  toward  the  structure  and  particle  size  of 
high  polymers. 

Attacking  this  new  field  with  his  usual  deep  insight  and 
characteristic  originality,  Debye  made  fundamental  and 
important  contributions  in  the  study  of  macromolecules. 
Now  professor  emeritus,  Debye  is  in  great  demand  as  a 
consultant  and  lecturer.  He  has  rare  ability  in  presenting 
the  most  complicated  subjects  in  a  fashion  that  gets  to  the 
heart  of  the  problem  with  penetrating  clarity.  Whenever  he 
speaks  at  a  scientific  meeting,  the  auditorium  is  filled  to 
capacity  with  an  audience  confident  they  will  hear  new  and 
interesting  ideas.  Inevitably  they  leave  inspired  and  stimu- 
lated by  their  contact  with  this  great  scientist — a  man  who 
can  delve  into  the  most  profound  aspects  of  nature  and 
bring  to  them  light  and  understanding. 


CHAPTER 


18 


The  Chemistry  of 
Carbon  Compounds 


The  synthesis  of  brazilin  would  have  no  industrial  value;  its  biological 
importance  is  problematical,  but  it  is  worthwhile  to  attempt  it  for  the 
sufficient  reason  that  we  have  no  idea  how  to  accomplish  the  task. 

ROBERT     ROBINSON,     1947 


The  compounds  of  carbon  furnish  one  of  the 
most  intriguing  aspects  of  all  of  chemistry.  One 
reason  they  interest  us  is  that  they  play  a  domi- 
nant role  in  the  chemistry  of  living  things,  both 
plant  and  animal.  Another  reason  is  that  there 
are  innumerable  carbon  compounds  useful  to 
man — dyes,  drugs,  detergents,  plastics,  perfumes, 
fibers,  fabrics,  flavors,  fuels — many  of  them 
tailored  to  suit  particular  needs.  Manufacture  of 


these  compounds  has  given  rise  to  a  huge  chemi- 
cal industry  requiring  millions  of  tons  of  raw 
materials  every  year. 

Where  do  we  find  the  enormous  quantities  of 
carbon  and  carbon  compounds  needed  to  feed 
this  giant  industry?  Let's  begin  our  study  of 
carbon  chemistry  by  taking  a  look  at  the  chief 
sources  of  carbon  and  carbon  compounds. 


18-1    SOURCES  OF  CARBON  COMPOUNDS 


18-1.1    Coal 

Coal,  a  black  mineral  of  vegetable  origin,  is 
believed  to  have  come  from  the  accumulation  of 
decaying  plant  material  in  swamps  during  pre- 
historic eras  when  warm,  wet  climatic  conditions 
permitted  rapid  growth  of  plants.  The  cycles  of 
decay,  new  growth,  and  decay,  caused  successive 
layers  of  plant  material  to  form  and  gradually 
build  up  into  vast  deposits.  The  accumulation  of 
top  layers  of  this  material  and  of  sedimentary 


rocks  excluded  air  from  the  lower  material  and 
subjected  it  to  enormous  pressures.  In  time  the 
layers  were  compressed  into  hard  beds  composed 
chiefly  of  the  carbon  that  was  present  in  the  origi- 
nal plants,  and  containing  appreciable  amounts 
of  oxygen,  hydrogen,  nitrogen,  and  some  sulfur. 
Thus,  coal  is  not  pure  carbon.  The  "hardest" 
coal,  anthracite,  may  contain  from  85  to  95% 
carbon;  the  "softest,"  peat,  is  not  really  coal  at 
all  but  one  of  the  early  stages  in  the  geological 

321 


322 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


history  of  coal.  Peat  still  contains  unchanged 
plant  remains  and  may  contain  no  more  than  50 
to  60%  carbon. 

When  coal  is  heated  to  a  high  temperature  in 
the  absence  of  air,  it  undergoes  decomposition: 
volatile  products  (coal  gas  and  coal  tar)  distill 
away  and  a  residue  called  coke  remains.  Coke  is 
a  valuable  industrial  material  which  finds  its 
chief  use  in  the  reduction  of  iron  ore  (iron  oxide) 
to  iron  for  the  manufacture  of  steel.  Coke  is 
essentially  carbon  that  still  contains  the  mineral 
substances  that  are  present  in  all  coals  (and  form 
the  ash  that  results  when  coal  or  coke  is  burned). 

About  eight  gallons  of  coal  tar  are  obtained 
from  a  ton  of  coal.  Coal  tars  are  very  complex 
mixtures ;  over  200  different  carbon  compounds 
have  been  isolated  from  them.  While  the  great 
value  of  coal  to  mankind  has  been  as  a  fuel,  a 
source  of  energy,  the  many  substances  in  coal 
gas  and  coal  tar  make  coal  also  an  important 
source  of  chemical  raw  materials. 

18-1.2    Petroleum 

Petroleum  is  a  complex  mixture  which  may 
range  from  a  light,  volatile  liquid  to  a  heavy, 
tarry  substance.  Petroleum  also  has  its  origin  in 
living  matter  that  has  undergone  chemical 
changes  over  the  course  of  geological  time.  It  is 
found  in  porous  rock  formations  called  oil  pools, 
between  impervious  rock  formations  that  seal 
off  the  pools.  When  a  pool  is  tapped,  the  oil 
flows  through  the  porous  structure  (driven  by 
subterranean  gas  or  water  pressure)  and  so  is 
brought  to  the  surface. 


18-1.3    Natural  Gas 

Natural  gas  is  a  mixture  of  low  molecular  weight 
compounds  of  hydrogen  and  carbon  (hydrocar- 
bons) found  in  underground  "fields"  of  sand- 
stone or  other  porous  rock.  This  gas  escapes  to 
the  surface  of  the  earth  when  the  field  is  tapped 
by  drilling. 

18-1.4    Certain  Plant  and  Animal  Products 

Plants  and  animals  are  themselves  highly  effec- 
tive chemical  factories  and  they  synthesize  many 
carbon  compounds  useful  to  man.  These  include 
sugars,  starches,  plant  oils  and  waxes,  fats,  gela- 
tin, dyes,  drugs,  and  fibers. 

Because  all  of  these  sources  of  carbon  com- 
pounds ultimately  find  their  origin  in  living 
matter,  plant  or  animal,  the  chemistry  of  carbon 
is  called  organic  chemistry.  Compounds  contain- 
ing carbon  are  called  organic  compounds. 
This  term  includes  all  compounds  of  carbon 
except  C02,  CO,  and  a  handful  of  ionic  sub- 
stances (for  example,  sodium  carbonate,  Na2C03, 
and  sodium  cyanide,  NaCN).  You  may  wonder 
how  many  organic  substances  are  known.  The 
number  is  actually  so  large  it  is  difficult  to  pro- 
vide a  reliable  estimate.  A  great  many  more 
compounds  of  carbon  have  been  studied  than  of 
any  other  element  except  hydrogen  (hydrogen  is 
present  in  most  carbon  compounds).  There  are 
undoubtedly  over  one  million  different  carbon 
compounds  known.  The  number  of  new  organic 
compounds  synthesized  in  one  year  (about 
100,000  per  year)  exceeds  the  total  number  of 
compounds  known  that  contain  no  carbon! 


18-2     MOLECULAR  STRUCTURES  OF  CARBON  COMPOUNDS 


How  can  there  be  so  many  compounds  contain- 
ing this  one  element?  The  answer  lies  in  the 
molecular  structures.  We  shall  find  that  carbon 
atoms  have  an  exceptional  tendency  to  form 
covalent  bonds  to  other  carbon  atoms,  forming 
long  chains,  branched  chains,  and  rings  of  atoms. 
Each  different  atomic  arrangement  gives  a  mole- 


cule with  distinctive  properties.  To  understand 
why  a  particular  substance  has  its  characteristic 
properties,  its  structure  must  be  known.  Thus 
the  determination  of  the  molecular  structure  of 
carbon  compounds  is  one  of  the  central  prob- 
lems of  organic  chemistry.  Let's  see  how  it  is 
done. 


SEC.    18-2    I    MOLECULAR    STRUCTURES    OF    CARBON    COMPOUNDS 


323 


18-2.1    The  Composition  and  Structure 
of  Carbon  Compounds 

Ethane  and  ethanol*  are  two  common  carbon 
compounds.  Ethane  is  a  gas  that  usually  makes 
up  about  10%  of  the  household  gas  used  for 
heating  and  cooking.  Its  useful  chemistry  is  al- 
most wholly  restricted  to  the  combustion  reac- 
tion. Ethanol  is  a  liquid  that  takes  part  in  a 
variety  of  useful  chemical  reactions.  It  has  great 
value  in  the  manufacture  of  chemicals  and  it 
bears  little  chemical  resemblance  to  ethane.  Yet, 
the  similarity  of  the  two  names,  ethane  and 
ethanol,  suggests  that  these  compounds  are  re- 
lated. This  is  so.  To  understand  how  they  are 
related  and  why  their  chemistries  are  so  different, 
we  must  learn  their  molecular  structures.  We 
must  find  out  what  kinds  of  atoms  are  present  in 
each  substance,  how  many  atoms  there  are  per 
molecule,  and  their  bonding  arrangement.  Usu- 
ally many  experiments  must  be  performed  before 
the  molecular  structure  of  a  compound  is  known 
with  certainty.  This  fascinating  problem  of  car- 
bon chemistry  involves  three  basic  experimental 
steps:  to  determine  the  empirical  formula,  then 
the  molecular  formula,  and  finally  the  structural 
formula.  First  we  shall  review  the  information 
conveyed  by  each  of  these  formulas,  using  eth- 
ane as  an  example.  Then,  in  Section  18-2.2  we 
will  consider  what  experiments  are  used  in  the 
determination  of  each  type  of  formula,  using 
ethanol  as  an  example. 

EMPIRICAL    FORMULA 

The  empirical  formula  tells  only  the  relative 
number  of  atoms  of  each  element  in  a  molecule. 
For  example,  consider  ethane.  Analysis  shows 
that  this  is  a  compound  of  carbon  and  hydrogen 
and  that  there  are  three  hydrogen  atoms  for 

*  Ethanol  is  another  name  for  the  substance  ethyl 
alcohol. 


H 

\ 

H C 

/ 
H 


/* 

C—H 


CH3  CH3 


Fig.  18-1.  Structural  formulas  of  ethane,  CV/e. 


every  carbon  atom.  Its  empirical  formula,  there- 
fore, is  CH3. 

MOLECULAR    FORMULA 

The  molecular  formula  tells  the  total  number  of 
atoms  of  each  element  in  a  molecule.  Ethane  is 
found  to  have  a  molecular  weight  of  30.  This 
molecular  weight  together  with  the  empirical 
formula  tells  us  the  molecular  formula.  It  cannot 
be  CH3 — this  compound  would  have  a  molecular 
weight  of  15.  The  molecular  formula  C2H6  also 
has  three  hydrogen  atoms  per  carbon  atom. 
Since  it  has  a  molecular  weight  of  30,  it  has  both 
the  correct  empirical  formula  and  the  correct 
molecular  weight.  Ethane  has  the  molecular  for- 
mula C2H6. 


EXERCISE  18-1 

Write  the  molecular  formula  for  the  carbon- 
hydrogen  compound  containing  two  carbon 
atoms  and  having  empirical  formula  CH2.  What 
is  its  molecular  weight? 


STRUCTURAL    FORMULA 

The  structural  formula  tells  which  atoms  are 
connected  in  the  molecule.  In  ethane,  the  two 
carbon  atoms  are  linked  and  three  hydrogen 
atoms  are  attached  to  each  carbon  atom.  Various 
ways  of  representing  its  structural  formula  are 
shown  in  Figure  18-1. 

The  formulas  in  Figure  18-1  all  represent  the 
same  structure;  the  choice  of  which  formula  to 
use  depends  upon  what  feature  of  the  structure 
is  to  be  emphasized.  The  first  and  second  draw- 
ings emphasize  the  three-dimensional  nature  of 


324 


THB  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


ethane;  the  third  is  a  simpler  way  of  doing  the 
same  thing;  and  the  last  formula  merely  shows 
that  three  hydrogens  are  attached  to  each  carbon 
atom.  It  is  not  at  all  difficult  to  decide  that 
CH3CH3  must  be  the  structural  formula  for  eth- 
ane. By  the  bonding  rules  developed  in  Chapter 
16,  we  know  that  carbon  is  always  surrounded 
by  four  electron  pair  bonds  and  that  a  hydrogen 
atom  forms  only  one  covalent  bond.  There  is  no 
structure  other  than  the  one  shown  in  Figure 
18-1  in  which  two  carbon  atoms  and  six  hydro- 
gen atoms  can  be  bound  together  and  satisfy  all 
the  bonding  rules. 

18-2.2    Experimental  Determination 
of  Molecular  Structure 

We  have  seen  three  steps  in  fixing  molecular 
structure.  What  experiments  are  involved  in  each 
of  these  steps?  Let's  investigate  the  nature  of 
these  experiments  using  ethanol  as  a  second  ex- 
ample. 

DETERMINING   THE   EMPIRICAL   FORMULA 

In  order  to  determine  the  empirical  formula  of  a 
compound,  you  must  first  find  out  just  what  ele- 
ments are  present  in  it.  Sometimes  this  is  done 
simply  by  burning  some  of  the  compound  in 
pure  oxygen.  If  the  compound  contains  only 
carbon  and  hydrogen,  only  carbon  dioxide  and 
water  will  be  produced.  If  the  compound  con- 
tains some  nitrogen  as  well,  nitrogen  gas  or  one 
of  the  nitrogen  oxides  will  be  produced  in  the 
combustion  and  can  be  identified.  Another  way 
of  finding  out  which  elements  are  in  a  compound 
is  to  allow  the  compound  to  react  with  hot, 
liquid  sodium  metal.  If  the  compound  contains 
nitrogen,  sodium  cyanide,  NaCN,  will  be 
formed;  if  it  contains  sulfur,  sodium  sulfide, 
Na2S,  will  be  a  product.  Once  such  reactions 
show  which  elements  are  in  the  compound,  rela- 
tive numbers  of  atoms  of  each  element  (the 
empirical  formula)  can  be  determined. 

The  usual  method  for  finding  the  empirical 
formula  is  simply  illustrated  with  ethanol,  a 
compound  containing  only  carbon,  hydrogen, 
and  oxygen.  A  weighed  amount  of  the  pure  com- 
pound is  completely  burned  in  oxygen  to  give 
carbon  dioxide  (from  the  carbon)  and  water 


(from  the  hydrogen).  The  weight  of  the  carbon 
dioxide  reveals  how  much  carbon  was  in  the 
weighed  amount  of  sample.  The  weight  of  water 
reveals  how  much  hydrogen  was  in  the  sample. 
The  remainder  of  the  sample  must  have  been 
oxygen.  (There  is  no  need  to  discuss  here  the 
procedures  for  compounds  containing  halogens, 
nitrogen,  or  sulfur,  for  they  are  quite  similar.) 
Suppose  we  burn  46  grams  of  ethanol.  Collec- 
tion of  the  products  yields  88  grams  of  carbon 
dioxide  and  54  grams  of  water.  We  wish  to  learn 
the  relative  numbers  of  carbon,  hydrogen,  and 
oxygen  atoms  in  the  compound,  and  we  can  do 
this  by  calculating  the  number  of  moles  of  car- 
bon dioxide  and  water  produced  by  the  combus- 
tion of  the  46  gram  sample.  Therefore,  we 
calculate: 

Number  of     =        88  g        =       88  g 
moles  C02      mol  wt  C02      44  g/mole 


=  2.0  moles  C02 


54  g 


54  g 


Number  of     _ 
moles  H20      mol  wt  H20       18  g/mole 

=  3.0  moles  H20 

Now  we  can  make  the  following  statements 
about  the  compound  ethanol: 

46  grams  of  ethanol  yield  two  moles  of  C02  and 
three  moles  of  H20 

or 

46  grams  of  ethanol  contain  two  moles  of  carbon 
atoms  and  six  moles  of  hydrogen  atoms 

or 

46  grams  of  ethanol  contain  24  grams  of  carbon 
atoms  and  6  grams  of  hydrogen  atoms. 

Thus  we  have  accounted  for  (24  +  6)  =  30  g 
of  the  46  g  we  started  with.  The  remainder  of  the 
sample,  (46  —  30)  =  16  g,  must  have  been  oxy- 
gen. This  is  just 

16  g  16  g 


at.  wt  oxygen      16  g/mole 


=  1.0  mole  oxygen  atoms 


Summarizing,  we  know  that  46  grams  of  the 
compound  ethanol  contain 

two  moles  of  carbon  atoms, 
six  moles  of  hydrogen  atoms, 
one  mole  of  oxygen  atoms. 


SEC.     18-2    I    MOLECULAR    STRUCTURES    OF    CARBON    COMPOUNDS 


325 


Since  the  relative  number  of  atoms  of  each  ele- 
ment in  the  compound  is  the  same  as  the  relative 
number  of  moles  of  atoms  of  each  element  in  the 
sample,  we  can  say  that  the  empirical  formula 
of  ethanol  is 

QHeO,    or    C2HeO 

This  example  has  been  much  simplified  by  our 
selection  of  46  grams  of  sample.  In  practice,  less 
than  a  gram  of  sample  is  used  and  whole  num- 
bers of  moles  are  not  obtained.  A  typical  set  of 
experimentally  obtained  data  is  given  in  Exer- 
cise 18-2. 


EXERCISE  18-2 

Automobile  antifreeze  often  contains  a  com- 
pound called  ethylene  glycol.  Analysis  of  pure 
ethylene  glycol  shows  that  it  contains  only  car- 
bon, hydrogen,  and  oxygen.  A  sample  of  ethylene 
glycol  weighing  15.5  mg  is  burned  and  the 
weights  of  C02  and  H20  resulting  are  as  follows: 

weight  of  sample  burned  =  15.5  mg, 
weight  of  C02  formed  =  22.0  mg, 
weight  of  H20  formed      =  13.5  mg. 

What  is  the  empirical  formula  of  ethylene  glycol? 


DETERMINING    THE    MOLECULAR 
FORMULA 

Now  we  know  that  the  relative  numbers  of  atoms 
in  ethanol  are  two  carbon  to  six  hydrogen  to  one 
oxygen.  We  do  not  know  yet  whether  the  molec- 
ular formula  is  C2HeO,  C,Hi202,  C6Hi,03,  or 
some  other  multiple  of  the  empirical  formula, 
C2HeO. 

This  returns  us  to  the  problem  of  the  deter- 
mination of  molecular  weight.  Avogadro's  Hy- 
pothesis provides  one  of  the  convenient  ways  of 
measuring  molecular  weight  if  the  substance  can 
be  vaporized. 

This  was  exactly  the  measurement  you  made 
in  Experiment  6 — it  is  called  the  vapor-density 
method  for  molecular  weight  determination.  To 


apply  the  method  to  a  liquid,  such  as  ethanol, 
a  temperature  above  the  boiling  point  is  needed. 
A  weighed  amount  of  liquid  is  placed  in  a  gas 
collecting  device  held  at  an  easily  regulated 
temperature.  For  example,  a  steam  condenser 
around  the  device  provides  a  convenient  way  of 
holding  the  temperature  at  100°C.  When  the 
substance  has  vaporized  completely,  its  pressure 
and  volume  are  measured  (perhaps  using  equip- 
ment like  that  shown  in  Figure  9-1  of  the  Labo- 
ratory Manual).  This  provides  a  measurement  of 
the  weight  per  unit  volume  of  gaseous  ethanol 
at  a  known  temperature  and  pressure.  Again, 
this  weight  is  compared  with  the  weight  of  the 
same  volume  of  a  reference  gas  (usually  02)  at 
the  same  temperature  and  pressure. 

Suppose  such  a  vapor-density  measurement 
shows  that  a  given  volume  of  ethanol  at  100°C 
and  one  atmosphere  weighs  1.5  times  as  much 
as  the  same  volume  of  oxygen  gas  at  100°C  and 
one  atmosphere.  Since  equal  volumes  contain 
equal  numbers  of  molecules  at  the  same  tem- 
perature and  pressure  (Avogadro's  Hypothesis), 
one  molecule  of  the  unknown  gas  must  weigh 
1.5  times  the  weight  of  a  molecule  of  02.  There- 
fore, 

mol  wt  of  the  unknown  gas  =  (1.5)  X  (molwt02) 

=  1.5  X  32  =  48g/mole 

Even  though  this  number  is  not  very  accurate, 
it  will  suffice  for  the  purpose  of  deciding  that  the 
molecular  formula  is  C2HeO  (with  molecular 
weight  46.07  g/mole),  not  (C2H60)2  (with  mo- 
lecular weight  92.14  g  mole),  or  (C2H60)3  (with 
molecular  weight  138.21  g  mole),  or  any  higher 
multiple  of  empirical  formula  units. 

There  are  two  other  common  methods  for  learning  the 
molecular  weight  of  an  unknown  compound.  They  are 
important  in  the  study  of  compounds  which  are  not 
readily  vaporized  (as  is  required  in  the  vapor-density 
method).  These  methods  are  called  the  boiling  point  eleva- 
tion and  freezing  point  lowering  methods.  We  have  already 
mentioned  in  Section  5-2.1  that  a  solution  of  salt  water 
has  a  higher  boiling  point  than  that  of  pure  water.  The 
boiling  point  elevation  for  a  given  solute  concentration 
(expressed  in  moles)  is  almost  independent  of  the  choice 
of  solute.  Hence  this  temperature  measurement  can  be 
readily  interpreted  in  terms  of  a  molar  concentration. 
From  the  weight  of  sample  used  in  preparation  of  the 
solution,  the  molecular  weight  can  be  calculated. 


326 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


Exactly  the  same  type  of  behavior  is  found  for  the 
freezing  point  of  a  solution  except  that  the  freezing 
point  is  lower  than  that  of  the  pure  solvent.  Thus  we  have 
two  methods  for  molecular  weight  determination  which 
are  applicable  to  compounds  with  such  low  vapor  pres- 
sure or  which  decompose  so  readily  that  the  vapor  density 
method  cannot  be  used. 


EXERCISE  18-3 

Ethylene  glycol,  the  example  treated  in  Exercise 
18-2,  has  an  empirical  formula  of  CHsO.  (Is  this 
what  you  obtained?)  A  sample  weighing  0.49 
gram  is  vaporized  completely  at  200°C  and  at 
one  atmosphere  pressure.  The  volume  measured 
under  these  conditions  is  291  ml.  This  same  vol- 
ume, 291  ml,  of  oxygen  gas  at  200°C  and  one 
atmosphere  weighs  0.240  gram.  What  is  the 
molecular  formula  for  ethylene  glycol,  CH^O, 
C2H602,  C3H9O3,  C4Hl2Ou  or  some  higher  mul- 
tiple of  CH30? 


ethane  is  a  saturated*  compound.  Neither  can 
the  oxygen  atom  just  be  attached  somehow  to  a 
hydrogen  atom.  Each  hydrogen  atom  in  ethane 
has  its  bonding  capacity  already  satisfied. 

Rather  than  trying  to  find  the  structural  for- 
mula of  ethanol  by  tacking  an  oxygen  atom  to 
ethane,  let  us  start  with  the  oxygen  atom  and  see 
how  we  might  build  a  molecule  around  it,  using 
the  two  carbon  atoms  and  six  hydrogen  atoms 
which  are  at  our  disposal.  We  already  know  that 
the  oxygen  atom  is  commonly  divalent,  and  that, 
as  in  water,  it  makes  bonds  to  hydrogen  atoms. 
Let  us  start  our  molecular  construction  with  a 
bond  between  one  hydrogen  atom  and  the  oxy- 
gen atom: 

O— H 

The  other  bond  which  the  oxygen  atom  can 
make  must  be  to  a  carbon  atom,  since  if  it  were 
to  another  hydrogen  atom  we  would  simply  have 
a  water  molecule.  Therefore  we  write 


determining  the  structural 
formula:  ethanol 

The  determination  of  how  the  atoms  of  a  mole- 
cule are  connected  is  the  most  important  prob- 
lem in  identifying  an  unknown  compound.  It  can 
be  as  exciting  as  a  detective  story,  with  the 
chemical  and  physical  properties  furnishing  the 
clues.  With  the  right  collection  of  clues,  the 
chemist  can  ascertain  the  identity  of  the  mole- 
cule. 

What,  then,  is  the  structure  of  ethanol?  First 
we  must  learn  its  empirical  formula.  Analysis 
shows  that  the  empirical  formula  is  C2H60.  The 
molecular  formula  is  fixed  by  a  vapor  density 
measurement.  The  molecular  weight  is  found  to 
be  46,  showing  that  the  molecular  formula  is  the 
same  as  the  empirical  formula,  C2H60.  It  remains 
to  discover  the  arrangement  and  connections  of 
the  atoms.  We  might  begin  by  eliminating  some 
structures  which  we  can  be  sure  are  incorrect. 
Ethanol  is  not  simply  ethane  with  an  oxygen 
atom  somehow  attached  to  a  carbon  atom,  be- 
cause in  ethane  all  of  the  four  bonds  of  each 
carbon  are  satisfied,  and  so  there  is  no  way  in 
which  an  additional  bond  can  form.  We  say  that 


O— H 

The  carbon  atom  we  have  added  must  form 
three  additional  bonds  to  satisfy  its  tetravalent 
bonding  capacity.  If  all  of  these  bonds  were  to 
hydrogen  atoms,  we  would  have  the  completed 
molecule  CH3OH,  and  would  also  have  two  hy- 
drogen atoms  and  a  carbon  atom  left  over. 
Therefore,  one  of  the  bonds  our  first  carbon 
atom  forms  must  be  to  the  other  carbon  atom, 
and  the  two  other  bonds  must  be  to  hydrogen 
atoms.  We  have  then 

H 

/ 
C— C— H 

\ 
O— H 

We  can  easily  complete  the  structure  by  adding 
three  bonds  from  our  last  carbon  atom  to  the 


*  This  usage  of  the  word  "saturated"  shows  that  chem- 
ists, like  other  people,  sometimes  use  the  same  word  with 
two  entirely  different  meanings.  On  p.  164  this  word  was 
used  to  describe  a  solution  which  contains  the  equilib- 
rium concentration  of  a  dissolved  substance.  As  used 
here,  in  reference  to  organic  compounds,  it  means  that 
all  bonds  to  carbon  are  single  bonds  and  they  are  all 
formed  with  hydrogen  or  other  carbon  atoms. 


SEC.    18-2    I   MOLECULAR    STRUCTURES   OF    CARBON    COMPOUNDS 


327 


three  hydrogen  atoms  we  have  left.  The  result  is 

H  H 

\  / 

H— C— C— H 

H  O— H 

We  have  now  used  all  six  hydrogen  atoms,  the 
two  carbon  atoms  and  the  oxygen  atom  which 
the  molecular  formula  of  ethanol  requires.  All 
the  bonding  rules  are  satisfied,  so  the  structure 
we  have  written  must  be  taken  as  a  possible 
structural  formula  for  ethanol.  However,  we 
must  also  decide  whether  this  is  the  only  possible 
structural  formula  for  a  molecule  with  molecular 
formula  C2H60.  A  little  reflection  shows  it  is  not. 
Instead  of  having  the  oxygen  atom  form  one 
bond  to  carbon  and  one  to  hydrogen,  why  not 
start  with  two  oxygen-carbon  bonds? 

O 
/   \ 

c       c 

We  have  six  hydrogen  atoms  at  our  disposal,  and 

each  of  the  carbon  atoms  must  form  three  more 

bonds.  Therefore  we  complete  the  structure  by 

writing 

O 

H  / 

\   / 
C 

/  \ 
H  H 


H 


H 


\ 


H 


Satisfy  yourself  that  this  structure  violates  no 
bonding  rules,  and  conforms  to  the  empirical 
and  molecular  formula  of  ethanol. 

We  have  now  found  all  possible  structural 
formulas  for  the  ethanol  molecule.  The  oxygen 
atom  is  either  directly  bonded  to  one  carbon 
atom  or  to  two  carbon  atoms.  Once  a  choice 
between  these  two  possibilities  is  made,  the 
structure  of  the  rest  of  the  molecule  can  be  deter- 
mined from  the  molecular  formula  and  the  bond- 
ing rules.  The  two  possible  structures  are  shown 
in  Figure  18-2.  Such  compounds  with  the  same 
molecular  formula  but  different  structural  formu- 
las are  called  structural  isomers.  The  existence 
of  the  two  compounds  1  and  2  was  known  long 
before  their  structures  were  clarified.  Hence  the 
existence  of  these  isomers  perplexed  chemists  for 
decades.  Now  we  recognize  the  crucial  impor- 


CH3  CH2  OH 


(1) 


CH30CH3  (2) 

Fig.  18-2.  The  structures  of  the  C~HaO  isomers. 


tance  of  learning  the  structural  formulas  (as  well 
as  molecular  formulas)  to  identify  the  isomers. 

So  our  problem  is  to  decide  whether  ethanol 
has  structure  1  or  structure  2.  How  can  we  tell 
which  is  correct?  Let  us  see  what  preliminary 
ideas  we  can  get  from  an  examination  of  the 
structural  formulas. 

In  structure  2,  all  of  the  hydrogen  atoms  are 
the  same — each  hydrogen  atom  is  bonded  to  a 
carbon  which  is,  in  turn,  bonded  to  the  oxygen 
atom.  In  structure  1,  one  of  the  hydrogen  atoms 
is  quite  different  from  any  of  the  others:  it  is 
bonded  to  oxygen  and  not  to  carbon.  Of  the 
remaining  five,  two  are  similarly  placed,  on  the 
carbon  bonded  tc  oxygen,  and  three  are  on  the 
other  carbon.  Structures  1  and  2  should  have 


328 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  |  CHAP.  18 


quite  different  chemistries.  Which  one  should 
correspond  to  the  chemistry  of  ethanol? 

We  can  offer  several  kinds  of  evidence.  Some 
comes  from  the  behavior  of  ethanol  in  chemical 
reactions  and  some  from  the  determination  of 
certain  physical  properties.  Let's  consider  the 
reactions  first. 

Clean  sodium  metal  reacts  vigorously  with 
ethanol,  giving  hydrogen  gas  and  an  ionic  com- 
pound, sodium  ethoxide,  with  empirical  formula 
C2H5ONa.  The  reaction  is  quite  similar  to  the 
behavior  of  sodium  and  water  which  yields  hy- 
drogen and  the  ionic  compound  sodium  hy- 
droxide, NaOH.  This  suggests,  but  certainly  does 
not  prove,  that  ethanol  shows  some  structural 
similarity  to  water.  In  water  we  have  two  hydro- 
gen atoms  bonded  to  oxygen  atoms,  and  in  struc- 
ture 1  we  have  one  hydrogen  atom  bonded  to 
oxygen.  This  bit  of  chemical  evidence  suggests 
ethanol  has  structure  1 . 

More  quantitative  evidence  can  be  obtained 
by  carrying  out  the  reaction  between  an  excess 
of  sodium  and  a  weighed  amount  of  ethanol  and 
measuring  the  amount  of  hydrogen  gas  evolved. 
When  this  is  done  it  is  found  that  46  grams  of 
ethanol  (one  mole)  will  produce  only  §  mole  of 
hydrogen  gas.  We  can  therefore  write  a  balanced 
chemical  equation  for  the  reaction  of  sodium 
with  ethanol: 

NafsJ  +  C2H60(J J  — ►-  \Wg)  +  C2H5ONa(sj     (7) 

This  equation  expresses  the  fact  that  one  mole 
of  ethanol  produces  \  mole  of  hydrogen  gas. 
Hence,  one  mole  of  ethanol  must  contain  one 
mole  of  hydrogen  atoms  that  are  uniquely  ca- 
pable of  undergoing  reaction  with  sodium.  Ap- 
parently one  molecule  of  ethanol  contains  one 
hydrogen  atom  that  is  capable  of  reacting  with 
sodium  and  five  that  are  not.  Let  us  now  consider 
structures  1  and  2  in  the  light  of  this  information. 
In  structure  2  all  six  of  the  hydrogen  atoms  are 
structurally  equivalent,  whereas  in  structure  1 
there  is  one  hydrogen  atom  in  the  molecule 
which  is  structurally  unique — that  is,  the  one 
bonded  to  the  oxygen  atom.  Structure  1  is  there- 
fore consistent  with  the  experimental  fact  that 
only  one  hydrogen  atom  per  molecule  of  ethanol 
will  react  with  sodium  and  structure  2  is  not. 


We  can  find  further  evidence  that  structure  1, 
CH3CH2OH,  is  the  correct  structural  formula 
for  the  substance  known  as  ethanol.  It  is  known 
that  compounds  which  contain  only  carbon  and 
hydrogen  (such  as  ethane,  C2H6)  do  not  react  at 
all  readily  with  metallic  sodium  to  produce  hy- 
drogen gas.  In  these  compounds  the  hydrogen 
atoms  are  all  bonded  to  carbon  atoms  (see  Fig- 
ure 18-1),  so  we  can  make  the  deduction  that,  in 
general,  hydrogen  atoms  which  are  bonded  to 
carbon  atoms  do  not  react  with  sodium  to 
produce  hydrogen  gas.  In  structure  2,  CH3OCH3, 
all  the  hydrogen  atoms  are  bonded  to  carbon 
atoms,  so  we  do  not  expect  a  compound  with 
this  structure  to  react  with  sodium.  Ethanol  re- 
acts with  sodium,  so  it  is  unlikely  that  ethanol 
has  structure  2. 

Let  us  consider  one  other  reaction  of  ethanol. 
If  ethanol  is  heated  with  aqueous  HBr,  we  find 
that  a  volatile  compound  is  formed.  This  com- 
pound is  only  slightly  soluble  in  water  and  it 
contains  bromine:  its  molecular  formula  is  found 
by  analysis  and  molecular  weight  determination 
to  be  C2H5Br  (ethyl  bromide,  or  bromoethane). 
With  the  aid  of  the  bonding  rules,  we  can  see 
that  there  is  only  one  possible  structure  for  this 
compound.  This  result  is  verified  by  the  fact  that 
only  one  isomer  of  C2H5Br  has  ever  been  dis- 
covered. 

Now  we  can  ask  how  this  chemical  reaction 
furnishes  a  clue  to  the  structure  of  ethanol. 
Structure  1  could  give  structure  3  in  Figure  18-3 
merely  by  breaking  the  carbon-oxygen  bond. 


Fig.  18-3.  The  structural  formula  of  ethyl  bromide 
(bromoethane). 


CHjCH2Bf         (3) 


SEC.     18-2    I    MOLECULAR    STRUCTURES    OF    CARBON    COMPOUNDS 


329 


Convince  yourself  of  this  fact  by  writing  an 
equation  using  the  structural  formulas  1  and  3. 
In  contrast,  bromoethane  can  be  obtained  from 
structure  2  only  through  a  complicated  rear- 
rangement. Two  carbon-oxygen  and  one  carbon- 
hydrogen  bond  would  have  to  be  broken.  Expe- 
rience shows  that  such  complicated  reshufflings 
of  atoms  rarely  occur.  Therefore,  the  reaction 
between  ethanol  and  hydrobromic  acid,  HBr,  to 
form  bromoethane  provides  more  evidence  that 
ethanol  has  structure  1. 

The  evidence  cited  so  far  has  been  associated 
with  the  chemistry  of  ethanol.  Its  boiling  point 
provides  a  different  sort  of  information  also 
leading  to  structure  1.  Ethanol  is  a  liquid  with  a 
boiling  point  of  78°C.  This  can  be  compared 
with  the  boiling  point  of  ethane,  C2HC,  which  is 
-172°C,  and  to  that  of  water,  100°C.  Plainly, 
the  substance  ethanol  is  more  like  water  than 
like  ethane,  as  far  as  boiling  point  is  concerned. 
Once  again  this  can  be  understood  better  in 
terms  of  structure  1.  Structure  1  has,  in  common 
with  H20,  oxygen  linked  to  hydrogen.  The  high 
boiling  point  of  water  is  explained  in  terms  of  an 
abnormally  large  intermolecular  attraction  of 
such  an  O — H  group  to  surrounding  water  mole- 
cules. The  interaction  is  called  hydrogen  bonding 
(see  Section  17-2.6).  If  ethanol  also  has  the  O — H 
group  (as  in  structure  1)  then  it  too  can  exert  the 
same  abnormally  large  attraction  to  neighboring 
ethanol  molecules.  Thus  structure  1  provides  an 
explanation  of  the  fact  that  the  boiling  point  of 
ethanol  is  so  high. 

This  possibility  of  forming  hydrogen  bonds 
should  cause  a  strong  attraction  between  water 
and  a  compound  of  structure  1.  If  there  is  strong 
attraction,  then  ethanol  should  have  high  solu- 
bility in  water.  Experiment  shows  that  they  are 
miscible — they  dissolve  in  all  proportions.  Again 
the  evidence  tends  to  strengthen  belief  in  struc- 
ture 1. 

Notice  that  our  attempt  to  determine  the 
structural  formula  of  ethanol  has  involved  the 
consideration  of  a  variety  of  types  of  evidence. 
Others  could  be  listed  as  well — for  example,  the 
infrared  spectrum  of  the  liquid  and  the  X-ray 
diffraction  pattern  of  the  solid  add  strong  sup- 
port for  structure  1 .  No  one  fact  by  itself  gives 


absolute  proof  of  the  structure,  but  all  the  facts 
considered  together  show  that  1  is  unquestion- 
ably the  correct  structure  for  ethanol.  A  com- 
parable set  of  experiments  shows  that  another 
compound  with  the  formula  C2H60  has  proper- 
ties consistent  with  structure  2.  This  compound 
is  called  dimethyl  ether. 

EXERCISE  18-4 

Ethylene  glycol  has  empirical  formula  CH30  and 
molecular  formula  C2H602.  Using  the  usual 
bonding  rules  (carbon  is  tetravalent;  oxygen  is 
divalent;  hydrogen  is  monovalent),  draw  some 
of  the  structural  formulas  possible  for  this  com- 
pound. 

EXERCISE  18-5 

Decide  which  of  your  structures  in  Exercise  18-4 
best  fits  the  following  list  of  properties  observed 
for  pure  ethylene  glycol. 

(a)  It  is  a  viscous  (syrupy)  liquid  boiling  at 
197°C. 

(b)  It  is  miscible  with  water,  that  is,  it  dissolves, 
forming  solutions,  in  all  proportions. 

(c)  It  is  miscible  with  ethanol. 

(d)  It  reacts  with  sodium  metal,  producing  hy- 
drogen gas. 

(e)  A  6.2  gram  sample  of  ethylene  glycol  reacts 
with  an  excess  of  sodium  metal  to  produce 
2.4  liters  of  hydrogen  gas  at  one  atmosphere 
pressure  and  25°C. 


18-2.3    The  Ethyl  Group 

All  of  the  reactions  and  the  physical  properties 
of  ethanol  have  been  explained  on  the  basis  of 
the  behavior  of  the  OH  group  in  structure  1, 
CH3CH2OH.  This  is  true  of  most  of  the  reactions 
of  ethanol — the  reaction  centers  at  the  OH  group 
(which  is  called  the  hydroxyl  group),  and  the 
remainder  of  the  molecule,  C2H5 — ,  remains  in- 
tact. The  reactions  lead  to  the  suggestion  that 
there  are  two  parts  in  the  molecule  of  ethanol,  the 
H    H 

I       I 
H — C — C —  group,  which  is  unchanged  during 

H    H 


330 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


reactions,  and  the  —OH  group,  which  can 
change.  This  concept  of  the  structural  integrity 
of  the  hydrocarbon  group  is  an  important  one 
in  organic  chemistry.  It  focuses  attention  on  the 
groups  that  do  change,  the  so-called  functional 
groups.  If  the  chemistry  of  a  particular  functional 
group  is  understood  for  one  compound,  it  is  a 
good  assumption  that  this  same  chemistry  will  be 
found  for  another  compound  containing  this 
same  functional  group.  Thus,  compounds  with 
the  OH  group  are  given  a  family  name,  alcohols. 
The  rest  of  the  molecule,  the  carbon  skeleton, 
has  relatively  little  effect  and  it  remains  intact 
through  the  reactions  of  the  functional  group. 
We  have  mentioned  earlier  that  when  ethanol 
reacts  with  hydrogen  bromide,  ethyl  bromide  is 
formed.  Similar  treatment  of  ethanol  with  hy- 
drogen chloride  or  hydrogen  iodide  gives  us  the 
corresponding  ethyl  halides: 

CH3CH>OH  +  HBr  — >-  CH3CH2Br  +  H20    (2) 


CH3CH2OH  +  HC1 
CH3CH2OH  +  HI 


CH3GH2CI  +  H20    (3) 
CH3CH2I     +  H20    (4) 


We  say  that  the  hydroxyl  group  has  been  dis- 
placed, and  the  halogen  atom  substituted  for  it. 
You  can  see  that  the  group  CH3CH2—  has  re- 
mained intact  in  all  of  these  reactions.  Indeed, 
this  group  has  appeared  in  most  of  our  discussion 
so  far,  sometimes  attached  to  oxygen  (as  in 
ethanol  and  sodium  ethoxide),  sometimes  at- 
tached to  other  atoms  (as  in  the  ethyl  halides). 
You  will  recall  that  earlier  we  became  acquainted 


with  ethane,  C2H6.  Looking  at  the  structural 
formula  of  ethane,  you  see  that  it  is  simply  the 
CH3CH2 —  group  attached  to  hydrogen: 

H    H 


H— C— C— H     or    CH3CH2 

I       I 
H    H 


H 


This  group,  CH3CH2 —  (also  written  C2H5 — ),  is 
called  the  ethyl  group. 

Because  ethyl  bromide,  ethyl  alcohol  (etha- 
nol), etc.,  can  be  thought  of  as  being  derived 
from  ethane  by  the  substitution  of  one  of  its 
hydrogens  by  — Br,  — OH,  etc.,  we  speak  of  these 
as  derivatives  of  ethane,  and  we  say  that  ethane 
is  the  parent  hydrocarbon  for  a  series  of  related 
compounds. 

The  name  ethyl  is  derived  from  the  name  of 
the  parent  hydrocarbon,  ethane.  In  the  same 
way  the  name  of  the  methyl  group  (CH3 — )  is 
derived  from  that  of  methane,  CH4 ;  the  name  of 
the  propyl  group  (CH3CH2CH2 — )  is  derived 
from  propane,  CH3CH2CH3;  etc. 

It  is  important  to  recognize  that  these  groups 
are  not  substances  that  can  be  isolated  and 
bottled.  They  are  simply  parts  of  molecules  that 
remain  intact  in  composition  and  structure  dur- 
ing reactions.  We  find  this  way  of  classifying 
organic  groups  a  useful  and  convenient  one,  but 
we  must  keep  in  mind  that  in  the  reactions  we 
have  described,  the  ethyl  group  is  not  actually 
formed  as  a  distinct  substance.  Table  18-1  gives 
more  examples  of  group  names  (see  p.  338). 


18-3    SOME  CHEMISTRY  OF  ORGANIC  COMPOUNDS 


18-3.1    Chemical  Behavior  off  Ethyl  and 
Methyl  Bromide 

We  can  use  these  bromine  compounds  to  illus- 
trate one  kind  of  organic  reaction.  Ethyl  bromide 
is  not  particularly  reactive  but  it  does  react  with 
bases  such  as  NaOH  or  NH3.  If  we  mix  ethyl 
bromide  and  aqueous  sodium  hydroxide  solu- 
tion and  heat  the  mixture  for  an  hour  or  so,  we 
find  that  sodium  bromide  and  ethanol  are 
formed. 


C2H5Br  +  OH(aq)  — >■  C2H5OH  +  Br-(aq)  (5) 
This  reaction  may  seem  similar  to  the  reaction 
between  aqueous  HBr  and  NaOH  but  there  are 
two  important  differences.  The  ethyl  bromide 
reaction  is  very  slow  (about  one  hour  is  needed 
for  the  reaction)  and  it  occurs  between  a  covalent 
molecule  (C2H5Br)  and  an  ion  (OH").  In  con- 
trast, the  reaction  between  HBr  and  NaOH  in 
water  occurs  in  a  fraction  of  a  second  and  it 
involves  ions  only,  as  shown  in  reaction  (<5). 


SEC.     18-3    I    SOME    CHEMISTRY    OF    ORGANIC    COMPOUNDS 


331 


H+(aq)  +  OH(aq)  — ►-  H20  (6) 

Let's  describe  the  course  of  reaction  (5)  in 
terms  of  a  model.  We  will  use  methyl  bromide 
to  make  the  description  simpler  but  the  reaction 
of  ethyl  bromide  is  of  the  same  type. 

CH3Br  +  OH"  — -*-  CH3OH  +  Br"         (7) 

First  of  all,  let  us  recount  a  few  of  the  experi- 
mental facts. 

(1)  Methyl  bromide  is  a  compound  in  which  the 
chemical  bonds  are  predominantly  covalent. 
An  aqueous  solution  of  methyl  bromide  does 
not  conduct  electricity,  hence  it  does  not 
form  ions  (such  as  CH3+  and  Br~  ions)  in 
aqueous  solutions. 

(2)  The  reaction  takes  a  measurable  time  for 
completion. 

(3)  Experiments  show  that  the  rate  of  the  reac- 
tion is  increased  by  increasing  the  concentra- 
tion of  OH-  and  also  by  raising  the  tem- 
perature. 


OH-(<tq) 


CH5Br- 


A..  Re  octants  approaching 


These  observations  remind  us  of  Chapter  8, 
in  which  we  considered  the  factors  that  deter- 
mine the  rate  of  a  chemical  reaction.  Of  course, 
the  same  ideas  apply  here.  We  can  draw  qualita- 
tive information  about  the  mechanism  of  the 
reaction  by  applying  the  collision  theory.  With 
quantitative  study  of  the  effects  of  temperature 
and  concentration  on  the  rate,  we  should  be 
able  to  construct  potential  energy  diagrams  like 
those  shown  in  Figure  8-6  (p.  134). 

Figure  18-4  illustrates  the  mechanism  chemists 
have  deduced  for  this  reaction.  This  picture 
shows:  (A)  the  approach  of  the  hydroxide  ion, 
(B)  the  atomic  arrangement  thought  to  be  the 
activated  complex,  and  (Q  the  final  products. 
In  the  activated  complex  the  O — C  bond  is  be- 
ginning to  form  and  the  C— Br  bond  is  beginning 

Fig.  18-4.  The  mechanism  and  potential  energy  dia- 
grams for  the  reaction 
CHaBr  +  OH(aq)  — »»  CH3OH  +  Br-(aq) 


Reaction    coordinate 


Reaction,    coordinate 
S,    A  possible  form,  of  the   activated  complex   (Notice  the  unstable  positions  of  the  hydrogen   atoms) 


HOCH3 


Sr-(a<J) 


Reaction    coordinate 


C.    Products  separating  (Notice  the  new,   stable  positions  of  the  hydrogen    atoms.) 


332 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


to  break.  The  potential  energy  curves  for  the 
reaction  are  shown  alongside  the  molecular 
models.  The  slow  rate  shows  that  activation 
energy  is  needed.  One  of  the  reasons  why  activa- 
tion energy  must  be  supplied  is  that  in  the  ac- 
tivated complex  the  bond  angles  have  been 
distorted  from  their  favorable  (stable)  configura- 
tions and  forced  into  an  unstable  condition. 


18-3.2    Oxidation  of  Organic  Compounds 

By  far  the  majority  of  the  million  or  so  known 
compounds  of  carbon  also  contain  hydrogen  and 
oxygen.  There  are  several  important  types  of 
oxygen-containing  organic  compounds  and  they 
can  be  studied  as  an  oxidation  series.  For  in- 
stance, the  compound  methanol,  CH3OH,  is  very 
closely  related  to  methane,  as  their  structural 
formulas  show.  Methanol  can  be  regarded  as  the 
first  step  in  the  complete  oxidation  of  methane 
to  carbon  dioxide  and  water. 


Fig.  18-5.  Structural  formulas  of  methane  and 
methanol. 


// 


H C 


\ 


H 


formaldehyde,  HCHO 


H^ 


H- 


metha-nol,    CH3OH 


/ 
c 

acetaldehyde,    CH3CH0 


Fig.  18-6.  Structural  formulas  of  formaldehyde  and 
acetaldehyde. 


ALDEHYDES 

Methanol  (and  other  alcohols)  react  with  com- 
mon inorganic  oxidizing  agents  such  as  potas- 
sium dichromate,  K2Cr207.  When  an  acidic, 
aqueous  solution  of  potassium  dichromate  reacts 
with  methanol,  the  solution  turns  from  bright 
orange  to  muddy  green,  owing  to  the  production 
of  the  green  chromic  ion,  Cr+3.  The  solution  then 
has  a  strong  odor  easily  identified  as  that  of 
formaldehyde,  CH20.  This  formula  represents 
the  structure  at  the  top  in  Figure  18-6.  Notice 
that  the  bond  between  carbon  and  oxygen  is  a 
double  bond  (see  Section  16.5),  and  that  all  the 
atoms  lie  in  the  same  plane. 


SEC.     18-3    I    SOME    CHEMISTRY    OF    ORGANIC    COMPOUNDS 


333 


The  balanced  net  reaction  for  the  formation  of 
formaldehyde  is 

3CH3OH  +  Cr2Of  2(aq)  +  8H +(aq)  — >- 

3CH.O  +  2Cr  "(aq)  +  7H,0     (5) 

Since  the  dichromate  ion  on  the  left  side  of  the 
equation  has  been  reduced  to  chromic  ion,  Cr+3, 
on  the  right  side,  the  conversion  of  methanol  to 
formaldehyde  must  involve  oxidation.  To  show 
more  clearly  that  methanol  has  been  oxidized, 
let  us  balance  this  reaction  by  the  method  of 
half-reactions.  We  have  encountered  the  half- 
reaction  involving  dichromate  and  chromic  ions 
before  (Problem  20b  in  Chapter  12).  It  is 

Cr207-2(aq)  +  \4H+(aq)  +  6e~  — ►- 

2Cr+3(aq)  +  7H20    (9) 

To  balance  the  methanol-formaldehyde  half- 
reaction  we  write,  as  a  start, 

CH3OH  gives  CH20  (10a) 

This  statement  does  not  yet  show  the  fact  that 
hydrogen  atoms  are  conserved  in  the  reaction, 
since  there  is  a  deficiency  of  two  hydrogen  atoms 
on  the  right.  This  can  be  remedied  by  adding 
two  hydrogen  ions, 

CH3OH  gives  CH20  +  2U+(aq)        (10b) 

Now  the  equation  is  chemically  balanced,  but 
not  electrically  balanced.  The  addition  of  two 
electrons  to  the  right-hand  side  completes  the 
balancing  procedure,  and  the  completed  half- 
reaction  is 

CH3OH  — >-  CH20  +  2H+( aq)  +  2e~    (10c) 

This  equation  shows  that  the  methanol  molecule 
has  lost  electrons  and  thus  has  been  oxidized. 
Formaldehyde  is  the  second  member  in  the  oxi- 
dation series  of  methane. 

In  a  similar  manner,  ethanol  can  be  oxidized 
by  the  dichromate  ion  to  form  a  compound 
called  acetaldehyde,  CH3CHO.  The  molecular 
structure  of  acetaldehyde,  which  is  similar  to 
that  of  formaldehyde,  is  shown  at  the  bottom  in 
Figure  18-6.  We  see  that  the  molecule  is  struc- 
turally similar  to  formaldehyde.  The  methyl 
group,  — CH3,  replaces  one  of  the  hydrogens  of 
formaldehyde.  The  balanced  equation  for  the 
formation  of  acetaldehyde  from  ethanol  is 


3CH3CH2OH  +  Cn07-2(aq)  +  m+(aq)  —+ 

3CH3CHO  +  2Cr+3(aq)  +  7H20 

CARBOXYLIC    ACIDS 


(11) 


Another  oxidation  product  can  be  obtained  from 
the  reaction  of  an  acidic  aqueous  solution  of 
potassium  permanganate  with  methanol.  The 
product  has  the  formula  HCOOH,  and  is  called 
formic  acid.  The  structural  formula  of  formic 
acid  is  shown  in  Figure  18-7.  The  structure  of 
formic  acid  is  also  related  to  the  structure  of 
formaldehyde.  If  one  of  the  hydrogen  atoms 
of  formaldehyde  is  replaced  by  an  OH  group,  the 


Fig.  18-7.  Structural    formulas    of   formic   acid   and 
acetic  acid. 


// 


H C 


\ 


,H 


formic  acid,  HCOOH 


H— 


// 


H 


acetic  acid,    CHjCOOH 


334 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


resulting  molecule  is  formic  acid.  The  balanced 
equation  for  its  formation  from  methanol  is 

5CH3OH  +  4Mn04"  (aq)  +  \2H+(aq)  —*- 

5HCOOH  +  4Mn+*(aq)  +  11H20    (12) 

The  half-reaction  involving  methanol  and  formic 
acid  can  be  obtained  by  using  the  three  steps 
outlined  in  our  previous  example: 

CH3OH  gives  HCOOH  (13a) 

Chemical  balance: 
CH3OH  +  H20  gives  HCOOH  +  4H+(  aq)    (13b) 

Charge  balance:  CH3OH  +  H20  — >- 

HCOOH  +  4H+(aq)  +  4e~    (13c) 

From  this  completed  half-reaction  we  see  that 
the  conversion  of  methanol  to  formic  acid  in- 
volves the  loss  of  four  electrons.  Since  the  oxida- 
tion of  methanol  to  formaldehyde  was  only  a 
two-electron  change,  it  is  clear  that  formic  acid 
is  a  more  highly  oxidized  compound  of  carbon 
than  formaldehyde  or  methanol. 


EXERCISE  18-6 

Balance  the  half-reaction  for  the  conversion  of 
formaldehyde,  HCHO,  to  formic  acid,  HCOOH. 


Just  as  methanol  can  be  oxidized  to  formic 
acid  [reaction  (72)],  ethanol  can  be  oxidized  to 
an  acid,  CH3COOH,  called  acetic  acid.  The 
molecular  structure  of  acetic  acid  is  shown  in 
Figure  18-7.  The  atomic  grouping  — COOH  is 
called  the  carboxyl  group  and  acids  containing 
this  group  are  called  carboxylic  acids. 

The  balanced  equation  for  production  of 
acetic  acid  from  ethanol  is 


5CH3CH2OH  +  4Mn04"  (aq)  +  UH+(aq)  — ■ 
5CH3COOH  +  AMn^aq)  +  11H20 


(14) 


Acetic  acid  can  also  be  obtained  by  the  oxidation 
of  acetaldehyde,  CH3CHO: 

5CH3CHO  +  2Mn04-faqJ  +  6H+(aq)  — ►- 

5CH3COOH  +  2Mn+i(aq)  +  3H20    (75) 

The  oxidation  of  acetic  acid  is  difficult  to  accom- 
plish. It  does  not  react  in  solutions  of  K2Cr207 
or  KMn04.  Vigorous  treatment,  such  as  burning, 


causes  its  complete  oxidation  to  carbon  dioxide 
and  water.  Formic  acid  also  can  be  oxidized  to 
carbon  dioxide  and  water  by  combustion  with 
oxygen. 


EXERCISE  18-7 

There  is  a  compound  called  propanol  with  struc- 
tural formula  CH3CH2CH2OH.  If  it  is  oxidized 
carefully,  an  aldehyde  called  propionaldehyde  is 
obtained.  Vigorous  oxidation  gives  an  acid  called 
propionic  acid.  Draw  structural  formulas  like 
those  shown  in  Figures  18-6  and  18-7  for  pro- 
pionaldehyde and  propionic  acid. 

EXERCISE  18-8 

Balance  the  half-reaction  involved  in  the  oxida- 
tion of  ethanol  to  acetic  acid.  Compare  the 
number  of  electrons  released  per  mole  of  ethanol 
with  the  number  per  mole  of  methanol  in  the 
equivalent  reaction  (75c).  How  many  electrons 
would  be  released  per  mole  of  propanol  in  the 
oxidation  to  propionic  acid? 


KETONES 

The  bonding  rules  permit  us  to  draw  two  accept- 
able structural  formulas  for  an  alcohol  contain- 
ing three  carbon  atoms,  CH3CH2CH20H  and 
CH3CHOHCH3.  In  the  first  of  these  isomers 
(the  one  considered  in  Exercises  18-7  and  18-8), 
the  OH  group  is  attached  to  the  end  carbon 
atom.  In  the  second,  the  OH  group  is  attached 
to  the  second  carbon  atom.  They  are  both  called 
propanol  because  they  are  both  derived  from 
CH3CH2CH3,  propane.  They  are  distinguished 
by  numbering  the  carbon  atom  to  which  the 
functional  group,  the  OH,  is  attached.  Thus, 
CH3CH2CH2OH  is  called  1 -propanol  because 
the  OH  is  on  the  first  (the  end)  carbon  atom  in 
the  chain.  The  other  alcohol,  CH3CHOHCH3, 
is  called  2-propanol  because  the  OH  is  on  the 
second  carbon  atom.  The  structures  of  these  two 
alcohols  are  shown  in  Figure  18-8. 

We  have  already  considered  the  oxidation  of 
1 -propanol  in  Exercise  18-7.  The  second  isomer, 


SEC.     18-3    I    SOME    CHEMISTRY    OF    ORGANIC    COMPOUNDS 


335 


l-propanol,   CH3CH2CH2OH 


Z-propanol.    CH3CHOHCH3 

Fig.  18-8.  The    molecular    structures    of    l-propanol 
and  2-propanol. 


2-propanol,  can  also  be  oxidized  and  the  product 
is  called  acetone: 

3CH3CHOHCH3  +  Cr  JO  f2(aq)  +  SH+(aq)  — >- 

3CH3COCH3  +  2Ct+3(aq)  +  7H20    {16) 

Acetone  has  the  structure  shown  in  Figure  18-9. 
Acetone  is  the  simplest  member  of  a  class  of 
compounds  called  ketones.  They  are  quite  similar 
in  structure  to  the  aldehydes,  since  each  contains 
a  carbon  atom  doubly  bonded  to  an  oxygen 
atom.*  They  differ  in  that  the  aldehyde  has  a 

*  The  group     C=0  is  called  the  carbonyl  group. 


hydrogen  atom  attached  to  this  same  carbon 
atom  whereas  the  ketone  does  not.  (Compare 
Figures  18-6  and  18-9.)  Since  this  hydrogen  atom 
is  not  present,  a  ketone  cannot  be  oxidized 
further  to  an  acid. 

Figure  18-10  summarizes  the  successive  oxida- 
tion products  that  can  be  obtained  from  alcohols. 
When  the  hydroxyl  group,  OH,  is  attached  on 
an  end  carbon  atom,  an  aldehyde  and  a  car- 
boxylic  acid  can  be  obtained  through  oxidation. 
When  the  hydroxyl  group  is  on  a  carbon  atom 
attached  to  two  other  carbon  atoms,  oxidation 
gives  a  ketone.  Huge  amounts  of  aldehydes  and 
ketones  are  used  industrially  in  a  variety  of 
chemical  processes.  Furthermore,  these  func- 
tional groups  are  important  in  chemical  syn- 
theses of  medicines,  dyes,  plastics,  and  fabrics. 

18-3.3    The  Functional  Group 

The  reactive  groups  we  have  encountered  thus 
far,  such  as  —Br,  —OH,  — CHO,  — COOH,  are 
called  functional  groups.  They  are  the  parts  from 
which  the  molecules  get  their  characteristic 
chemical  behavior. 

For  example,  it  is  a  general  behavior  of  alco- 
hols to  undergo  a  reaction  in  which  the  — OH 
group  is  displaced  by  a  halogen  atom,  such  as 
— Br  [as  in  reaction  (77)]. 

Fig.  18-9.  The  molecular  structure  of  acetone,   the 

simplest  ketone. 


acetone,    CH3C0CH3 


336 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


1-  alcohol        ».         aldehyde       ^.        car-boxylic   acid    — »~    carbon  dioxide     -h   tva-ter- 


-2e- 


:> 


^> 


+ 


\\ 


c  — 


2  ~ alcohol 


C 

II 
O 

ketone 


o=c=o   +  H\ 


carbon,  dioxide  ~h  ->va~ter 


It  is  a  general  characteristic  of  aldehydes  to  be 
oxidizable  to  acids,  as  in  reaction  (79). 

These  common  types  of  behavior  are  shown 
by  using  the  general  symbol,  R — ,  to  stand  for 
the  part  of  the  molecule  which  does  not  change, 
and  writing  a  reaction  in  such  a  way  as  to  focus 
attention  on  the  functional  group.  For  example: 

RCH2OH  +  HBr  — ■»-  RCH2Br  +  H20    (77) 

RCH2Br  +  OU(aq)  — *■ 

RCH2OH  +  Br-(aq)    (75) 

3RCHO  +  Cr,07-2(aq)  +  SH+(aq)  — >- 

3RCOOH  +  2Cr+*(aq)  +  4H20    (79) 

The  symbol  R —  represents  any  alkyl  group 
(such  as  CH3 — ,  C2H6 — )  in  these  formulas. 

18-3.4    Amines 

Alcohols  can  be  related  to  water  by  imagining 
that  an  alkyl  group  (such  as  — CH3)  has  been 
substituted  for  one  of  the  two  hydrogen  atoms 


Fig.  18-10.  Summary:  The  successive  steps  in  the  oxi- 
dation of  an  alcohol. 


of  water.  In  the  same  way,  amines  are  related  to 
ammonia: 

H  H  H 


N— H 


N— R 


H 


H 


H 


an  amine 


N— CH2CH3 


ethylamine 


Amines  can  be  prepared  by  direct  reaction  of 
ammonia  with  an  alkyl  halide,  such  as  CH3Br  or 
CH3CH2I.  Iodides  react  fastest  and  an  excess  of 
ammonia  is  often  used  to  help  control  formation 
of  undesired  alternate  products: 

CH3CH2NH2  +  NHJ    (20) 

ethylamine 

RNH2  +  NHJ  (27) 


CH3CH2I  +  2NH3 

ethyl  iodide 

R— I  +  2NH3 


Equations  (20)  and  (27)  represent  a  net  change 
that  occurs  when  an  excess  of  ammonia  reacts 


SEC.     18-3    I    SOME    CHEMISTRY    OF    ORGANIC    COMPOUNDS 


337 


with  an  alkyl  iodide.  The  actual  reaction  goes  by 
two  successive  steps.  The  first  step  is  analogous 
to  the  attack  of  the  hydroxide  ion  on  an  alkyl 
halide  (see  Figure  18-4): 

NH3  +  RI  — »-  RNH3+  (aq)  +  l(aq)      (22) 

The  second  step  is  a  proton  transfer  reaction  (see 
Section  11-3.3): 
NH3  +  RNH3+  (aq)  — ►-  RNH2  +  NH +  (aqj    (23) 

18-3.5    Acid  Derivatives:  Esters  and  Amides 

We  see  from  reactions  (75)  and  (79)  that  oxida- 
tion of  an  aldehyde  gives  an  organic  acid.  All 
of  these  acids  contain  the  functional  group 
— COOH,  the  carboxyl  group.  The  bonding  in 
this  group  is  as  follows: 


O 


— C 


• 


O— H 


The  carboxyl  group  readily  releases  a  proton,  so 
it  is  an  acid.  For  example,  acetic  acid  dissolves 
in  water  and  the  solution  is  conducting,  it  turns 
blue  litmus  red,  it  is  sour,  and  it  shows  all  the 
other  properties  of  an  acid.  The  reaction 

CH3COOH  +  H20  — >- 

CH3COO(aq)  +  H30+(aq)    (24) 

has  an  equilibrium  constant  of  1.8  X  10~5. 

In  addition  to  this  acidic  behavior,  an  im- 
portant characteristic  of  carboxylic  acids  is  that 
the  entire  OH  group  can  be  replaced  by  other 
groups.  The  resulting  compounds  are  called  acid 
derivatives.  We  will  consider  only  two  types  of 
acid  derivatives,  esters  and  amides. 

ESTERS 

Compounds  in  which  the  — OH  of  an  acid  is 
transformed  into  —OR  (such  as  — OCH3)  are 
called  esters.  They  can  be  prepared  by  the  direct 
reaction  between  an  alcohol  and  the  acid.  For 
example, 


CH,OH  +  CH3C 


/ 


OH 


methyl     ,        acetic 
alcohol  acid 


The  method  of  naming  is  indicated  by  the  bold 
face  parts  of  the  names  of  the  reactants. 

EXERCISE  18-9 

Write  equations  for  the  reaction  of  (a)  ethanol 
and  formic  acid;  (b)  propanol  and  propionic 
acid;  (c)  methanol  and  formic  acid.  Name  the 
esters  produced. 

When  equilibrium  is  reached  in  reaction  (25), 
appreciable  concentrations  of  all  of  the  reactants 
may  be  present.  If  methyl  acetate  (the  product 
on  the  right)  alone  is  dissolved  in  water,  it  will 
react  with  water  slowly  to  give  acetic  acid  and 
methanol  until  equilibrium  is  attained: 


O 


CH3C 


/ 


+  H20 


O— CH3 


O 


CH3OH  +  CH3C 


./ 


(26) 


OH 


Of  course,  the  usual  equilibrium  considera- 
tions apply.  For  example,  if  we  add  the  substance 
methanol,  equilibrium  conditions  will  shift,  con- 
suming the  added  reagent  (methanol)  and  acetic 
acid  to  produce  more  methyl  acetate  and  water, 
in  accord  with  Le  Chatelier's  Principle.  Thus  a 
large  excess  of  methanol  causes  most  of  the 
acetic  acid  to  be  converted  to  methyl  acetate. 


EXERCISE  18-10 

Write  the  equilibrium  expression  relating  the 
concentrations  of  reactants  and  products  in  re- 
action (26).  Notice  that  the  concentration  of 
water  must  be  included  because  it  is  not  neces- 
sarily large  enough  to  be  considered  constant. 


O 


CH3C 


/ 


+  H,0 


(25) 


O—  CH3 
methyl 
acetate 


+  water 


338 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  f  CHAP.  18 


Reaction  (25)  between  methanol  and  acetic  acid 
is  slow,  but  it  can  be  speeded  up  greatly  if  a 
catalyst  is  added.  For  example,  addition  of  a 
strong  acid  such  as  hydrochloric  acid  or  sulfuric 
acid  will  speed  up  the  reaction  by  catalysis.  As 
mentioned  in  Section  9-1.4,  the  catalyst  does  not 
alter  the  equilibrium  state  (that  is,  the  concentra- 
tions of  the  reactants  at  equilibrium),  but  only 
permits  equilibrium  to  be  attained  more  rapidly. 


EXERCISE  18-11 

A  strong  acid  such  as  hydrochloric  acid  or  sul- 
furic acid  will  catalyze  reaction  (25).  Explain 
why  this  implies  that  these  acids  will  catalyze 
reaction  (26)  as  well.  (Consult  Section  8-2.3.) 


Esters  are  important  substances.  The  esters  of 
the  low  molecular  weight  acids  and  alcohols  have 
fragrant,  fruit-like  odors  and  are  used  in  per- 
fumes and  artificial  flavorings.  Esters  are  useful 
solvents;  this  is  the  reason  they  are  commonly 
found  in  "model  airplane  dope"  and  fingernail 
polish  remover. 


AMIDES 

A  compound  in  which  the  — OH  group  of  an 
acid  is  replaced  by  — NH2  is  called  an  amide. 
When  the  —OH  is  replaced  by  — NHR,  the 
product  is  called  a  nitrogen-substituted  amide  or, 
abbreviated,  an  N-substituted  amide.  One  way 
to  make  amides  is  to  react  ammonia  (or  an 
amine)  with  an  ester: 


CH3C 


/ 


O 


+  NH3 


O— CH3 


methyl 

o 

acetate 

/ 

CH3C             +  CHjOH    (27) 

\ 

ace  t  amide       NHj 

o 

// 

CH3C 

+  H2N— CH2CH3  — >- 

\ 

ethylatnine 

O— (  H 

methyl 

o 

acetate 

/ 

CH3C                          +  CH3OH    (28) 

\ 

N— CH2CHS 

JV-ethyl              1 
acetamide       n. 

Table  18-1.     regularities  in  names  of  alkanes,  alcohols,  and  amines 


NUMBER   OF 
CARBON   ATOMS 


ALCOHOLS 


AMINES 


1 

CH< 

CH,OH 

CH,NH2 

methane 

methanol 
methyl  alcohol 

methylamine 

2 

CH.CH, 

CH,CH2OH 

CH,CH2NH, 

ethane 

ethanol 
ethyl   alcohol 

ethylamine 

3 

CHjCHjCHi 

CH,CH2CH2OH 

CH,CH2CH2NH, 

propane 

1-propanol 
propyl  alcohol 

1 -propylamine 

4 

CH3CHSCH2CH1 

CH,CH2CH2CH2OH 

CH3CH2CH2CH2NH2 

butane 

1-butanol 
butyl  alcohol 

1-butylamine 

8 

chkch^ch, 

CH,(CH2),CH2OH 

CH,(CH2)6CH2NHf 

octane 

1-octanol 
octvl  alcohol 

1-octylamine 

SEC.     18-4    I    NOMENCLATURE 


339 


Note  the  similarity  of  the  two  reactions.  Amides 
are  of  special  importance  because  the  amide 


grouping 


O 


— C 


/ 


\ 


NH— 


is  the  basic  structural  element  in  the  long-chain 
molecules  that  make  up  proteins  and  enzymes 
in  living  matter.  Hydrogen  bonding  between  two 
amide  groups  helps  determine  the  protein  struc- 
ture, a  topic  that  will  be  dealt  with  later,  in 
Chapter  24. 


18-4    NOMENCLATURE 

The  names  of  organic  compounds  have  some 
system.  Each  functional  group  defines  a  family 
(for  example,  alcohols,  amines)  and  a  specific 
modifier  is  added  to  identify  a  particular  example 
(for  example,  ethyl  alcohol,  ethyl  amine).  As  an 
alternate  naming  system,  the  family  may  be 
named  by  a  general  identifying  ending  (for  ex- 
ample, alcohol  names  end  in  -ol)  and  a  particular 
example  is  indicated  by  an  appropriate  stem 
(ethyl  alcohol  would  be  ethanol).  These  naming 
systems  are  illustrated  in  Tables  18-1  and  18-11. 
Scrutiny  of  these  tables  reveals  that  the  key 


to  the  system  is  the  name  of  the  alkane  which  is 
modified  in  a  systematic  way  to  get  the  names 
that  carry  over  into  the  acid  derivatives.  Starting 
at  pentane  (C6Hi2)  and  hexane  (C6Hu)  the  alkane 
names  are  themselves  quite  regular.  They  are 
derived  from  Greek  words  for  the  number  of 
carbon  atoms. 

The  compounds  with  more  complicated  shapes 
and  more  than  one  functional  group  are  de- 
scribed by  a  straightforward  numbering  system 
that  you  will  learn  in  later  chemistry  courses. 
Other  functional  groups  will  be  studied  then  too. 


Table  18-11.     regularities  in  names  of  acids,  amides,  and  esters 


NUMBER  OF 

ESTERS  (ACID 

CARBON  ATOMS                           ACIDS 

AMIDES 

WITH  METHANOL) 

1 

HCOOH 

formic  acid 

HCONH2 

formamide 

HCOOCH, 

methyl  formate 

2 

CH.COOH 

acetic  acid 

CH,CONH, 

acetamide 

CH,COOCH, 

methyl  acetate 

3 

CH,CH2COOH 

propionic  acid 

CH,CH2CONH2 

propionamide 

CH,CH2COOCH, 

methyl  propionate 

4 

CHjCH^CHjCOOH 

butyric  acid 

CH,CH,CH2CONH2 

butyramide 

CH,CH2CH2COOCH, 
methyl  butyrate 

8 

CH,(CH,),COOH 

octanoic  acid 
caprylic  acid 

CH^CH^CONH, 

octanamide 
caprylamide 

CH,(CH2),COOCH, 

methyl  octanoate 
methyl  caprylate 

340 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


18-5     HYDROCARBONS 

Compounds  that  contain  only  hydrogen  and  carbon 
are  called  hydrocarbons.  The  hydrocarbons 
that  have  only  single  bonds  all  have  similar 
chemistry  and  they  are  called,  as  a  family,  the 
saturated  hydrocarbons.  If  there  are  carbon- 
carbon  double  bonds,  the  reactivity  is  much 
enhanced.  Hence  hydrocarbons  containing  one 
or  more  double  bonds  are  named  as  a  distinct 
family,  unsaturated  hydrocarbons.  Both  saturated 
and  unsaturated  hydrocarbons  can  occur  in 
chain-like  structures  or  in  cyclic  structures.  Each 
of  these  families  will  be  considered. 


18-5.1    Saturated  Hydrocarbons 

We  have  already  remarked  that  ethane  is  a  mem- 
ber of  a  family  of  compounds  called  the  saturated 
hydrocarbons.  This  term  identifies  compounds 
that  contain  only  carbon  and  hydrogen  in  which 
all  bonds  to  carbon  are  single  bonds  formed  with 
hydrogen  or  other  carbon  atoms.  They  occur  in 
chains,  branched  chains,  and  cyclic  structures. 


Fig.  18-11.  Structural  formulas  for  some  five-carbon 
saturated  hydrocarbons. 


n  -pentane 


H    H 

\/ 
C 

/\ 
H—C  C 

r/\  A 


H     H 

\/ 

C 


/  \ 


K 


C—K 


H   H 


H 


w        \/ 

i 

c          c- 

-H 

/  \  / 

H- 

-c         c 

HX\           /V 

H  ^c 

H  A 

a  k 

ieo  -pe.rtta.viG 


H    H 

\/ 


\  I 


9—h 


K' 


\ 


K 


H 


H 


cycLope.yita.YLe 


SEC.     18-5    I    HYDROCARBONS 


341 


The  chain  and  branched  chain  saturated  hydro- 
carbons make  up  a  family  called  the  alkanes. 
Some  saturated  hydrocarbons  with  five  carbon 
atoms  are  shown  in  Figure  18-11.  The  first  ex- 
ample, containing  no  branches,  is  called  normal- 
pentane  or,  briefly,  /i-pentane.  The  second  ex- 
ample has  a  single  branch  at  the  end  of  the  chain. 
Such  a  structural  type  is  commonly  identified  by 
the  prefix  "iso-".  Hence  this  isomer  is  called 
/50-pentane.  The  third  example  in  Figure  18-11 
also  contains  five  carbon  atoms  but  it  contains 
the  distinctive  feature  of  a  cyclic  carbon  struc- 
ture. Such  a  compound  is  identified  by  the  prefix 
"cyclo"  in  its  name — in  the  case  shown,  cyclo- 
pentane. 


normal  conditions — their  boiling  points  are 
shown  in  Table  18-111.  The  composition  of  gaso- 
line is  mainly  highly  branched  alkanes  with  from 
six  to  ten  carbon  atoms.  Paraffin  waxes  are  usu- 
ally alkanes  with  from  twenty  to  thirty-five 
carbon  atoms. 

The  saturated  hydrocarbons  are  relatively  in- 
ert except  at  high  temperatures.  For  example, 
sodium  metal  is  usually  stored  immersed  in  an 
alkane  such  as  kerosene  (8  to  14  carbon  atoms) 
to  protect  it  from  reaction  with  water  or  oxygen. 
Combustion  is  almost  the  only  important  chemi- 
cal reaction  of  the  alkanes.  That  reaction,  how- 
ever, makes  the  hydrocarbons  one  of  the  most 
important  energy  sources  of  our  modern  tech- 
nology. 


EXERCISE  18-12 

What  are  the  empirical  formulas  of  the  three 
compounds  shown  in  Figure  18-11?  The  molecu- 
lar formulas?  Which  are  structural  isomers? 

EXERCISE  18-13 

There  is  one  more  alkane  with  molecular  formula 
C6Hi2,  called  neopentane.  Draw  its  structural 
formula. 


EXERCISE  18-14 

Using  the  data  given  in  the  last  column  of  Table 
18-111,  plot  the  heat  released  per  carbon  atom 
against  the  number  of  carbon  atoms  for  the 
normal  alkanes.  Consider  the  significance  of  this 
plot  in  terms  of  the  molecular  structures  of  these 
compounds. 


The  alkanes  are  the  principal  compounds  pres- 
ent in  natural  gas  and  in  petroleum.  The  low 
molecular  weight  compounds  are  gases  under 


In  a  sense,  the  absence  of  reactivity  of  satu- 
rated hydrocarbons,  whether  cyclic  or  not,  is  a 
crucial  aspect  of  their  chemistry.  This  inertness 
accounts   for   the   fact   that  the  chemistry   of 


Table  18-111.     some  properties  of  saturated  hydrocarbons 


SATURATED 
HYDROCARBON 


MOLECULAR 
FORMULA 


MELTING 
POINT 


BOILING 
POINT 


HEAT  OF 
COMBUSTION  OF 

gas  (kcal/mole) 


methane 

CH< 

-182.5°C 

-161.5°C 

-212.8 

ethane 

CHjCH, 

-183.3°C 

-88.6°C 

-372.8 

propane 

CH3CH2CH} 

-187.7°C 

-42.  rc 

-530.6 

n-butane 

CH3CH2CH2CH, 

-138.4°C 

-0.5°C 

-687.7 

isobutane 

CH,— CH— CH3 
CH, 

-159.6°C 

-11.7°C 

-685.7 

n-hexane 

CtHu 

-95.3°C 

68.7°C 

-1002.6 

cyclohexane 

C«Hi2 

+6.6°C 

80.7°C 

-944.8 

H-octane 

CsHu 

-56.8°C 

125.7°C 

-1317.5 

w-cctadecane 

CisHj8 

+28.2°C 

316.  rc 

-2891.9 

342 


THB  CHBMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


organic  compounds  is  mainly  concerned  with  the 
functional  groups.  The  functional  groups  are 
usually  so  much  more  reactive  than  the  carbon 
"skeleton"  that  it  can  be  assumed  that  the  skele- 
ton will  remain  intact  and  unchanged  through 
the  reaction. 


18-5.2    Unsaturated  Hydrocarbons 

Unsaturated  compounds  are  those  organic  com- 
pounds in  which  less  than  four  other  atoms  are 
attached  to  one  or  more  of  the  carbon  atoms. 
Ethylene,  C2H4,  is  an  unsaturated  compound. 
Because  ethylene  involves  only  carbon  and  hy- 
drogen, it  is  an  unsaturated  hydrocarbon.  Pro- 
pylene, the  next  more  complicated  unsaturated 
hydrocarbon,  has  the  molecular  formula  C3H6. 


The  structural  formulas  of  ethylene  and  propyl- 
ene are  shown  in  Figure  18-12.  Cyclic  hydrocar- 
bons also  can  involve  double  bonds.  The 
structural  formula  of  a  cyclic  unsaturated  hydro- 
carbon is  shown  also  in  Figure  18-12. 

Unsaturated  hydrocarbons  are  quite  reactive 
— in  contrast  to  the  relatively  inert  saturated 
hydrocarbons.  This  reactivity  is  associated  with 
the  double  bond.  In  the  most  characteristic  re- 
action, called  "addition,"  one  of  the  bonds  of 
the  double  bond  opens  and  a  new  atom  becomes 
bonded  to  each  of  the  carbon  atoms.  Some  of 
the  reagents  that  will  add  to  the  double  bond  are 


Fig.  18-12.  Structural  formulas  of  some  unsaturated 
hydrocarbons. 


H 


/ 


H 


H 


\ 


H 


H  H 

\        / 

C  =  C 


H C 

\ 
H 


A 


H 


propylene 


\7 


\     / 


H 


H 


\ 


H 


SEC.     18-5    I    HYDROCARBONS 


343 


H2,  Br2,  HC1,  and  H20.  Examples  are  shown 
below  for  ethylene. 


H  H 

\  / 

C=C 

/  \ 

H  H 

H  H 

\  / 

C=C 

/         \ 
H  H 


+  H2 


+  Br2 


H 

V 

/ 

H 

H 
\ 

/ 
H 


\  / 

C=C         +  HC1 

/  \ 


\  / 

/     \ 


H 
H 

/ 
\ 

H 


+  H20 


H 

H 

\          / 
H— C— C— H 

/ 
H 

\ 
H 

H 

Br 

\ 

H— C- 

/ 
-C— H 

/ 
Br 

\ 
H 

H 

H 

\ 

H— C- 

/ 

-C— H 

/ 
H 

\ 
CI 

H 

H 

\ 
H— C- 

/ 

-C— H 

/ 
H 

\ 
OH 

(29a) 


(29b) 


(29c) 


(29d) 


Oxidizing  agents  also  attack  the  double  bond. 
When  a  reaction  between  an  unsaturated  com- 
pound and  the  permanganate  ion  occurs,  the 
violet  color  of  permanganate  fades.  This  reac- 
tion, as  well  as  reaction  (29b)  in  which  a  color 
change  also  occurs,  is  used  as  a  qualitative  test 
for  the  presence  of  double  bonds  in  compounds. 

18-5.3    Benzene  and  Its  Derivatives 

There  is  another  important  class  of  cyclic  com- 
pounds that  is  still  different  from  the  two  classes 
just  described.  The  simplest  example  is  the  com- 
pound benzene,  a  cyclic  compound  with  six  car- 
bon atoms  in  the  ring  and  formula  C6H6.  The 
benzene  ring  is  found  to  be  planar  with  120° 
angles  between  each  pair  of  bonds  formed  by  a 
given  carbon  atom.  Thus,  experiment  tells  us  the 
molecule  is  a  regular  hexagon  with  the  following 
atomic  arrangement: 

H 

I 
C 

/    \ 
H— C  C— H 

I 
H— C  C— H 

\    / 
C 

I 
H 


H 

H 

H 

| 

H 

H 

| 

H 

\ 

A* 

( 

or 

X( 

S** 

f 

A 

v* 

\ 

/ 

v 

\ 

H 

1 

H 

H 

H 

1 

H 

H 

A  difficulty  arises  when  we  attempt  to  represent 
the  bonding  in  benzene  and  its  derivatives.  We 
might  write 


(30) 


Both  of  these  structures  satisfy  the  formal  va- 
lence rules  for  carbon,  but  each  has  a  serious 
fault.  Each  structure  shows  three  of  the  carbon- 
carbon  bonds  as  double  bonds,  and  three  are 
shown  as  single  bonds.  There  is  a  wealth  of 
experimental  evidence  to  indicate  that  this  is  not 
true.  Any  one  of  the  six  carbon-carbon  bonds  in 
benzene  is  the  same  as  any  other.  Apparently  the 
fourth  bond  of  each  carbon  atom  is  shared 
equally  with  each  adjacent  carbon.  This  makes 
it  difficult  to  represent  the  bonding  in  benzene 
by  our  usual  line  drawings.  Benzene  seems  to  be 
best  represented  as  the  "superposition"  or  "aver- 
age" of  the  two  structures.  For  simplicity,  chem- 
ists use  either  one  of  the  structures  shown  in 
(30)  usually  expressed  in  a  shorthand  form  (31) 
omitting  the  hydrogen  atoms: 


or 


(3D 


Still  another  shorthand  symbol  sometimes  used 
is 

(32) 

Whichever  symbol  is  used,  (30),  (31),  or  (32),  the 
chemist  always  remembers  that  the  carbon- 
carbon  bonds  are  actually  all  the  same  and  that 
they  have  properties  unlike  either  simple  double 
bonds  or  simple  single  bonds. 

THE    SUBSTITUTION    REACTION 
OF    BENZENE 

Benzene  shows  neither  the  typical  reactivity  nor 
the  usual  addition  reaction  of  ethylene.  Benzene 
does  react  with  bromine,  Br2,  but  in  a  different 
type  of  reaction: 


344 


THE    CHEMISTRY    OF    CARBON    COMPOUNDS    I    CHAP.    18 


H 


H 


H 


+  Br2 


H  H 

H 


H 


H 


Br 


+  HBr    (33) 

'\ 
H 
H 

bromobenzene 

In  this  reaction,  called  bromination,  one  of  the 
hydrogen  atoms  has  been  replaced  by  a  bromine 
atom.  Notice  that  the  double  bond  structure  is 
not  affected— this  is  not  an  addition  reaction. 
Nitric  acid  causes  a  similar  reaction,  called  nitra- 
tion: 


H 


H 


H 

H 

/ 


+  HON02 


H 


H 


NO, 


+  H20    (34) 


Reactions  of  the  type  shown  in  (33)  and  (34)  are 
called  substitution  reactions.  The  substitution  re- 
action is  the  characteristic  reaction  of  benzene 
and  its  derivatives  and  is  the  way  in  which  a 
multitude  of  compounds  are  prepared  by  the 
organic  chemist.  By  this  means  he  is  able  to 
introduce  functional  groups,  which  can  then  be 


modified  in  various  ways.  Benzene  and  its  deriva- 
tives are  commonly  called  aromatic  com- 
pounds. 

MODIFICATION   OF   FUNCTIONAL 
GROUPS   ON   THE   BENZENE   RING 

One  of  the  most  important  derivatives  of  benzene  is 
nitrobenzene.  The  nitro  group  is  — NOj.  Nitrobenzene  is 
important  chiefly  because  it  is  readily  converted  into  an 
aromatic  amine,  aniline,  by  reduction.  One  preparative 
procedure  uses  zinc  as  the  reducing  agent: 

NOj 

3ZnfsJ  +  f'  J  +  6H+(aq)  —*- 

nitrobenzene 

NH, 


3Za»(aq)  +  2H,0  +  J 

aniline 

Aniline  and  other  aromatic  amines  are  valuable  indus- 
trial raw  materials.  They  form  an  important  starting  point 
from  which  many  of  our  dyestuffs,  medicinals,  and  other 
valuable  products  are  prepared.  For  example,  you  have 
used  the  indicator,  methyl  orange,  in  your  laboratory 
experiments.  Methyl  orange  is  an  example  of  an  aniline- 
derived  dye,  although  it  is  used  more  as  an  acid-base 
indicator  than  for  dyeing  fabrics.  The  structure  of  methyl 
orange  is  as  follows: 


CH, 


\        /= 


N 


CH, 


y\^ 


N=N 


_i       i 

methyl  orange 


J 


SO,H 


The  portions  of  the  methyl  orange  molecule  set  off  by 
the  dotted  lines  come  from  aromatic  amines  like  aniline. 
Aniline  is  indeed  the  starting  material  from  which  methyl 
orange  and  related  dyes  ("azo  dyes")  are  made. 

Another  useful  aniline  derivative  is  acetanilide,  which 
is  simply  the  amide  formed  from  aniline  and  acetic  acid : 


CH3C 


S 


O         H 


+         N- 


OH       H 

acetic  acid   +         aniline 


CH,C 


S 


N 


/ 
H 

acetanilide 


+  H,0 


+  water 


(36) 


SEC.     18-5    I    HYDROCARBONS 


345 


Acetanilide  has  been  used  medicinally  as  a  pain-killing      zene,  which  is  chlorinated  as  a  first  step.  Reaction  of 
remedy.  chlorobenzene  with  base  gives  phenol: 


PHENOL    AND    ITS    USES 

Another  important  constituent  of  coal  tar  is  hydroxy- 
benzene,  or  phenol: 

OH 


phenol 


Most  of  our  phenol  is  now  made  industrially  from  ben- 


Cl 
CI    — ►-   |         i  +HC1         (37) 

chlorobenzene 


CI 


.    _..    ,       .     heat 

+  OH(aq)  ►- 

pressure 


OH 


+  a~(aq)     (38) 


Table  18-IV.     structures  and  uses  of  some  benzene  derivatives 

STRUCTURE  NAME  USE 


OH 


/ 


OCH, 


C 

/  \ 
H  O 


CH3CH,0 


vanillin 


flavoring  material 


phenacetin 


I 
N  O 

/  \  • 

H  C 

\ 
CH, 


pain-reliever  (in  headache 
remedies) 


OH 


OH 


hydroquinone  photographic  developer 


NH2 

1 

n 

"Novocaine" 

local  anaesthetic 

Ks 

(procaine) 

C                           QH6 

•  \                    / 

O          OCH2CH2N 

\ 

CH=CH2 

a 

styrene 

monomer  for  preparation 

\s> 

of  polystyrene  plastics 

346 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


Phenol  is  a  germicide  and  disinfectant,  and  was  first 
used  by  Lister  in  1867  as  an  antiseptic  in  medicine.  More 
effective  and  less  toxic  antiseptics  have  since  been  dis- 
covered. 

Perhaps  the  most  widely  known  compound  prepared 
from  phenol  is  aspirin.  If  phenol,  sodium  hydroxide,  and 
carbon  dioxide  are  heated  together  under  pressure,  sali- 
cylic acid  is  formed  (as  the  sodium  salt): 


OH 


+  CO,  +  NaOH 


OH 
— COO~Na+ 


{39) 


0%™+™ 


r-  OH 
I— COOH 

salicylic  acid 


+  Na+fooJ    (40) 


Salicylic  acid  is  quite  useful.  Its  methyl  ester  has  a  sharp, 
characteristic  odor  and  is  called  "oil  of  wintergreen." 


The  acid  itself  (or  the  sodium  salt)  is  a  valuable  drug  in 
the  treatment  of  arthritis.  But  the  most  widely  known 
derivative  of  salicylic  acid  is  aspirin,  which  has  the  fol- 
lowing structure: 

O 

o— c 

y     \ 

CH, 

'\ 

COOH 

aspirin 

You  will  see,  by  examining  this  structure,  that  aspirin  is 
an  ester  of  acetic  acid.  Aspirin  is  mankind's  most  widely 
used  drug.  Somewhat  over  20  million  pounds  of  aspirin 
are  manufactured  each  year  in  the  United  States  alone! 
This  amounts  to  something  like  150  five-grain  tablets  for 
every  person  in  the  country ! 

Table  18-IV  shows  the  structures  of  a  few  simple  ben- 
zene derivatives  that  are  important  commercial  products. 
Study  these  structures  so  that  you  can  see  their  relation- 
ship with  the  simple  compounds  from  which  they  are 
derived. 


18-6    POLYMERS 


Table  18-111,  p.  341,  shows  that  the  melting 
points  of  the  normal  alkanes  tend  to  increase  as 
the  number  of  carbon  atoms  in  the  chain  is 
increased.  Ethane,  C2H6,  is  a  gas  under  normal 
conditions;  octane,  C8Hi8,  is  a  liquid;  octadec- 
ane,  Ci8H38,  is  a  solid.  We  see  that  desired  physi- 
cal properties  can  be  obtained  by  controlling  the 
length  of  the  chain.  Functional  groups  attached 
to  the  chain  provide  additional  variability,  in- 
cluding chemical  reactivity.  In  fact,  by  adjusting 
the  chain  length  and  composition  of  high  mo- 
lecular weight  compounds,  chemists  have  pro- 
duced a  multitude  of  organic  solid  substances 
called  plastics.  These  have  been  tailored  to  meet 
the  needs  of  a  wide  variety  of  uses,  giving  rise 
to  an  enormous  chemical  industry. 

The  key  to  this  chemical  treasure  chest  is  the 
process  by  which  extended  chains  of  atoms  are 
formed.  Inevitably  it  is  necessary  to  begin  with 
relatively  small  chemical  molecules — with  car- 
bon chains  involving  only  a  few  atoms.  These 
small  units,  called  monomers,  must  be  bonded 
together,  time  after  time,  until  the  desired  mo- 
lecular weight  range  is  reached.  Often  the  desired 


properties  are  obtained  only  with  giant  mole- 
cules, each  containing  hundreds  or  even  thou- 
sands of  monomers.  These  giant  molecules  are 
called  polymers  and  the  process  by  which  they 
are  formed  is  called  polymerization. 


18-6.1    Types  of  Polymerization 

Polymerization  involves  the  chemical  combina- 
tion of  a  number  of  identical  or  similar  molecules 
to  form  a  complex  molecule  of  high  molecular 
weight.  The  small  units  may  be  combined  by 
addition  polymerization  or  condensation  poly- 
merization. 

Addition  polymers  are  formed  by  the  reaction 
of  the  monomeric  units  without  the  elimination 
of  atoms.  The  monomer  is  usually  an  unsatu- 
rated organic  compound  such  as  ethylene, 
H2C=CH2,  which  in  the  presence  of  a  suitable 
catalyst  will  undergo  an  addition  reaction  to 
form  a  long  chain  molecule  such  as  polyethylene. 
A  general  equation  for  the  first  stage  of  such  a 
process  is 


SEC.     18-6       POLYMERS 


347 


H 


H 


c=c 


H       H 


+ 

H       H 


o=c 


H 


H 


H 

I 
H— C- 


H    H 


C— C=C 


H    H 


H 


H 


The  same  addition  process  continues  and  the 
final  product  is  the  polymer,  polyethylene: 


H 

I 
H— C- 


I— C=C 


H 


H      \H/  „ 

in  which  n  is  a  very  large  number. 


"Lucite,"  and  "Plexiglas"  result.  It  is  thus  pos- 
sible to  create  molecules  with  custom-built  prop- 
erties for  various  uses  as  plastics  or  fibers. 

Condensation  polymers  are  produced  by  reac- 
tions during  which  some  simple  molecule  (such 
as  water)  is  eliminated  between  functional  groups 
(such  as  alcoholic  OH  or  acidic  COOH  groups). 
In  order  to  form  long  chain  molecules,  two  or 
more  functional  groups  must  be  present  in  each 
of  the  reacting  units.  For  example,  when  ethylene 
glycol,  HOCH2CH2OH,  reacts  with  />araphthalic 


acid  HOOC 


COOH  a  polyester  of 


high  molecular  weight  called  "Dacron"  is  pro- 
duced. The  equation  below  shows  the  first 
stages  of  this  process: 


H- 


H    H        r 

-O— C— C— O-f-H  + 


H    H 


HO-f-C— /     V-C— ' 

!l    \=/ll' 
'  O  O  l. 


,  OH  +  H+> 


H    H 


H    H 


When  one  or  more  of  the  hydrogens  are  re- 
placed by  groups  such  as  fluorine,  F;  chlorine, 
CI;  methyl,  CH3  or  methyl  ester,  COOCH3; 
synthetic  polymers  such  as  "Teflon,"  "Saran," 


Fig.   18-13.   Molecular  structures  of  a-amino  acids. 

group  :    a -ammo   acid  group 


an     tx  -  amino  acid, 
(general  formula) 


glycine 


\ 


/ 


C CHZ CH2 


glutawiic   acid 


18-6.2    "Nylon,"  a  Polymeric  Amide 

"Nylon,"  the  material  widely  used  in  plastics 
and  fabrics,  is  a  condensation  polymer.  It  con- 
sists of  molecules  of  extremely  high  molecular 
weight  and  it  is  made  up  of  small  units  joined 
one  to  another  in  a  long  chain  of  atoms.  The 
reaction  by  which  the  units  become  bonded  to- 
gether is  a  conventional  amide  formation.  There 
is,  however,  the  additional  requirement  that  the 
reaction  must  take  place  time  after  time  to  form 
an  extended,  repeating  chain.  To  accomplish 
this,  we  select  reactants  with  two  functional 
groups.  Thus  polyamides  can  be  made  from  one 
compound  with  two  acid  groups, 

HOOC— CH2— CH2— CHr- CH*— COOH 

adipic  acid 

and  another  with  two  amine  groups, 
H2N— CH2— CH2— CH2— CHz— CH2— CH2— NH2 

1 , 6-diami  noh  e  xane 


348 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  1 


These  molecules  can  react  repeatedly,  each  time 
removing  water,  and  forming  amide  linkages  at 
both  ends: 


O  O 

C-(CH2)4-C 


\ 


C-(CH2)4-C 


H— O 


N-(CH2)6-N 


N— (CH2)6— N 

I  I 

H  H 


O  O 

\  S 

C— (CH2)4— C  H 

/  \  / 

N— (CH2)6— N 


H 


H 


Other  polyamides  can  be  made  from  different 
acids  and  other  amines,  giving  a  variety  of  prop- 
erties suited  to  a  variety  of  uses. 


18-6.3    Protein,  Another  Polymeric  Amide 

A  most  important  class  of  polyamides  is  that  of 
the  proteins,  the  essential  structures  of  all  living 
matter.  In  addition,  they  are  a  necessary  part  in 
the  diet  of  man  because  they  are  the  source  of 
the  "monomeric"  units,  the  amino  acids,  from 
which  living  protein  materials  are  made. 

Proteins  are  polyamides  formed  by  the  poly- 
merization, through  amide  linkages,  of  a-amino 
acids.  Three  of  the  25-30  important  natural 
a-amino  acids  are  shown  in  Figure  18-13.  Each 
acid  has  an  amine  group,  — NH2,  attached  to 
the  a-carbon,  the  carbon  atom  immediately  ad- 
jacent to  the  carboxylic  acid  group. 

The  protein  molecule  may  involve  hundreds 
of  such  amino  acid  molecules  connected  through 
the  amide  linkages.  A  portion  of  this  chain  might 
be  represented  as  shown  in  Figure  18-14. 


Fig.  18-14.  The  structure  of  protein  showing  the  am- 
ide chain. 


An  amide  is  decomposed  by  aqueous  acids  to 
an  acid  and  an  amine  (produced  as  the  ammo- 
nium salt): 


O 


CH3C 


y 


+  H30+(aq) 


NHCH3 


A'-methyl 

acetamide 


o 


CH3C 


• 


\ 


+  CH3NH3+(aqJ     (41) 


OH 


acetic 
acid 


methyl 

ammonium 

ioD 


In  this  same  type  of  reaction,  a  protein  can  be 
broken  down  into  its  constituent  amino  acids. 

It  is  in  this  way  that  most  of  what  we  call  the 
"natural"  amino  acids  have  been  discovered. 
Proteins  from  many  sources — egg  yolk,  milk, 
animal  tissues,  plant  seeds,  gelatin,  etc. — have 
been  studied  to  learn  what  amino  acids  compose 
them.  In  this  way  about  thirty  of  the  "natural" 
amino  acids  have  been  identified. 

When  you  consider  how  many  different  ways 
thirty  (or  even  fewer)  different  amino  acids  can 
be  combined  in  long  chains  of  a  hundred  or 
more,  you  can  see  why  there  are  so  many  pro- 
teins known,  and  why  different  living  species  of 


glycine 


QUESTIONS    AND    PROBLEMS 


349 


plants  and  animals  can  have  in  their  tissues  a 
great  many  different  proteins.  Enzymes,  the  bio- 
logical catalysts,  are  also  proteins.  Each  enzyme, 
of  course,  has  its  own  particular  structure,  de- 
termined by  the  order  and  spatial  arrangement 
of  the  amino  acids  from  which  it  is  formed. 
Perhaps  the  most  marvelous  part  of  the  chem- 
istry of  living  organisms  is  their  ability  to  syn- 
thesize just  the  right  protein  structures  from  the 
myriad  of  structures  possible. 


EXERCISE  18-15 

Take  the  letters  A,  B,  C,  and  see  how  many  dif- 
ferent three  unit  combinations  you  can  make; 
for  example,  ABC,  BAC,  AAC,  CBC,  etc.  This 
will  convince  you  that  a  chain  made  of  hundreds 
of  groups  with  up  to  thirty  different  kinds  of 
units  in  each  group  can  have  an  almost  unlimited 
number  of  combinations. 


QUESTIONS  AND  PROBLEMS 


1.  What  information  is  revealed  by  the  empirical 
formula?  The  molecular  formula?  The  struc- 
tural formula?  Demonstrate,  using  ethane,  C2H6. 

2.  Write  the  balanced  equation  for  the  complete 
burning  of  methane. 

3.  Draw  the  structural  formulas  for  all  the  C2H3C13 
compounds. 

4.  Draw  the  structures  of  two  isomeric  compounds 
corresponding  to  the  empirical  formula  C3HgO. 

5.  Draw  the  structural  formulas  of  the  isomers  of 
butyl  chloride. 

6.  What  angle  would  you  expect  to  be  formed  by 
the  C,  O,  H  nuclei  in  an  alcohol  molecule? 
Explain. 

7.  When  0.601  gram  of  a  sample  having  an  em- 
pirical formula  CH20  was  vaporized  at  200°C, 
and  one  atmosphere  pressure,  the  volume  occu- 
pied was  388  ml.  This  same  volume  was  occupied 
by  0.301  gram  of  ethane  under  the  same  condi- 
tions. What  is  the  molecular  formula  of  CH20? 

One  mole  of  the  sample,  when  reacted  with 
zinc  metal,  liberated  (rather  slowly)  \  mole  of 
hydrogen  gas.  Write  the  structural  formula. 

Answer.  The  molecular  formula  is  C2H402. 

8.  A  100  mg  sample  of  a  compound  containing  only 
C,  H,  and  O  was  found  by  analysis  to  give  149 
mg  C02  and  45.5  mg  H20  when  burned  com- 
pletely. Calculate  the  empirical  formula. 

9.  How  much  ethanol  can  be  made  from  50  grams 
of  ethyl  bromide?  What  assumptions  do  you 
make  in  this  calculation? 

10.  Write  the  balanced  equation  for  the  production 
of  pentanone  from  pentanol,  using  dichromate 
ion  as  the  oxidizing  agent. 


11.  One  mole  of  an  organic  compound  is  found  to 
react  with  \  mole  of  oxygen  to  produce  an  acid. 
To  what  class  of  compounds  does  this  starting 
material  belong? 

12.  Using  the  information  given  in  Table  7-II,  deter- 
mine the  reaction  heat  per  mole  of  C-Mt(g)  for 
the  complete  combustion  of  ethane. 

13.  An  aqueous  solution  containing  0.10  mole/liter 
of  chloroacetic  acid,  ClH2CCOOH,  is  tested  wiih 
indicators  and  the  concentration  of  H+(aq)  is 
found  to  be  1.2  X  10~2  M.  Calculate  the  value 
of  KA  (if  necessary,  refer  back  to  Section  11-3.2). 
Compare  this  value  with  KA  for  acetic  acid — the 
change  is  caused  by  the  substitution  of  a  halogen 
atom  near  a  carboxylic  acid  group. 

14.  Give  simple  structural  formulas  of 

(a)  an  alcohol, 

(b)  an  aldehyde,  and 

(c)  an  acid, 

each  derived  from  methane;  from  ethane;  from 
butane;  from  octane. 

15.  Write  the  equations  for  the  preparation  of 
methylamine  from  methyl  iodide. 

16.  Write  equations  to  show  the  formation  of  the 
esters,  methyl  butyrate  and  butyl  propionate. 

17.  Given  the  structural  formula 

H  O    H    H    H 

I        /         I       I       I 
H— C— C— O— C— C— C— H 

I  III 

H  H    H    H 

for  an  ester,  write  the  formula  of  the  acid  and 
the  alcohol  from  which  it  might  be  made. 


350 


THE  CHEMISTRY  OF  CARBON  COMPOUNDS  I  CHAP.  18 


18.  How  much  acetamide  can  be  made  from  3.1 
grams  of  methyl  acetate?  See  equation  (27), 
p.  338.  Assume  the  ester  is  completely  converted. 

Answer.  2.5  grams  acetamide. 

19.  An  ester  is  formed  by  the  reaction  between  an 
acid,  RCOOH,  and  an  alcohol,  R'OH,  to  form 
an  ester  RCOOR'  and  water.  The  reaction  is 
carried  out  in  an  inert  solvent. 

(a)  Write  the  equilibrium  relation  among  the 
concentrations,  including  the  concentration 
of  the  product  water. 

(b)  Calculate  the  equilibrium  concentration  of 
the  ester  if  K  =  10  and  the  concentrations 
at  equilibrium  of  the  other  constituents  are: 

[RCOOH]  =  0.1  M; 
[R'OH]  =  0.1  M; 
[H20]  =  1.0  M.  ' 

(c)  Repeat  the  calculation  of  part  b  if  the  equi- 
librium concentrations  are: 


[RCOOH]  =  0.3  M; 
[R'OH]  =  0.3  M\ 
[H20]  =  1.0  M. 

20.  Give  the  empirical  formula,  the  molecular  for- 
mula, and  draw  the  structural  formulas  of  the 
isomers  of  butene. 

21.  There  are  three  isomers  of  dichlorobenzene  (em- 
pirical formula  C3H.C1).  Draw  the  structural 
formulas  of  the  isomers. 

/OH 

22.  Consider  the  compound  phenol, 


(a)  Predict  the  angle  formed  by  the  nuclei  C,  O, 
H.  Explain  your  choice  in  terms  of  the 
orbitals  used  by  oxygen  in  its  bonds. 

(b)  Predict  qualitatively  the  boiling  point  of 
phenol.  (The  boiling  point  of  benzene  is 
80°C.)  Explain  your  answer. 

(c)  Write  an  equation  for  the  reaction  of  phenol 
as  a  proton  donor  in  water. 

(d)  In  a  LOW  aqueous  solution  of  phenol, 
[H+]  -  1.1  X  10-5.  Calculate  K. 


SIR    ROBERT    ROBINSON,   !•••- 

Sir  Robert  Robinson  will  always  be  recognized  as  one  of 
the  outstanding  British  scientists.  With  admiration  and 
pride,  his  country  knighted  him  in  1937.  His  international 
recognition  in  organic  chemistry  was  signaled  in  1947,  when 
he  received  the  Nobel  Prize  in  Chemistry. 

The  son  of  a  surgical  dressing  manufacturer,  Robinson 
was  born  in  Chesterfield,  Derbyshire,  England.  He  received 
the  Ph.D.  at  the  University  of  Manchester,  and  his  early 
scientific  promise  led  to  a  Professorial  appointment  at  the 
University  of  Sydney  when  he  was  only  26.  Three  years 
later,  he  returned  to  England  as  Professor  at  the  University 
of  Liverpool.  As  his  reputation  grew,  he  moved  from  uni- 
versity to  university,  until,  in  1930,  he  accepted  one  of  the 
most  coveted  academic  positions  in  England— Professor  of 
Chemistry  at  Oxford. 

During  a  prolific  career  that  has  produced  an  astounding 
bibliography  of  about  600  publications,  Robinson  has  found 
leisure  time  to  scale  many  of  the  mountain  peaks  in  Switzer- 
land, Norway,  and  New  Zealand.  This  zeal  for  moun- 
taineering couples  well  with  his  interest  in  photography. 
He  is  a  formidable  chess  player  and  a  lover  of  music.  His 
wife,  also  a  chemist  of  note,  aided  him  in  his  career  while 
raising  his  family,  a  son  and  a  daughter. 

Sir  Robert  is  a  master  of  organic  synthesis.  His  research 
entered  on  the  synthesis  and  the  determination  of  struc- 

res  of  biochemically  important  substances.  These  include 

tremely  complicated  molecules  with  as  many  as  five 

sed  rings  and  a  variety  of  functional  groups  whose  posi- 
tioning and  spatial  relationships  are  critically  important 
in  fixing  biological  activity.  His  name  is  linked  with  our 
knowledge  of  certain  alkaloids— a  family  of  nitrogen  con- 
taining compounds  that  includes  some  powerful  drugs,  such 
as  strychnine  and  morphine.  His  crowning  achievements 
may  have  been  in  the  total  synthesis  of  such  important 
compounds  as  cholesterol,  cortisone,  and  other  related 
structures,  called  steroids.  During  the  second  world  war, 
Professor  Robinson  led  the  Oxford  scientific  team  that 
studied  the  chemistry  and  structure  of  the  antibiotic, 
penicillin. 

These  are  but  a  few  pinnacles  of  success  in  a  brilliant 
scientific  career.  Like  the  Alpine  peaks  that  Sir  Robert 
Robinson  loves  to  climb,  these  accomplishments  tower  high, 
to  challenge  and  inspire  this  and  future  scientific  genera- 
tions. 


CHAPTER 


19 


The  Halogens 


, 

H 

A 

A 

3 

Li 

4 

Be 

5 
B 

6 
C 

7 

N 

6 
0 

9 

F 

i 

10 

Ne 

11 
Ha 

12 

Mg 

15 

At 

14 

Si 

15 
P 

16 

s 

17 

CI 

18 
Ar 

19 

K 

20 

Ca 

21 

Sc 

22 

Ti 

23 
V 

24 
Cr 

25 

Mn 

26 

Fe 

27 

Co 

28 
Hi 

29 
Cu 

30 
Zn 

31 

Ga 

32 
Ge 

33 
As 

34- 
Se 

35 
Br- 

S3 
I 

85 

At 

36 
Kr 

37 

Fb 

30 
Sr 

39 

Y 

40 
Zr 

■41 

Hb 

4-2 

Mo 

43 
Tc 

44 
Fu 

•45 

Fh 

4-b 

Td 

4-7 

A9 

48 
Ctt 

49 

In 

50 

Sn 

51 

Sb 

52 

Te. 

54 

Xe 

55 
Cs 

5b 

Ba 

57-71 

72 
Hf 

73 

Ta 

74 

w 

75 

Fe 

7b 

Os 

77 
If 

70 
Pt 

79 
Au 

80 

61 

Tl 

82 
Pb 

83 

Bi 

64 
Po 

66 
Fn 

67 

Fr 

68 
Fa 

89- 

i        >        ' 

>        i        • 

■ ■ j j 

The  halogens  are  a  family  of  elements  appearing 
on  the  right  side  of  the  periodic  table,  in  the 
column  just  before  the  inert  gases.  The  elements 
in  this  group — fluorine,  chlorine,  bromine,  io- 
dine, and  astatine — show  some  remarkable  simi- 
larities and  some  interesting  trends  in  chemical 
behavior.  The  similarities  are  expected  since  the 


electron  populations  of  the  outer  levels  are 
analogous.  Each  element  has  one  electron  less 
than  an  inert  gas  arrangement.  The  trends,  too, 
are  understandable  in  terms  of  the  increases  in 
nuclear  charge,  number  of  electrons,  and  atomic 
size,  going  from  top  to  bottom  of  this  column 
of  the  periodic  table. 


19-1  PROPERTIES  OF  THE  HALOGENS 

You  have  already  studied  some  properties  of  the 
halogens.  They  are  very  reactive  elements  that 
exist  under  normal  conditions  as  diatomic  mole- 
cules with  covalent  bonds.  These  molecules  all 
are  colored.  Gaseous  fluorine  is  pale  yellow; 
gaseous  chlorine  is  yellow-green;  gaseous  bro- 
mine is  orange-red  (remember  the  film  equilib- 
rium); gaseous  iodine  is  violet.  The  halogens  are 
all  toxic  and  dangerous  substances.  Fluorine,  F2, 
is  the  most  hazardous;  the  danger  decreases  as 
the  atomic  number  of  the  halogen  becomes 
352 


larger.  Even  iodine,  I2,  should  be  handled  with 
caution.* 

19-1.1    Electron  Configurations  off  the  Halogens 

Table  19-1  shows  the  electron  configurations  of 
the  halogens,  using  the  symbols  introduced  in 


*  These  elements  produce  nasty  "burns"  that  are  slow 
to  heal.  The  mucous  membranes  are  attacked  especially, 
and  chlorine  "poisoning"  is  really  a  lung  inflammation. 
Under  no  circumstances  should  inexperienced  people 
handle  these  substances  without  close  guidance. 


SEC.     19-1    !    PROPERTIES    OF    THE    HALOGENS 


353 


Is 


2s 


Fz 


3s 


3p 


Br 


Br2 


all    inner 
orbital s 
filled 


Fig.   19-1.  Orbital  representations  of  the  bonding  in 
Fa  and  Br2. 


Section  15-1.6.  The  superscripts  give  the  number 
of  electrons  in  a  particular  type  of  orbital;  the 
letters  s,  p,  and  d  indicate  the  shape  of  the 
orbital;  the  number  before  the  letter  indicates 
the  principal  quantum  number  of  the  orbital. 
The  important  point  to  note  is  that  each  of  these 
halogens  has  one  less  electron  than  the  number 
required  to  fill  the  outermost  cluster  of  energy 
levels.  In  each  case  the  shortage  is  in  the  outer- 
most p  orbitals  in  which  six  electrons  can  be 
accommodated.  Therefore,  one  electron  could  be 
added  to  each  of  the  halogen  atoms  without  re- 
quiring the  population  of  additional,  higher  en- 
ergy orbitals.  Sharing  an  electron  that  spends 
some  time  in  a  valence  orbital  of  another  atom 
produces  a  covalent  bond  to  the  halogen  atom. 
Organic  compounds  (such  as  ethyl  bromide)  and 
the  halogen  molecules  (F2,  Cl>,  Br2,  I2)  contain 
covalently  bonded  halogen  atoms.  The  similar 
orbital  representations  of  the  bonding  in  F2  and 
Br2  are  contrasted  in  Figure  19-1. 
Electron  dot  representations  of  the  electron 


3d 


4s 


Ap 


sharing  in  Br2  and  I2  raise  the  question  of  how 
many  valence  electrons  need  be  shown.  Each  of 
these  two  elements  appears  in  a  row  of  the 
periodic  table  in  which  the  inert  gas  element  has 
18  electrons  in  the  orbitals  that  fix  the  chemistry 
of  the  row.  Nevertheless,  it  is  a  convenience  to 
emphasize  the  similarity  among  the  halogens  by 
showing  only  the  valence  electrons  in  the  outer- 
most s  and  p  orbitals.  Thus  the  3d  (valence) 
electrons  of  bromine  are  usually  omitted  so  that 
the  electron  dot  representations  of  F2  and  Br2 
will  appear  alike,  as  shown  in  Figure  19-2. 
Table  19-1  also  shows  the  ionization  energies 

:  F  :  F  : 


:  Br   '  Br   : 

Fig.  19-2.  Electron  dot  representations  of  the  bond- 
ing in  Fs  and  Br2.  Note  the  omission,  for 
convenience,  of  the  3d  valence  electrons  of 
bromine. 


Table  19-1.     electron  configurations  and  ionization  energies  of  the 

H  A  LO  G  E  N  S 


ELECTRON  CONFIGURATION 


NUCLEAR 
CHARGE 


inner 
electrons 


valence 
electrons 


IONIZATION 
ENERGY,  E\ 

(kcal/mole) 


fluorine 

F 

+9 

Is2 

2s-2p> 

401.5 

chlorine 

CI 

+  17 

\s22s-2fr 

3^3^ 

300 

bromine 

Br 

+  35 

■  ■  ■  3^3/j6 

3d*4s*4pr 

273 

iodine 

1 

+53 

■■■4s-Aff 

ldl05s-5fr 

241 

354 


THE    HALOGENS   I    CHAP.    19 


of  the  gaseous  halogen  atoms.  They  decrease 
significantly  as  we  move  downward  in  the  pe- 
riodic table.  Nevertheless,  all  of  these  ionization 
energies  are  very  large  compared  with  those  of 
the  alkali  metals  (compare  sodium,  whose  ioni- 
zation energy  is  118.4  kcal).  Hence  when  any  of 
the  halogens  reacts  with  an  alkali  metal,  an 
ionic  solid  is  formed.  These  ionic  solids,  or  salts, 
contain  halide  ions,  F-,  Cl~,  Br-,  or  I~,  each 
with  the  appropriate  inert  gas  electron  popula- 
tion. 

19-1.2    The  Sizes  of  Halogen  Atoms  and  Ions 

The  "size"  assigned  to  an  atom  or  ion  requires 
a  decision  about  where  an  atom  "stops."  From 
quantum  mechanics  we  learn  that  an  atom  has 
no  sharp  boundaries  or  surfaces.  Nevertheless, 
chemists  find  it  convenient  to  assign  sizes  to 
atoms  according  to  the  observed  distances 
between  atoms.  Thus,  atomic  size  is  defined  op- 
erationally— it  is  determined  by  measuring  the 
distance  between  atoms. 

For  example,  Figure  19-3  contrasts  the  dimen- 
sions assigned  to  the  halogens  in  the  elementary 
state.  One-half  the  measured  internuclear  distance 
is  called  the  covalent  radius.  This  distance 
indicates  how  close  a  halogen  atom  can  approach 


Fig.  19-3.  Covalent   radii  and   van   der   Waals  radii 

(in  parentheses)  of  the  halogens  (in  Ang- 
stroms). 


another  atom  to  which  it  is  bonded.  To  atoms  to 
which  it  is  not  bonded,  a  halogen  atom  seems 
to  be  larger.  We  can  take  as  a  measure  of  this 
size  one-half  the  distance  between  neighboring 
molecules  in  the  solid  state.  This  defines  an 
effective  radius,  the  van  der  Waals  radius,  and 
it  is  shown  by  the  black  lines  in  Figure  19-3. 
It  is  an  effective  radius,  not  a  real  radius,  be- 
cause the  electron  distribution  actually  extends 
far  out  from  the  atom. 

These  distances  aid  us  in  explaining  and  pre- 
dicting bond  lengths  in  other  covalent  halogen 
compounds.  For  example,  when  a  chlorine  atom 
is  bonded  to  a  carbon  atom  (as  in  carbon  tetra- 
chloride, CCLi),  the  bond  length  can  be  expected 
to  be  about  the  sum  of  the  covalent  radius  of  the 
carbon  atom  plus  the  covalent  radius  of  the 
chlorine  atom.  The  covalent  radius  of  carbon  is 
taken  as  0.77  A  (from  diamond),  so  the  C — CI 
bond  length  might  be  near  (0.77  +  0.99)  = 
1.76  A.  Experiment  shows  that  each  bond  length 
in  COWs  1.77  A. 

EXERCISE  19-1 

Using  the  carbon  atom  covalent  radius  0.77  A 
and  the  covalent  radii  given  in  Figure  19-3,  pre- 
dict the  C — X  bond  length  in  each  of  the  follow- 
ing molecules:  CF4,  CBr4,  CI4.  Compare  your 
calculated  bond  lengths  with  the  experimental 
values:  C— F  in  CF4  =  1.32  A,  C— Br  in  CBr4  = 
1.94  A,  C— IinCI4  =  2.15  A. 


ct, 


Br, 


SEC.     19-1    I    PROPERTIES    OF    THE    HALOGENS 


355 


F~  cr 

Fig.  19-4.  Ionic   radii   of  the  halide  ions   (in   Ang- 
stroms). 


Figure  19-4  contrasts  the  effective  sizes  of  the 
halide  ions.  Each  of  these  dimensions  is  obtained 
from  the  examination  of  crystal  structures  of 
many  salts  involving  the  particular  halide  ion. 
The  effective  size  found  for  a  given  halide  ion  is 
called  its  ionic  radius.  These  radii  are  larger  than 
the  covalent  radii  but  close  to  the  van  der  Waals 
radii  of  neutral  atoms. 

The  covalent,  van  der  Waals,  and  ionic  radii 
are  collected  in  Table  19-11  together  with  some 
physical  and  chemical  properties  of  the  halogens. 
We  see  some  interesting  trends.  For  each  type 
of  radius  we  find  a  progressive  increase  from  the 
top  of  the  column  to  the  bottom.  This  increase 
in  size  reflects  the  fact  that  as  atomic  number 
rises,  higher  energy  levels  are  used  to  accommo- 
date the  electrons.  In  addition,  Table  19-11  shows 
a  trend  of  increasing  melting  and  boiling  points 
as  we  move  downward  in  the  periodic  table. 
This  trend  is  appropriate  for  a  series  of  molecular 
solids  in  which  van  der  Waals  forces  are  the 
principal  ones  holding  the  molecules  in  proxim- 


ity.  This  type  of  force  is  higher  for  more  complex 
molecules  with  more  electrons. 

The  last  column  of  Table  19-11  shows  the  dis- 
sociation constants  for  the  reactions  of  the  type 

Xt(g)  +=£  2X(g)  U) 

with  equilibrium  constant 

where 

[X]  =  partial  pressure  of  X  atoms, 

[Xz]  =  partial  pressure  of  X2  molecules. 

Again  the  available  data  show  a  trend;  K  in- 
creases in  the  series  CU,  Br2,  I2.  The  equilibrium 
conditions  are  fixed  by  two  factors:  tendency  to 
minimum  energy  and  tendency  to  maximum 
randomness.  The  randomness  factor  is  about  the 
same  for  the  three  halogens,  so  the  trend  in  equi- 
librium constants  is  largely  determined  by  the 
energy  effects.  In  each  dissociation  reaction  (of 
CI2,  Br2,  and  I2)  energy  is  absorbed.  However, 
more  energy  is  absorbed  to  break  the  Br2  bond 
than  to  break  the  I2  bond,  and  still  more  to 
break  the  Cl2  bond.  The  energy  absorbed  ac- 
counts for  the  trend  in  equilibrium  constants. 


Table  19-11.     sizes  of  the  halogen  atoms,  melting  points,  boiling  points, 

AND    DISSOCIATION    PROPERTIES    OF    THE    HALOGEN    MOLECULES 


HALOGEN,  X 


COVALENT 
RADIUS 
IN  Xi 


VAN  DER  WAALS 
RADIUS  IN  Xt 


IONIC 
RADIUS 

(-1  ion) 


M.P.  OF 

X, 


B.P.  OF 

X, 


BOND 

DISSOCIATION 

ENERGY 

CONSTANT  OF 

OF  Xt 

Xt  AT  1000°C 

fluorine,  F 

0.72  A 

1.35  A 

1.36  A 

55°K 

85°K 

36     kcal 

— 

chlorine,  CI 

0.99 

1.80 

1.81 

172°K 

238.9°K 

57.1 

io-» 

bromine,  Br 

1.14 

1.95 

1.95 

265.7°K 

331.8°K 

45.5 

8  X  10-» 

iodine,  I 

1.33 

2.15 

2.16 

387°K 

457°K 

35.6 

10-' 

356 


THE    HALOGENS    I    CHAP.    19 


Why  should  the  bond  energy  be  greater  for  Cl2 
than  for  Br2,  and  greater  for  Br2  than  for  I2? 
Presumably  the  size  effect  is  a  factor.  Two  halo- 
gen atoms  remain  together  because  a  pair  of 
electrons  is  simultaneously  near  both  nuclei. 
However,  the  larger  the  halogen  atoms,  the  more 
distant  are  those  bonding  electrons  from  the 
nuclei.  Since  the  electrical  forces  decrease  with 
distance,  the  bond  energy  lessens. 

The  dissociation  constant  of  F2  at  1000°C  is 
not  known  and  the  bond  energy  of  F2  was  only 
learned  recently.  Almost  all  chemists  were  sur- 
prised when  experiment  showed  that  the  energy 


necessary  to  break  the  bond  in  F2  is  much  lower 
than  that  in  Cl2.  This  is  still  not  well  explained 
so,  rather  than  abandon  familiar  arguments  con- 
cerning halogen  properties  based  on  the  trend 
in  size,  chemists  treat  fluorine  as  a  special  case. 


EXERCISE  19-2 

On  the  basis  of  the  trend  in  atomic  size,  what 
trend  is  expected  in  the  ionization  energy  Ex  of 
the  halogen  atoms?  Compare  your  prediction 
with  the  actual  trend  in  Eu  given  in  Table  19-1. 


19-2    HALOGEN  REACTIONS  AND  COMPOUNDS 


Most  of  the  reactions  of  the  halogens  are  of  the 
oxidation-reduction  type.  The  halogens  are  so 
reactive  that  they  do  not  occur  uncombined  in 
nature  and  they  must  be  made  from  halide  com- 
pounds (salts).  We  shall  consider  briefly  the 
preparation  of  the  elements  and  then  explore 
some  of  the  very  interesting  chemistry  of  this 
family. 


made  by  electrolysis  of  molten  sodium  chloride 
(dissolved  in  CaCl2  to  lower  the  melting  point). 
Figure  19-5  shows  the  components  of  the  elec- 
trolysis cell.  At  one  electrode  molten  sodium  is 
produced,  and  at  the  other,  chlorine  gas  is 
collected. 


19-2.1    Preparation  of  the  Halogens 

Oxidation  through  electrolysis  is  used  to  make 
fluorine  and  chlorine.  Chlorine,  for  example,  is 


Fig.  19-5.  Preparation    of    chlorine    and    sodium    by- 
electrolysis  of  molten  NaCl. 


Na(t) 


EXERCISE  19-3 

At  the  left  electrode  in  Figure  19-5  the  half- 
reaction  occurring  is  CI"  — *-  \C\i(g)  +  e~,  and 
at  the  right  electrode  the  half-reaction  is 
Na+  +  e~  — >-  Na(l).  Which  electrode  is  the 
anode  and  which  is  the  cathode?  With  these 
half-reactions,  balance  the  net  reaction  occurring 
in  the  electrolysis  cell. 


Gaseous  fluorine  is  also  prepared  by  electroly- 
sis of  molten  fluoride  salts  but  simpler  methods 
are  available  for  the  preparation  of  bromine  and 
iodine.  Chemical  oxidation,  usually  with  chlo- 
rine as  the  oxidizing  agent,  provides  Br2  and  I2 
economically  because  chlorine  is  a  relatively  in- 
expensive chemical.  The  reactions  are 


I/a  CI  (t)  in   molten  CaCl2 


2l-(aq)  +  Ch(g) 
2Br(aq)  +  Cl2(g) 


h(s)  +  2Cl-(aq)       (2) 

Bu(g)  +  2C\-(aq)    (3) 


SEC.     19-2    I    HALOGEN    REACTIONS    AND    COMPOUNDS 


357 


EXERCISE  19-4 


From  the  £°'s  for  the  half-reactions  of  the  type 
2X~  — ►■  X2  +  2e~,  show  that  Cl2  can  be  used 
to  produce  Br.  from  Br~  and  I2  from  I~  but  not 
to  produce  F2  from  F~. 


19-2.2    Reduction  of  the  Halogens 

A  substance  that  is  readily  reduced  is  a  good 
oxidizing  agent.  The  oxidizing  abilities  of  the 
halogens  vary  in  a  regular  manner,  fluorine  being 
the  strongest  and  iodine  the  weakest.  On  the 
other  hand,  the  iodide  ion  sometimes  acts  as  a 
reducing  agent,  while  fluoride  ion  never  does. 
These  statements  are  reflected  in  the  £°  values 
for  the  oxidation  half-reactions: 


2F" 
2C1- 
2Br" 

21" 


F,  +  2e~ 
Cl,  +  2e~ 
Br .  +  2e- 
h  +  2e~ 


E° 

E° 
E° 
E° 


-2.87  volts  (4) 

-1.36  (5) 

-1.07  (6) 

-0.53  (7) 


The  trend  in  oxidizing  ability  is  quantitatively 
expressed  in  the  trend  in  £°'s.  The  E°  for  reac- 
tion (4),  —2.87  volts,  indicates  that  fluoride  ion, 
F~,  has  little  tendency  to  release  electrons.  Con- 
versely, F2  has  a  high  tendency  to  acquire  elec- 
trons. Hence,  F2  is  readily  reduced — F2  is  a 
powerful  oxidizing  agent.  At  the  bottom  of  the 
list,  E°  for  reaction  (7)  is  —0.53  volt.  Iodide 
ion,  I-,  has  a  moderate  tendency  to  release 
electrons,  oxidizing  I-  to  I2.  Thus  iodide  ion  has 
a  moderate  tendency  to  be  oxidized,  acting  as  a 
reducing  agent.  Conversely,  I2  has  only  a  moder- 
ate tendency  to  acquire  electrons.  Iodine  has  a 
moderate  tendency  to  be  reduced — 12  can  act  as 
an  oxidizing  agent. 


How  can  this  trend  in  half-cell  potentials  and  oxidizing 
abilities  be  explained?  Let  us  imagine  that  reaction  (7), 
as  an  example,  is  carried  out  in  a  hypothetical  series  of 
simpler  steps: 


2l-(aq) 

2l~(g) 

21(g) 

h(g) 


2\~(g) 

2\(g)  +  2e~(g) 

Ug) 

U(s) 


Dehydration  (5) 

Electron  removal  (9) 

Molecule  formation  (70) 

Condensation  (77) 


2\~(aq) 


h(s)  +  2e~(g)     Net  reaction 


U2) 


We  can  consider  the  energy  effect  that  accompanies  the 
net  reaction  (72)  in  terms  of  the  energy  effects  of  the 
hypothetical  steps.  Since  the  tendency  toward  minimum 
energy  is  one  of  the  factors  that  fixes  equilibrium,  the 
energy  change  is  one  of  the  factors  influencing  E°.  Step 
(8)  is  a  dehydration  step  in  which  the  I~  ions  are  pulled 
out  of  the  water  into  the  gas  phase.  In  this  step,  heat  is 
absorbed;  since  the  system  gains  energy,  AH  is  positive 
for  step  (8).  In  step  (9),  electrons  are  pulled  off  the  gaseous 
I~  ions  to  give  neutral  gaseous  I  atoms;  this  step  also 
requires  energy,  and  again  AH  is  positive.  In  step  (70), 
two  atoms  of  iodine  come  together  to  form  a  diatomic 
molecule.  This  is  the  opposite  of  breaking  the  bond  in  I»; 
energy  is  liberated  and  the  system  loses  energy,  so  AH  is 
negative  for  this  step.  In  step  (77),  the  gaseous  U  mole- 
cules form  solid  iodine  crystals;  this  is  the  reverse  of 
vaporizing  the  solid  and  also  corresponds  to  liberation  of 
energy.  The  question  as  to  which  halide  ion  is  the  best 
reducing  agent  depends  primarily  upon  the  net  energy 
change  considering  all  the  steps  involved  in  the  half- 
reaction.  The  halide  ion  that  requires  the  least  energy  to 
convert  it  to  the  halogen  should  be  the  best  reducing 
agent.  Table  19-111  shows  the  experimentally  measured 
requirement  for  each  of  the  above  steps  for  the  four 
halogens  (using  ^  as  a  general  symbol  for  a  halogen 
atom). 

Notice  that  in  Table  19-111  dehydration  and  electron 
removal  are  the  steps  that  involve  the  largest  energy 
changes.  The  amount  of  energy  required  for  each  of  these 
processes  diminishes  as  atomic  weight  increases  for  the 
three  large  halogens.  For  fluorine,  dehydration  accounts 
for  more  than  half  the  total  energy,  but  electron  removal 
is  still  a  major  part  of  the  positive  energy  change.  The 


Table  19-111.     the  energy  required  for  some  halogen  reactions 

ELECTRON  REMOVAL              MOLECULE              CONDENSATION 
HALO-  DEHYDRATION  2X~(g)  ►■  FORMATION  X%(g)  >-   X2  OVERALL  REACTION 

gen       2X~(aq)  — »-  2X~(g)         2X(g)  +  2e~(g)     ?X(g)  — >■  X2(g)     (normal  state)    2X~(aq)  — ►-  Xt  +  2e~ 


F 

+246  kcal 

+  162  kcal 

-37  kcal 

-1.6  kcal 

+  369  kcal 

CI 

+  178kcal 

+  171  kcal 

-58  kcal 

—  4.4  kcal 

+287  kcal 

Br 

+  162  kcal 

+  161  kcal 

-46  kcal 

-7     kcal 

+270  kcal 

I 

+  144  kcal 

+  146  kcal 

-36  kcal 

-10  kcal 

+244  kcal 

358 


THE    HALOGENS    I    CHAP.     19 


values  in  the  last  column  show  that  the  energy  change 
for  the  overall  reaction  forms  a  regular  series,  roughly 
comparable  to  the  variation  of  E°.  The  half-reactions 
having  the  most  negative  E°  are  those  that  require  the 
most  energy,  and  that  show  the  least  tendency  to  proceed 
as  written  from  left  to  right.  Therefore,  the  energies  listed 
in  the  last  column  do  help  to  explain  why  the  iodide  ion 
is  a  better  reducing  agent  than  the  fluoride  ion.  This  also 
explains  why  F2  oxidizes  compounds  better  than  do  the 
other  halogens.  In  aqueous  solution  the  affinity  of  fluorine 
for  electrons,  plus  the  rather  strong  attraction  between 
water  molecules  and  F~,  make  F2  a  good  oxidizing  agent. 

The  halogens  are  reactive  even  without  water. 
All  the  halogens  react  quite  vigorously  with  most 
of  the  metals  to  produce  simple  halide  salts. 
Copper  and  nickel,  however,  appear  to  be  quite 
inert  to  F2.  This  apparent  inertness  is  attributed 
to  the  fact  that  a  thin  layer  of  the  fluoride  salt 
forms  on  the  surface  of  each  of  those  metals  and 
protects  it  from  further  attack  by  fluorine. 

19-2.3    lodimetry 

The  I~-I2  half-reaction  has  many  applications  in 
aqueous  solution  chemistry.  The  use  of  I-  as  a 
reducing  agent  and  I2  as  an  oxidizing  agent, 
particularly  for  quantitative  purposes,  is  called 
iodimetry. 

This  half-reaction  possesses  an  E°  of  —0.53 
volt;  neither  is  I~  a  particularly  powerful  re- 
ducing agent  nor  is  I2  a  particularly  powerful 
oxidizing  agent: 


21 


I2  +  2e~        E°  =  -0.53  volt      (13) 


There  are  many  half-reactions  below  this  one  in 
Appendix  3,  so  there  are  quite  a  few  substances 
that  will  oxidize  I~.  For  example,  iodide  ion  can 
be  quantitatively  oxidized  to  I2  by  Fe+3,  Br2, 
Mn02,  Cr2Of2,  Cl2,  and  Mn04",  On  the  other 
hand,  there  are  many  half-reactions  above 
E°  =  —0.53  volt  in  Appendix  3.  For  example, 
I2  can  be  quantitatively  reduced  to  I-  by  Sn+2, 
H2S03,  and  Cr+2.  The  usefulness  of  the  /~-/2  re- 
action derives  from  the  fact  that  all  of  the  sub- 
stances mentioned  react  rapidly  and  without  side 
reactions. 

To  top  off  this  versatility,  iodine  possesses  an 
unusually  sensitive  and  specific  indicator.  Iodine 
reacts  with  starch  to  give  a  blue-colored  complex. 
This  complex  is  so  intensely  colored  that  I2  can 
be  detected  at  a  concentration  as  low  as  10-6  M 


and  it  furnishes  the  basis  for  the  qualitative  test 
for  iodine  known  as  the  "starch-iodine"  test. 
More  important,  the  complex  serves  as  a  sensi- 
tive indicator  in  oxidation-reduction  titrations 
based  upon  the  I~-I2  half-reaction. 


EXERCISE  19-5 

Balance  the  reaction  that  occurs  when  I-  is 
oxidized  to  I2  by  Mn04~  in  acid  solution,  produc- 
ing Mn+2. 


19-2.4    Positive  Oxidation  States 

of  the  Halogens:  The  Oxyacids 

The  halogens,  except  fluorine,  can  be  oxidized  to 
positive  oxidation  states.  Most  commonly  you 
will  encounter  these  positive  oxidation  states  in  a 
set  of  compounds  called  "halogen  oxyacids"  and 
their  ions. 

Compounds  of  the  type  HC103,  HC104,  etc., 
are  examples  of  the  halogen  oxyacids.  Chlorine 
and  iodine  form  a  series  of  these  acids  in  which 
the  halogen  oxidation  number  can  be  +1,  +3, 
+5,  or  +7.  For  chlorine  the  series  is  made  up 
of  HCIO  (hypochlorous  acid),  HC102  (chlorous 
acid),  HCIO3  (chloric  acid),  and  HC104  (per- 
chloric acid).  Although  it  is  not  easy  to  handle 
these  unstable  substances,  aqueous  solutions  of 
these  acids  have  been  examined  to  find  out  how 
strong  they  are  as  proton  donors.  HCIO  is  a  weak 
proton  donor,  HC102  is  somewhat  stronger, 
HCIO3  is  quite  strong,  and  HC104  is  the  strongest 
of  all.  (Perchloric  acid,  HC104,  is,  in  fact,  one 
of  the  strongest  acids  known.) 

Now  we  might  wonder  how  to  account  for  the 
observed  trend  in  terms  of  structure  and  bond- 
ing. Figure  19-6  shows  the  presumed  positions 
of  the  atoms  in  these  molecules.  It  is  evident  that 
in  each  case  we  need  to  break  a  hydrogen- 
oxygen  bond  to  split  off  the  proton.  A  regular 
decrease  in  the  strength  of  the  hydrogen-oxygen 
bond  as  we  proceed  from  chlorous  to  perchloric 
acid  would  explain  the  trend  in  acidity.  How  do 
we  account  for  the  fact  that  the  strength  of  this 
bond  varies  as  we  go  through  the  sequence? 
Formally,  we  say  that  the  oxidation  number  of 
chlorine  ranges  from  +1  to  +3  to  +5  to  +7 


SEC.     19-2    '    HALOGEN    REACTIONS    AND    COMPOUNDS 


359 


NAME 


FORMULA         BALL-AMD- STICK MODEL        SPACEFILLING  MODEL 


Hypochlorous  acid  HO  CI 


1 


Chlorous  acid  HO  CIO 

(HClOz) 


3 


Chloric    acid 


HOCIOZ 
(HClOs) 


*% 


Perchloric    acid  HOClOj 

(HCIO+) 


Fig.  19-6.  Presumed  structures  of  the  chlorine  oxy acids. 


360 


THE    HALOGENS    I    CHAP.     19 


across  the  set,  but  actually  there  are  no  Cl+1, 
Cl+3,  Cl+5,  or  Cl+7  ions  in  these  acids.  The  oxida- 
tion number  is  only  an  artificial  way  of  keeping 
count  of  electric  charges,  as  we  learned  in  Sec- 
tion 12-3.3.  What  is  more  to  the  point  (and  this 
is  really  why  the  oxidation  number  cnanges  in 
the  first  place)  is  that  there  is  an  increasing  num- 
ber of  oxygen  atoms  bonded  to  the  central 
chlorine.  Each  time  an  additional  oxygen  bonds 
to  the  chlorine  atom,  some  electron  charge  is 
drawn  off  the  chlorine  and  hence  away  from  the 
original  O — CI  bond.  This,  in  turn,  draws  elec- 
trons from  the  adjacent  H — O  bond  and  thereby 
weakens  it. 

This  increase  in  acid  strength  with  oxidation 
number  is  a  general  phenomenon.  For  example, 
nitric  acid  (HNO3,  in  which  the  oxidation  num- 
ber of  N  equals  +5)  is  stronger  than  nitrous 
(HN02,  oxidation  number  +3);  sulfuric  acid 
(H2SO4,  in  which  the  oxidation  number  of  S 
equals  +6)  is  stronger  than  sulfurous  (H2S03, 
oxidation  number  +4).  A  consistent,  hence  use- 
ful, explanation  is  found.  When  an  oxygen  atom 
is  added  to  the  central  atom,  there  is  a  reduction 
of  the  strength  of  O — H  bonds  in  attached  OH 
groups. 

The  halogen  oxyacids  and  their  anions  are 
quite  easily  reduced — they  are  good  oxidizing 
agents.  When  one  of  these  acts  as  an  oxidizing 
agent,  the  halogen  is  reduced  to  a  lower  oxida- 
tion number.  Just  what  oxidation  number  is 
attained  depends  upon  a  variety  of  factors,  in- 
cluding acidity  of  the  solution,  strength  of  the 
reducing  agent,  amount  of  reducing  agent,  and 
temperature. 

For  example,  the  common  use  of  sodium  hy- 
pochlorite solution,  NaOCl,  as  a  bleaching  solu- 
tion depends  upon  the  oxidizing  action  of  hypo- 
chlorite, OCl~.  Iodate  ion,  I03",  also  furnishes 
a  strong  oxidizing  power,  as  shown  by  E°  for  the 
half-reaction  12-10^ 

I2  +  6H20  — >-  2IO3-  +  12H+  +  10e- 

E°  =  -1.2  volts    (14) 

The  chlorine  and  bromine  counterparts,  chlo- 
rate, CIO3",  and  bromate,  Br03",  have  E0,s  that 
are  even  more  negative.  Hence  these  ions  are 
even  stronger  oxidizing  agents  than  iodate  ion. 
It  is  not  unusual  in  halogen  chemistry  to  find 
striking  differences  in  the  chemistry  of  acidic  and 


basic  solutions.  For  example,  iodine  in  an  acidic 
solution  is  quite  stable,  but  in  a  basic  solution 
it  is  spontaneously  oxidized  to  oxidation  number 
+  5  in  the  IO:i~  ion.  The  reason  for  this  can  be 
seen  by  considering  the  half-reactions  (15)  and 
(16): 

21-  — >-  h  +  2e~ 

E°  =  -0.53  volt      (15) 

h  +  6H-.0  — ►■  2IO3-  +  12H-  +  10«r 

E°  =  -1.2  volts      (16) 

The  E°  for  reaction  (75)  is  more  negative  than 
that  of  reaction  (75),  hence  iodate  ion  can  react 
with  I~.  In  1  M  acid  solution,  the  E°  for  the  net 
reaction 


6I2  +  6H20 


2IO3-  4-  101-  -h  12H+    (17) 


is  (-1.2  volts)  -  (-0.53  volt)  or  -0.7  volt. 
Since  E°  of  the  net  reaction  is  so  negative,  we 
expect  that  the  reaction  will  not  proceed  spon- 
taneously, and  it  doesn't.  Let  us  now  see  what 
happens  when  the  H+  concentration  becomes  so 
low  that  the  OH-  concentration  is  1  M.  When  a 
reaction  occurs  in  basic  solution  it  is  conven- 
tional to  show  OH-  (rather  than  H+)  in  the 
balanced  reaction.  Therefore,  for  the  reaction  of 
iodine  in  basic  solution,  half-reaction  (16)  be- 
comes 


I2  +  120H- 


2IO3-  +  6H20  +  10<?-    (18) 


Le  Chatelier's  Principle  aids  us  in  predicting  how 
the  tendency  for  I2  to  release  electrons  in  (18) 
will  be  affected  if  we  raise  the  hydroxide  ion 
concentration.  Raising  the  concentration  of  a 
reactant  (such  as  OH-)  tends  to  favor  products. 
Hence  E°  for  (18)  (1  M  OH~)  will  be  more  posi- 
tive than  E°  for  (76)  (1  M  H+).  In  1  M  OH~, 
E°  =  —0.23  volt.  Since  this  E°  is  now  more  posi- 
tive than  E°  for  reaction  (75)  (E°  =  -0.53  volt), 
the  net  reaction  (79)  can  occur: 


6I2  +  120H- 


21O3-  +  101-  +  6H20    (79) 


This  reaction  now  has  an  E°  of  (  —  0.23  volt)  — 
(-0.53  volt)  or  +0.30  volt.  With  this  positive 
value  of  E°,  we  can  expect  that  in  basic  solution 
the  reaction  will  proceed  spontaneously  (if  the 
rate  is  rapid),  and  it  does. 

The  industrial  preparation  of  bromine  takes 
advantage  of  this  effect  of  hydrogen  ion  concen- 
tration on  the  direction  of  the  spontaneous  re- 


SEC.     19-2    I    HALOGEN    REACTIONS    AND    COMPOUNDS 


361 


action.  A  dilute  solution  of  bromine  is  produced 
by  chlorination  of  salt  well  brines: 


EXERCISE  19-6 


Cl2  +  2Br- 


Br2  +  2C1- 


(20) 


The  liberated  bromine  is  carried  by  a  stream  of 
air  into  an  alkaline  solution  of  sodium  carbonate 
where  it  dissolves  as  a  mixture  of  bromide  and 
bromate  [the  analogy  of  reaction  (19)].  This  last 
step  serves  to  concentrate  the  product,  and  free 
bromine  is  obtained  by  subsequent  acidification 
of  the  solution,  through  the  reaction 


6H+  +  5Br~  +  BrOi 


3Br2  +  3H20    (27) 


An  alternative  process  for  preparing  bromine 
from  sea  water  begins  again  with  reaction  (20). 
The  liberated  bromine  is  produced  at  a  very  low 
partial  pressure  and  it  is  necessary  to  concentrate 
it.  This  is  accomplished  through  the  reaction 
between  S02  and  Br2 


Br2(gj  +  SO,(g)  +  2H20(g)  — »- 

2HBrfgJ  +  2HSO<(g) 


(22) 


The  resulting  acid  vapors  have  a  great  affinity 
for  water  (do  you  remember  how  rapidly  HC1 
dissolved  in  water  in  the  film,  gases  and  how 
they  combine?).  Hence  the  HBr  rapidly  dis- 
solves in  water  and  concentrations  as  high  as 
0.5  M  can  be  reached,  a  thousandfold  more 
concentrated  than  original  sea  water.  With  this 
concentration,  chlorine  is  again  introduced  to 
produce  Br2  by  reaction  (20). 


19-2.5     Self -Oxidation-Reduction: 
Disproportionate 

In  reaction  (79)  the  iodine  shown  on  the  left  has 
an  oxidation  number  of  zero.  After  the  reaction, 
some  of  the  iodine  atoms  have  oxidation  number 
+5  and  some  —1.  In  other  words,  the  iodine 
oxidation  number  has  gone  both  up  and  down 
in  the  reaction.  This  is  an  example  of  self- 
oxidation-reduction,  sometimes  called  dispropor- 
tionation.  It  is  a  reaction  quite  typical  of,  but  not 
at  all  restricted  to,  the  halogens. 

When  chlorine  gas  is  bubbled  into  a  solution 
of  NaOH,  self-oxidation-reduction  occurs  to  give 
hypochlorite  ion,  C10~,  by  the  reaction 

Cl2  +  20H-  — +■  a-  +  CIO"  +  H20     (23) 


Show  that  when  an  aqueous  solution  of  NaCl 
is  being  electrolyzed,  vigorous  stirring  in  the  cell 
might  permit  reaction  (23)  to  occur. 


Going  one  step  further,  if  a  basic  solution 
containing  hypochlorite  ion  is  heated,  the  CIO- 
can  again  disproportionate, 

3C10-  — ►-  2C1"  +  C103-  (24) 

this  time  to  produce  the  chlorate  ion,  C103". 

19-2.6    Special  Remarks  on  Fluorine 

Because  of  the  small  size  of  the  atom,  fluorine  is 
rather  special  in  the  halogen  group.  We  have 
already  seen  that  it  is  a  strong  oxidizing  agent 
in  aqueous  solution,  and  that  a  large  part  of  this 
arises  because  of  the  large  hydration  energy 
associated  with  the  fluoride  ion.  Another  way  in 
which  fluorine  reveals  its  special  character  is  in 
the  properties  of  hydrogen  fluoride  compared 
with  the  other  hydrogen  halides.  These  are  ex- 
plained in  terms  of  a  special  attraction  of  fluorine 
for  protons — an  attraction  called  hydrogen 
bonding  (reread  Section  17-2.6). 

The  strong  attraction  of  fluorine  for  protons 
shows  up  in  another  way.  In  aqueous  solution, 
HF  is  a  weak  acid  whereas  HC1,  HBr,  and  HI 
are  strong  acids.  The  dissociation  constant  of 
HF  is  6.7  X  10-4,  so  hydrofluoric  acid  is  less 
than  10%  dissociated  in  a  0.1  M  HF  solution. 

Another  contrast  between  HF  and  the  other 
hydrogen  halides,  HC1,  HBr,  and  HI,  is  found 
in  the  reactivity  with  glass.  Hydrofluoric  acid 
cannot  be  stored  in  glass  bottles  because  it  etches 
the  silica,  Si02,  in  glass.  On  the  other  hand,  even 
the  most  concentrated  hydrochloric  acid  solu- 
tions can  be  stored  indefinitely  in  glass  without 
any  evidence  of  a  comparable  reaction.  To  store 
HF  solutions,  we  must  use  either  polyethylene  or 
wax  containers  (rather  than  glass)  because  of 
this  reactivity  with  silica.  Silicon  bonds  more 
strongly  to  fluorine  than  to  oxygen  and  hence 
silica  dissolves  in  a  solution  of  HF  by  the  reac- 
tion 

Si02  +  6HF  — >-  SiF6"3  +  2H,0+        (25) 


362 


THE    HALOGENS    |    CHAP.    19 


Hydrofluoric  acid  is  a  polar  material,  as  water 
is,  and  it  behaves  as  an  ionizing  solvent  when  it 
is  scrupulously  free  of  water.  Salts  that  dissolve 
readily  in  liquid  HF  include  LiF,  NaF,  KF,  AgF, 
NaN03,  KN03,  AgN03,  Na2S04,  K2S04,  and 
Ag2S04.  Liquid  HF  also  dissolves  organic  com- 
pounds and  is  used  as  a  solvent  for  a  variety  of 
reactions. 

A  very  stable  bond  involving  fluorine  is  the 
carbon-fluorine  bond.  The  strength  of  this  C — F 
bond  is  comparable  to  the  C — H  bond,  and  has 
led  to  the  existence  of  a  series  of  compounds 
known  as  the  fluorocarbons.  These  are  analogous 
to  the  hydrocarbons  and  can  be  imagined  as 
being  derived  from  them  by  substituting  F  atoms 
for  H  atoms.  For  example, 


F— C— C— F 

/  \ 

F  F 

is  the  fiuorocarbon  analogue  of  ethane.  It  is 
called  perfluoroethane.  Many  of  the  fluorocar- 
bons are  quite  inert  and  their  uses  exploit  this 
property.  CC12F2,  Freon,  is  a  volatile,  nonpoison- 
ous,  noncorrosive  material  used  in  refrigerators 
and  as  the  propellant  in  some  aerosol  cans. 
Increasingly  important  also  are  the  polymeric 
fluorocarbons,  such  as  Teflon,  which  are  derived 
from  perfluoroethylene,  CF2=CF2,  by  polymeri- 
zation. They  can  be  used,  for  example,  as 
gaskets,  valves,  and  fittings  for  handling  ex- 
tremely corrosive  chemicals. 


QUESTIONS  AND  PROBLEMS 


1.  Give  the  electron  configuration  for  each  of  the 
trio  F-,  Ne,  Na+.  How  do  the  trios  CI",  Ar,  K+, 
and  Br-,  Kr,  Rb+  differ  from  the  above? 

2.  Table  19-11  contains  values  for  the  covalent 
radii  and  the  ionic  radii  of  the  halogens.  Plot 
both  radii  versus  row  number.  What  systematic 
changes  are  evident  in  the  two  curves  ? 

3.  Using  the  data  from  Table  19-11,  plot  on  one  set 
of  axes  the  melting  and  boiling  points  of  the 
halogens  versus  row  number. 

4.  For  astatine,  use  your  graphs  from  Problems  2 
and  3  as  a  basis  for  a  prediction  of  its  covalent 
radius,  ionic  radius  of  the  —1  ion,  melting  point, 
and  boiling  point. 

5.  Predict  the  molecular  structures  and  bond 
lengths  for  SiF4,  SiCLj,  SiBr4,  and  Sil4,  assuming 
the  covalent  radius  of  silicon  is  1.16  A. 

6.  Explain  in  terms  of  nuclear  charge  why  the  K+ 
ion  is  smaller  than  the  Cl~  ion,  though  they  are 
isoelectronic  (they  have  the  same  number  of 
electrons). 

7.  Can  aqueous  bromine,  Br2,  be  used  to  oxidize 
ferrous  ion,  Fe+*(aq),  to  ferric  ion,  Fe+3(aq) 
(use  Appendix  3)?  Aqueous  iodine,  I2? 

8.  What  will  happen  if  F2  is  bubbled  into  1  M  NaBr 
solution?  Justify  your  answer  using  E°  values. 


9.  Using  E°  values,  predict  what  will  happen  if,  in 
turn,  each  halogen  beginning  at  chlorine  is  added 
to  a  1  M  solution  of  ions  of  the  next  lower  halo- 
gen: Cl2  to  Br~,  Br2  to  I".  Which  halogen  is 
oxidized  and  which  is  reduced  in  each  case? 

10.  Write  a  balanced  equation  for  the  reaction  of 
dichromate  and  iodide  ions  in  acid  solution. 
Determine  E°  for  the  reaction 

Ct2Oj  2(aq)  +  l-(aq)  +  H+(aq)  gives 

Cr+3(aq)  +  I2  +  H20 

Answer.  E°  =  +0.80  volt. 

11.  Balance  the  equation  for  the  reaction  of  iodine 
with  thiosulfate  ion: 


k-2/ 


>-2/ 


I2  +  SA"  (aq)  gives  S4Oe    («<?J  +  ^~(»<l) 

thiosulfate  tetrathionate 

ion  ion 

What  is  the  oxidation  number  of  sulfur  in  the 
tetrathionate  ion? 

12.  How  many  grams  of  iodine  can  be  formed  from 
20.0  grams  of  KI  by  oxidizing  it  with  ferric 
chloride  (FeCl3)?  Determine  E°. 

Answer.  15.3  grams  of  I2. 

13.  Balance  the  equation  for  the  reaction  between 
S02  and  I2  to  produce  S04~2  and  I"  in  acid  solu- 
tion. Calculate  E°.  From  Le  Chatelier's  Prin- 
ciple, predict  the  effect  on  the  E°  in  this  reaction 
if  H+  =  10~7  M  is  used  instead  of  H+  -  1  M. 


QUESTIONS    AND    PROBLEMS 


363 


14.  What  is  the  oxidation  number  of  the  halogen 
in  each  of  the  following:  HF,  HBr02,  HI03, 
CIO3" ,  F2)  CIO,"  ? 

15.  Comparable  half-reactions  for  iodine  and  chlo- 
rine are  shown  below. 

|I,  +  3H.O  — >-  ICV  +  6H+  +  Se~ 

E°  =  -1.195  volts 
2C12  +  3H20  — *-  CIO3-  +  6H+  +  5e~ 

E°  =  -1.47  volts 

(a)  Which  is  the  stronger  oxidizing  agent,  iodate, 
IO3- ,  or  chlorate,  C103_  ? 

(b)  Balance  the  equation  for  the  reaction  be- 
tween chlorate  ion  and  I"  to  produce  I2  and 
Cl2. 

16.  Two  half-reactions  involving  chlorine  are 

2C1-  — ►-  Cl2  +  2tr 

E°  =  -1.36  volts 
Cl2  +  2H20  — ►-  2HOC1  +  2H+  +  2e~ 

E°  =  -1.63  volts 

(a)  Balance  the  reaction  in  which  self-oxidation- 
reduction  of  Cl2  occurs  to  produce  chloride 
ion  and  hypochlorous  acid,  HOC1. 

(b)  What  is  the  oxidation  number  of  chlorine  in 
each  species  containing  chlorine? 


(c)  What  is  £°  for  the  reaction  ? 

(d)  Explain,  using  Le  Chatelier's  Principle,  why 
the  self-oxidation-reduction  reaction  occurs 
in  1  M  OH~  solution  instead  of  1  M  H+. 

17.  How  many  grams  of  Si02  would  react  with 
5.00  X  102  ml  of  1.00  M  HF  to  produce  SiF4? 

18.  A  water  solution  that  contains  0.10  M  HF  is  8% 
dissociated.  What  is  the  value  of  its  KA1 

Answer.  6.9  X  10~4. 

19.  From  each  of  the  following  sets,  select  the  sub- 
stance which  best  fits  the  requirement  specified. 


(a)  Strongest  acid 

(b)  Biggest  atom 

(c)  Smallest  ionization 
energy 

(d)  Best  reducing  agent 

(e)  Weakest  acid 


HOC1,  HOCIO, 

HOCIO, 
F,  CI,  Br,  I 

F,  CI,  Br,  I 
F",  CI",  Br",  I- 
HF,  HC1,  HBr,  HI 


(0  Best  hydrogen  bonding    HF,  HC1,  HBr,  HI 

20.  Describe  two  properties  that  the  halogens  have 
in  common  and  give  an  explanation  of  why  they 
have  these  properties  in  common. 


CHAPTER 


20 

The  Third  Row  of 
the  Periodic  Table 


/ 
H 

z 
He 

3 

4 
** 

s       « 

B      C 

7 

N 

6 
O 

9 

F 

10 
Ne 

it 

Na 

m 

IS 
At 

14 
Si 

15  1  16 

P  \  s 

17 

CI 

A 

Ca 

Se 

22 

Ti 

23 
V 

24 

Cr 

29 
M* 

26 

27 
Co 

2B 

m 

29 

Cm 

SO 
In 

Si 

At 

32 

ss 

As 

34 
St 

Z5 
Br- 

36 
Kf 

SS 
Sr 

39 

Y 

4Q 

Zr 

4t 
Ub 

42 
Mo 

43 

44 

4$ 

Xh 

44 

47 

4$ 
Cd 

49 

In 

to 
Sn 

Si 
Sb 

32 

7* 

S3 
I 

54 

Xe 

ss 

Cs 

Ba 

ST-fi 

72 

73 

Ta 

74 

W 

rs 

76 
09 

77 
I* 

» 

79 

Au 

% 

St 

Ti 

St 

Tb 

03 

Bi 

«4 

Po 

85 

At 

66 

Ar 

so 

«*• 

*         1         *         » 

1 

m 


In  earlier  chapters  we  recognized  that  strong 
chemical  similarities  are  displayed  by  elements 
which  are  in  the  same  vertical  column  of  the 
periodic  table.  The  properties  which  chlorine 
holds  in  common  with  the  other  halogens  reflect 
the  similarity  of  the  electronic  structures  of  these 
elements.  On  the  other  hand,  there  is  an  enor- 
mous difference  between  the  behavior  of  ele- 
ments on  the  left  side  of  the  periodic  table  and 
those  on  the  right.  Furthermore,  the  discussions 
in  Chapter  15  revealed  systematic  modification 


in  certain  atomic  properties,  such  as  ionization 
energy,  as  we  proceed  from  left  to  right  along  a 
row  of  the  periodic  table.  Our  purpose  in  this 
chapter  is  to  examine  the  chemical  behavior  of 
the  elements  in  the  third  row  of  the  periodic 
table  to  look  for  trends  in  chemical  properties. 
Specifically,  we  will  consider  the  physical  prop- 
erties of  the  elements  themselves,  their  per- 
formance as  oxidizing  or  reducing  agents,  and 
the  acid-base  behavior  of  their  hydroxides. 


20-1    PHYSICAL  PROPERTIES  OF  THE  ELEMENTS 


All  the  elements  in  Row  3  are  commercially 
available  or  can  be  easily  prepared  in  the  labo- 
ratory. Try  to  examine  as  many  of  these  elements 
as  possible  in  the  laboratory  as  you  study  this 
chapter.  If  all  the  elements  are  available  to  you, 
364 


arrange  them  in  order  of  atomic  number  and 
compare  them.  You  can  hardly  imagine  a  more 
varied  set  of  appearances.  At  one  extreme  we 
have  the  metals  sodium,  magnesium,  and  alumi- 
num. When  they  are  freshly  cut  these  solids  show 


SEC.    20-1    |    PHYSICAL    PROPERTIES    OF    THE    ELEMENTS 


365 


the  bright  luster  or  reflectivity  typical  of  metals. 
They  are  soft:  sodium  is  so  soft  it  can  be  cut 
with  a  knife;  magnesium  and  aluminum  bend  in 
your  fingers  and  can  be  easily  scratched  by  a 
sharp  object.  Silicon  also  shows  the  metallic 
luster,  but  is  much  harder  than  magnesium  or 
aluminum.  Phosphorus  in  the  form  known  as 
white  phosphorus  is  a  yellowish,  waxy  solid  with 
a  distinctly  nonmetallic  appearance.  Black  phos- 
phorus, obtained  by  subjecting  white  phosphorus 
to  high  pressure,  is  a  dark  gray  solid  which  does 
have  some  of  the  luster  which  characterizes 
metals.  By  the  time  we  get  to  sulfur,  however, 
it  is  very  clear  we  have  a  nonmetal.  The  two 
gases  chlorine  and  argon  complete  the  trend 
away  from  metallic  appearance. 

20-1.1    Sodium,  Magnesium,  and  Aluminum: 
Metallic  Solids 

Can  we  explain  the  wide  variation  in  appearance 
and  physical  properties  of  these  elements?  We 
have  already  said  in  Chapter  17  that  metals  are 
found  at  the  left  of  the  periodic  table.  The  low 
ionization  energy  and  vacant  valence  orbitals  of 
one  of  these  elements  lead  to  a  sea  of  highly 
mobile  valence  electrons.  The  mobile  electrons 
hold  the  atoms  together  in  the  metallic  crystal 
and,  at  the  same  time,  are  responsible  for  the 
ease  of  conduction  of  heat  and  electricity.  We 
also  remarked  that  the  metallic  bond  becomes 
stronger  as  the  number  of  valence  electrons  per 
atom  and  their  ionization  energy  increase.  The 
trend  in  the  physical  properties  of  the  third-row 
metals  seems  to  be  well  explained  in  terms  of  the 
increasing  number  and  increasing  ionization  en- 
ergy of  the  valence  electrons. 


Table  20-1 

HEATS    OF    VAPORIZATION    AND 
BOILING    POINTS    OF    METALS 


Na 


Mg 


Al 


A//xap  (kcal/mole) 
b.p.  (°C) 


23.1 
889 


31.5 
1120 


67.9 
2327 


20-1.2    Silicon:  A  Network  Solid 

Even  though  silicon  is  metallic  in  appearance,  it 
is  not  generally  classified  as  a  metal.  The  elec- 
trical conductivity  of  silicon  is  so  much  less  than 
that  of  ordinary  metals  it  is  called  a  semicon- 
ductor. Silicon  is  an  example  of  a  network  solid 
(see  Figure  20-1) — it  has  the  same  atomic  ar- 
rangement that  occurs  in  diamond.  Each  silicon 
atom  is  surrounded  by,  and  covalently  bonded 
to,  four  other  silicon  atoms.  Thus,  the  silicon 
crystal  can  be  regarded  as  one  giant  molecule. 
Almost  all  of  the  valence  electrons  in  the 
silicon  crystal  are  localized  in  the  covalent  bonds 
and  are  not  free  to  conduct  heat  or  electricity  by 
moving  throughout  the  solid.  On  the  other  hand, 
in  the  solid  there  are  always  a  few  valence  elec- 
trons which  have  acquired  enough  energy  to  be 
nonlocalized  and  these  few  electrons  account  for 
the  small,  but  noticeable,  electrical  conductivity 
of  silicon.  Again,  we  can  rationalize  the  behavior 
of  silicon  in  terms  of  its  atomic  structure  and 
ionization  energy.  The  fact  that  the  silicon  atom 
has  four  electrons  (3s23p2)  in  its  valence  orbitals 
accounts  for  its  tendency  to  form  four  covalent 
bonds.  The  increase  in  ionization  energy  and 


Fig.  20-1 .  The  crystal  structures  of  silicon  and 
diamond. 


EXERCISE  20-1 

Write  out  the  electron  configuration  of  sodium, 
magnesium,  and  aluminum  and  find  the  ioniza- 
tion energies  for  all  their  valence  electrons  (Table 
20-IV,  p.  374).  Account  for  the  trend  in  the  heats 
of  vaporization  and  boiling  points  (Table  20-1) 
of  these  elements.  Compare  your  discussion  with 
that  given  in  Section  17-1.3. 


diamond 
C-C  distance  =1.54 A 


silicon 
Si -Si  dirtance  =  2.35 A. 


366 


THE   THIRD    ROW   OF   THE    PERIODIC    TABLE   I   CHAP.    20 


absence  of  vacant  valence  orbitals  as  we  proceed 
along  the  row  accounts  for  the  increasing  locali- 
zation of  the  valence  electrons  into  covalent 
bonds  and  the  almost  complete  disappearance  of 
electrical  conductivity. 


20-1.3    Phosphorus,  Sulfur,  and  Chlorine: 
Molecular  Solids 

Since  the  ionization  energy  of  the  phosphorus 
atom  is  still  higher  than  that  of  the  silicon  atom, 
it  is  not  surprising  that  the  common  forms  of 
phosphorus  are  nonmetallic  molecular  solids. 
White  phosphorus  consists  of  discrete  P4  mole- 
cules (see  Figure  20-2)  and  weak  van  der  Waals 
forces  between  the  separate  molecules  are  re- 
sponsible for  the  stability  of  the  solid.  The 
electronic  structure  of  the  phosphorus  atom 
provides  an  explanation  of  the  formula  and 
structure  of  the  P4  molecule.  Phosphorus  has  the 
electron  configuration  ls22s22/?63.s23/?3,  and  we 
can  suggest  that  since  the  3p  orbitals  are  half- 
filled,  phosphorus  should  be  able  to  form  three 
covalent  bonds.  The  geometry  should  be  like 
that  in  ammonia,  NH3,  in  which  the  three  N — H 
bonds  form  a  pyramid  with  a  three-sided  base 
(see  page  291).  As  shown  in  Figure  20-2,  in  the 
P4  molecule  each  phosphorus  atom  does  make 


Fig.  20-2.  The  structure  of  a  P«  molecule. 


Fig.  20-3.  The  structure  of  an  S.  molecule. 


three  bonds,  and  each  atom  is  at  one  apex  of  a 
pyramid. 

The  electron  configuration  in  the  valence  or- 
bitals of  the  sulfur  atom  (3s23/?4)  suggests  that  it 
will  form  two  covalent  bonds  by  making  use  of 
two  half-filled  7>p  orbitals.  This  is,  in  fact,  ob- 
served in  the  molecule  S8,  which  is  present  in  the 
common  forms  of  solid  sulfur.  The  S8  molecules 
assume  the  form  of  a  puckered  ring,  as  shown  in 
Figure  20-3.  As  with  the  phosphorus,  the  sta- 
bility of  this  crystalline  form  of  sulfur  is  due  to 
van  der  Waals  forces  between  discrete  molecules. 

The  electronic  structure  of  the  chlorine  atom 
(3s23/?5)  provides  a  satisfactory  explanation  of  the 
elemental  form  of  this  substance  also.  The  single 
half-filled  2>p  orbital  can  be  used  to  form  one 
covalent  bond,  and  therefore  chlorine  exists  as 
a  diatomic  molecule.  Finally,  in  the  argon  atom 
all  valence  orbitals  of  low  energy  are  occupied 
by  electrons,  and  the  possibility  for  chemical 
bonding  between  the  atoms  is  lost. 


EXERCISE  20-2 

Using  the  principles  discussed  in  Chapter  17, 
attempt  to  arrange  the  third-row  elements  from 
silicon  through  argon  in  order  of  increasing  boil- 
ing point,  starting  with  the  element  you  think 
has  the  lowest  boiling  point.  Be  prepared  to 
defend  in  a  class  discussion  your  choice  of  the 
position  you  assign  each  element  in  this  se- 
quence. 


SEC.    20-2    I   THE   ELEMENTS    AS   OXIDIZING    AND    REDUCING    AGENTS 


367 


20-2    THE  ELEMENTS  AS  OXIDIZING  AND  REDUCING  AGENTS 


In  the  last  section  your  attention  was  called  to 
the  extreme  variation  in  the  physical  appearance 
and  properties  of  the  third-row  elements.  We 
might  start  this  section  with  a  similar  statement 
about  the  behavior  of  the  elements  as  oxidizing 
and  reducing  agents.  The  outstanding  chemical 
characteristic  common  to  metallic  sodium,  mag- 
nesium, and  aluminum  is  strong  reducing  power. 
Their  tendency  to  lose  electrons  and  to  react  with 
other  elements  is  so  great  that  they  are  found  in 
nature  only  in  compounds,  never  as  free  ele- 
ments. All  three  of  these  metals  will  react  with 
water  to  give  hydrogen.  In  the  case  of  sodium 
this  reaction  is  fast  and  all  of  the  sodium  is  con- 
sumed if  sufficient  water  is  present.  For  mag- 
nesium and  aluminum,  the  reaction  produces  a 
thin  layer  of  oxide  on  the  metallic  surface.  This 
oxide  layer  has  low  solubility  in  water  and  the 
oxide  adheres  strongly  to  the  metal.  Hence  it 
forms  a  protective  layer  that  prevents  further 
contact  between  water  (or  air)  and  the  metal. 
This  protection  accounts  for  the  notable  resist- 
ance of  aluminum  to  weathering,  upon  which 
most  of  the  structural  uses  of  aluminum  depend. 
If  either  magnesium  or  aluminum  comes  in  con- 
tact with  mercury,  the  protective  layer  is  re- 
moved and  rapid  reaction  takes  place. 

As  we  saw  in  Chapter  19,  chlorine  represents 
the  other  extreme  in  chemical  reactivity.  Its  most 
obvious  chemical  characteristic  is  its  ability  to 
acquire  electrons  to  form  negative  chloride  ions, 
and,  in  the  process,  to  oxidize  some  other  sub- 
stance. Since  the  tendency  to  lose  or  gain  elec- 
trons is  a  result  of  the  details  of  the  electronic 
structure  of  the  atom,  let  us  try  to  explain  the 
chemistry  of  the  third-row  elements  on  this  basis. 

20-2.1     Sodium,  Magnesium,  and  Aluminum: 
Strong  Reducing  Agents 

A  glance  at  Appendix  3,  the  table  of  £°'s  for 
half-reactions,  should  convince  you  that  sodium, 
magnesium,  and  aluminum  are  among  the 
strongest  reducing  agents  available.  Their  £°'s 
are  also  listed  in  Table  20-11.  Part  of  this  strong 


Table  20-11 

THE    HALF-CELL    POTENTIALS 
OF    THIRD-ROW    METALS 


NafsJ  — >-  Na+(aq)  +  tr 

E°  =  2.71  volts 

Mg(s)  — >-  Mg^(aq)  +  2e~ 

E°  =  2.37  volts 

A\(s)  — >-  A\+3(aq)  +  3e~ 

E°  =  1.66  volts 

tendency  for  metals  to  lose  electrons  and  become 
positive  ions  in  aqueous  solutions  is  a  result  of 
the  fact  that  the  valence  electrons  of  their  atoms 
are  not  very  strongly  bound,  as  is  shown  by  their 
low  ionization  energies. 

However,  this  is  not  the  complete  explanation  for  their 
reducing  properties.  Let  us  analyze  the  energy  require- 
ments of  the  process 


Nafsj 


Na+(aq)  +  e~ 


U) 


in  much  the  same  way  as  we  treated  the  half-reactions  of 
the  halogens  in  Section  19-2.2.  Reaction  (/)  can  be  dis- 
cussed in  terms  of  a  series  of  hypothetical  steps: 


NafsJ 

Natej 

Na+f|?J 


Nafgj 


vaporization      (2) 


Na+fgJ  +  e~(g)    ionization 


Na+(aqJ 


hydration 


(i) 
(4) 


Step  (2)  is  just  the  vaporization  of  solid  sodium  to  a  gas. 
This  step  requires  energy ;  AH  is  positive  for  this  reaction. 
In  step  (3)  an  electron  is  removed  from  a  gaseous  atom 
to  form  a  gaseous  ion;  AH  is  positive  since  this  step  re- 
quires energy  also.  Step  (4)  is  the  hydration  of  a  positive 
ion;  energy  is  evolved  in  this  process,  so  AH  is  negative. 
Sodium  metal  is  a  good  reducing  agent,  primarily  because 
the  energy  required  to  carry  out  reaction  (/)  is  small.  In 
other  words,  the  energy  we  put  in  to  cause  steps  (2)  and 
(5)  is  small  and  is  somewhat  compensated  by  the  energy 
we  get  out  in  step  (4).  We  can  use  this  example  to  set  up 
general  criteria  for  a  good  metallic  reducing  agent:  (a)  the 
metallic  crystal  must  not  be  too  stable  [otherwise  the 
energy  required  for  step  (2)  will  be  large] ;  (b)  the  ioniza- 
tion energy  of  the  gaseous  atom  should  be  small;  (c)  the 
hydration  energy,  the  energy  evolved  in  step  (4),  should 
be  large. 

We  have  already  mentioned  that  the  stability  of  the 
metallic  crystal  and  the  ionization  energies  of  the  atom 
tend  to  increase  in  the  series  sodium,  magnesium,  and 
aluminum.  In  spite  of  this,  aluminum  is  still  an  excellent 
reducing  agent  because  the  hydration  energy  of  the  Al+I 
ion  is  very  large  (Table  20-111). 


368 


THE    THIRD    ROW    OF    THE    PERIODIC    TABLE    |    CHAP.    20 


Table  20-111 

HYDRATION    ENERGIES    OF    SOME 
THIRD-ROW    IONS    (kcal/mole) 


NaH 
97 


460 


Al+3 
1121 


electrolysis 

MgCW      — ►      Mg(s)  +  a,(g)         (5) 

The  magnesium  metal  is  thus  recovered  for  re- 
peated use  in  reaction  (7).  Chlorine  produced  in 
reaction  (8)  is  also  put  to  use  in  the  manufacture 
of  TiCU,  the  other  reactant  in  reaction  (7). 


The  third-row  metals  also  show  their  strong 
reducing  properties  in  reactions  which  do  not 
take  place  in  aqueous  solution.  For  instance, 
magnesium  metal  ignited  in  air  will  react  with 
carbon  dioxide,  reducing  it  to  elemental  carbon: 
2Mg(s)  +  C02(g)  —>-  2MgO(sj  +  C(s)    (5) 

If  aluminum  metal  is  mixed  with  a  metal  oxide 
such  as  ferric  oxide,  Fe203,  and  ignited,  the  oxide 
is  reduced  and  large  amounts  of  heat  are  evolved: 

2 Alfs;  +  Fe203(s)  — »-  2Fefsj  +  A\203(s) 

kcal 


AH  =  -203 


mole  Fe203 


(6) 


These  reactions  are  possible  because  of  the  great 
stability,  or  low  energy,  of  the  oxides  of  mag- 
nesium and  aluminum.  The  oxides  are  ionic  com- 
pounds whose  great  stability  can  be  attributed 
to  strong  electrostatic  forces  between  small,  posi- 
tively charged  Mg+2  ions  (or  A1+3  ions)  and 
negatively  charged  oxide  ions,  O-2.  The  fact  that 
the  Mg+2  and  Al+3  ions  are  very  small  allows 
these  positive  ions  to  approach  closely  the  nega- 
tive oxide  ions,  resulting  in  strong  attractive 
forces. 

The  ease  of  oxidation  of  magnesium  is  im- 
portant in  the  commercial  manufacture  of  tita- 
nium metal.  Titanium,  when  quite  pure,  shows 
great  promise  as  a  structural  metal,  but  the 
economics  of  production  have  thus  far  inhibited 
its  use.  One  of  the  processes  currently  used,  the 
Kroll  process,  involves  the  reduction  of  liquid 
titanium  tetrachloride  with  molten  metallic 
magnesium: 

TxCUl)  +  2Mg(l)  — >-  Ti(s)  +  2MgCl2r/J      (7) 

Titanium  has  a  very  high  melting  point  (1812°C), 
so  the  magnesium  chloride  can  be  vaporized  and 
distilled  away  from  the  solid  titanium.  The  gase- 
ous magnesium  chloride  is  condensed  and  then 
electrolyzed  to  regenerate  magnesium  and  chlo- 
rine: 


20-2.2    Silicon,  Phosphorus,  and  Sulfur: 
Oxidizing  and  Reducing  Agents 
of  Intermediate  Strengths 

Silicon  also  can  act  as  a  reducing  agent,  as  we 
might  expect  from  the  properties  of  sodium, 
magnesium,  and  aluminum.  It  reacts  with  mo- 
lecular oxygen  to  form  silicon  dioxide,  Si02.  This 
network  solid  is  held  together  by  very  strong 
bonds.  However,  because  of  the  rather  high 
ionization  energy  of  the  silicon  atom,  and  the 
great  stability  of  the  silicon  crystal,  its  reducing 
properties  are  considerably  less  than  those  of  the 
typical  metals. 

Phosphorus  continues  the  trend  away  from 
strong  reducing  properties.  Elemental  phospho- 
rus will  react  with  strong  oxidizing  agents  like 
oxygen  and  the  halogens, 


?<(s)  +  502(g) 
Pt(s)  +  6C\2(g) 


P«O10(sj 
4PC13(!J 


(9) 
(10) 


but  will  also  react  with  strong  reducing  agents 
such  as  magnesium  to  form  phosphides: 

Pt(s)  +  6Mg(s)  — *-  2Mg,P2(s;  (//) 

Therefore,  elemental  phosphorus  shows  both  re- 
ducing and  oxidizing  properties.  This  interme- 
diate behavior  can  be  explained  in  terms  of  the 
electron  occupancy  and  ionization  energy  of  the 
phosphorus  atom.  The  fact  that  the  ionization 
energy  of  phosphorus  is  greater  than  the  previous 
elements  in  Row  3  suggests  it  will  be  a  poorer 
reducing  agent  than,  for  example,  aluminum. 
The  combination  of  a  noticeable  affinity  for  elec- 
trons (as  evidenced  by  the  ionization  energy)  and 
three  half-filled  3p  orbitals  provides  an  explana- 
tion for  the  appearance  of  a  weak  oxidizing  tend- 
ency in  phosphorus.  Since  phosphorus  neither 
loses  nor  gains  electrons  readily,  it  is  neither  a 
strong  reducing  agent  nor  a  strong  oxidizing 
agent. 


SEC.    20-2    I    THE    ELEMENTS    AS    OXIDIZING    AND    REDUCING    AGENTS 


369 


From  this  discussion  you  should  not  conclude 
that  phosphorus  is  an  inert  element.  Many  of  the 
reactions  of  elemental  phosphorus  take  place 
very  rapidly.  For  instance,  when  white  phospho- 
rus is  exposed  to  the  air  it  reacts  rapidly  accord- 
ing to  reaction  (9).  Evidently  this  reaction  has  a 
low  activation  energy,  accounting  for  the  fact 
that  though  phosphorus  is  not  a  strong  reducing 
agent,  yet,  with  oxygen,  it  reacts  rapidly. 

The  ionization  energy  of  the  sulfur  atom  shows 
that  it  is  even  more  reluctant  than  phosphorus 
to  lose  electrons.  The  common  compounds  of 
sulfur  are  the  sulfides,  which  may  be  formed  by 
reactions  of  elemental  sulfur  with  a  large  number 
of  metals.  Typical  reactions  are 

SZn(s)  +  SgfsJ  — *-  SZnS(s)  (12) 

16Ag(sJ  +  Sa(s)  — »-  8Ag2S(s)  (13) 

These  reactions  show  sulfur  in  the  role  of  an 
oxidizing  agent.  The  properties  of  compounds 
such  as  ZnS  suggest  they  contain  the  sulfide  ion, 
S~2.  The  formation  of  this  ion  again  can  be  ex- 
pected on  the  basis  of  the  fact  that  the  neutral 
sulfur  atom  has  two  electrons  less  than  enough 
to  fill  the  valence  orbitals.  Acquisition  of  two 
electrons  completely  fills  the  low  energy  valence 
orbitals  and  solid  ionic  compounds  can  be 
formed. 

Sulfur  reacts  with  molecular  oxygen  to  form 
compounds  in  which  sulfur  is  assigned  positive 
oxidation  numbers,  +4  and  +6.  The  reactions 
are  those  used  in  the  manufacture  of  sulfuric 
acid  (see  Chapter  13): 


\S6(s)  +  02(g) 
S02(g)  +  iO/gj 


SCVgJ 
SQ3(g) 


(14) 
(15) 


Of  these  two  oxides,  S02  (sulfur  dioxide)  is  gase- 
ous at  ordinary  temperatures  and  pressures, 
while  S03  (sulfur  trioxide)  is  a  solid  with  a  rather 
high  vapor  pressure.  Gaseous  sulfur  trioxide  con- 
sists of  discrete  S03  molecules.  Even  though 
sulfur  forms  these  and  other  compounds  in  which 
it  is  assigned  a  positive  oxidation  number,  it  does 
so  only  by  reacting  with  the  strongest  of  oxidiz- 
ing agents.  Therefore  we  see  that  sulfur  is 
somewhat  like  phosphorus — its  strength  as  a 
reducing  agent  or  as  an  oxidizing  agent  is  in- 
termediate. 


EXERCISE  20-3 

Water  saturated  with  S02  gas  is  a  relatively  mild 
but  quite  useful  reducing  agent.  Which  of  the 
following  aqueous  ions  might  be  reduced  by  it? 

(a)  Fe+3  to  Fe+2 

(b)  Cu+2  to  Cu+ 

(c)  Sn+4  to  Sn+2 

(d)  Hg+2  to  Hg(l) 

20-2.3    Chlorine:  A  Strong  Oxidizing  Agent 

The  formation  of  several  oxidation  states  is 
typical  of  the  elements  on  the  right  side  of  the 
periodic  table.  We  have  already  discussed  in 
Chapter  19  the  fact  that  chlorine  can  exist  in  the 
+  1,  +3,  +5,  and  +7  oxidation  states  as  well  as 
in  the  —  1  state.  In  its  compounds,  chlorine  is 
most  often  found  in  the  —  1  state.  This  prepon- 
derance of  —  1  compounds  shows  that  elemental 
chlorine  behaves  as  an  oxidizing  agent  in  most 
of  its  reactions. 

When  we  review  the  oxidation-reduction  prop- 
erties of  the  Row  3  elements  we  do  see  a  rather 
smooth  trend  in  behavior  from  the  strong  reduc- 
ing agent  sodium,  through  the  elements  like 
phosphorus  and  sulfur  which  are  neither  strong 
reducing  agents  nor  strong  oxidizing  agents, 
finally  ending  with  the  strong  oxidizing  agent 
chlorine,  where  the  reducing  tendency  is  very 
low.  This  trend  is  quite  consistent  with  the 
ionization  energies  and  orbital  occupancy  of  the 
atoms.  This  rationalization  of  chemical  behavior 
in  terms  of  electronic  structure  aids  in  remember- 
ing the  chemistry  of  these  elements,  and  in 
making  predictions  of  the  chemistry  of  other 
elements  in  the  table. 

EXERCISE  20-4 

Water  containing  dissolved  Cl2  is  a  useful  oxidiz- 
ing agent.  Which  of  the  following  aqueous  ions 
might  be  oxidized  by  it? 

(a)  Fe+2  to  Fe+3 

(b)  Cu+  to  Cu+2 

(c)  Sn+2  to  Sn+4 

(d)  Mn+2  to  Mn02(s) 

(e)  Mn+2  to  Mn04_ 


370 


THE    THIRD    ROW    OF    THE    PERIODIC    TABLE    I    CHAP.    20 


20-3    THE  ACIDIC  AND  BA6IC  PROPERTIES  OF  THE  HYDROXIDES 


In  the  last  section  we  saw  that  the  variation  in 
chemical  reactivity  of  elements  in  the  third  row 
of  the  periodic  table  can  be  understood  in  terms 
of  electronic  structure  of  the  atoms.  Those  ele- 
ments which  have  both  high  ionization  energies 
and  vacant  valence  orbitals  tend  to  gain  electrons 
and  act  as  oxidizing  agents,  while  those  with 
lower  ionization  energies  tend  to  lose  electrons 
and  act  as  reducing  agents.  We  wish  now  to  ex- 
plain the  acid-base  behavior  of  the  hydroxides 
of  the  third-row  elements,  again  using  the  sim- 
plest possible  ideas  of  atomic  structure.  We 
already  encountered  a  similar  problem  in  Chap- 
ter 19  when  we  explained  the  acidity  of  the 
oxyacids  of  chlorine.  In  this  chapter  we  will 
again  deal  with  a  series  of  compounds  which 
contain  the  group 

M— O— H 

Here  M  might  be  any  of  the  third-row  elements. 
Compounds  of  this  structure  may,  of  course,  act 
as  bases  by  releasing  hydroxide  ions,  breaking 
the  M — OH  bond.  The  hydroxide  ion,  then,  can 
accept  a  proton  from  an  acid  HB 


M— O— H 

OH"  (aq)  +  KB 


M+(aq)  +  OH-(aq)     (16) 
H20  +  B-(aq)  (17) 


Chemists  recognize  another  way  in  which 
M — O — H  compounds  can  act  as  bases.  The 
base  M — O — H  can  react  directly  with  VLB: 

M— O— H  +  HS^± 

M(OH2)+(aq)  +  B-(aq)         (18) 

In  reaction  (18),  one  of  the  unshared  pairs  of 
electrons  on  the  oxygen  atom  of  the  MOH  group 
has  accepted  a  proton,  and  so  MOH  can  act  as 
a  base  without  actually  releasing  hydroxide  ion. 
In  addition,  MOH  compounds  can  act  as 
acids,  breaking  the  MO — H  bond: 

M— O— H  +±:  M-O-(aq)  +  H+(aq)     (19) 
or, 

M— O—  H(s)  +  H20  +±: 

M-O-(aq)  +  HaO+(aq)     (19a) 


We  can  attempt  to  predict  whether  a  com- 
pound containing  the  MOH  group  will  behave 
as  an  acid  or  base  by  considering  the  strength 
with  which  the  element  M  binds  electrons.  If  the 
atom  M  has  a  high  ionization  energy,  it  holds 
electrons  strongly,  so  we  would  not  expect  the 
M — O  bond  to  break  and  form  M+  and  OH- 
ions.  Also,  if  M  holds  electrons  strongly,  we 
expect  it  to  draw  electrons  away  from  the  oxygen 
atom  in  the  MOH  group,  with  the  result  that 
the  oxygen  atom  will  be  less  able  to  acquire  and 
bind  a  proton  as  indicated  in  reaction  (18). 
Therefore,  as  the  ionization  energy  of  M  in- 
creases, we  expect  MOH  to  be  less  able  to  act 
as  a  base,  either  by  reaction  (76)  or  (18). 

An  increase  in  strength  with  which  M  binds 
electrons  has  another  consequence,  as  we  men- 
tioned in  Chapter  19.  As  M  draws  electrons 
toward  itself,  the  O — H  bond  may  become 
weakened  and  the  compound  would  be  expected 
to  display  acidic  properties. 


20-3.1    Sodium  and  Magnesium  Hydroxides: 
Strong  Bases 

Let  us  apply  these  ideas  to  the  third-row  ele- 
ments. On  the  left  side  of  the  table  we  have  the 
metallic  reducing  agents  sodium  and  magnesium, 
which  we  already  know  have  small  affinity  for 
electrons,  since  they  have  low  ionization  energies 
and  are  readily  oxidized.  It  is  not  surprising, 
then,  that  the  hydroxides  of  these  elements, 
NaOH  and  Mg(OH)2,  are  solid  ionic  compounds 
made  up  of  hydroxide  ions  and  metal  ions. 
Sodium  hydroxide  is  very  soluble  in  water  and 
its  solutions  are  alkaline  due  to  the  presence  of 
the  OH-  ion.  Sodium  hydroxide  is  a  strong  base. 
Magnesium  hydroxide,  Mg(OH>2,  is  not  very 
soluble  in  water,  but  it  does  dissolve  in  acid 
solutions  because  of  the  reaction 

Mg(OH)2fS;  +  2H+(aq)  — >- 

Mg+2f  aq)  +  2H20    (20) 

Magnesium  hydroxide  is  a  strong  base. 


SEC.    20-3    |    THE    ACIDIC    AND    BASIC    PROPERTIES    OF    THE    HYDROXIDES 


371 


20-3.2     Aluminum  Hydroxide:  Both  an  Acid 
and  a  Base 

We  can  tell  from  the  ionization  energy  of  alumi- 
num that  this  atom  holds  its  second  and  third 
valence  electrons  rather  firmly.  With  this  fact  in 
mind,  we  can  see  why  aluminum  hydroxide, 
Al(OH)3,  would  not  be  as  strongly  basic  as  are 
the  hydroxides,  NaOH  and  Mg(OH)2.  Alumi- 
num hydroxide  has  extremely  low  solubility  in 
neutral  aqueous  solutions  but  does  react  with 
strong  acids  according  to  the  reaction 

Al(OH)3(s)  +  3H+(aq)  — ►-  Al+3(aq)  +  3H20    (27) 


or 


Al(OH)3f  sj  +  3H30+f  aq) 


A\(OH^3(aq)     (22) 


These  reactions  say  the  same  thing,  but  reaction 
(22)  emphasizes  the  fact  that  the  A1+3  ion  is 
hydrated  in  aqueous  solutions.  In  any  case,  the 
fact  that  Al(OH)3  reacts  with  acids  shows  that 
it  has  the  properties  of  a  base. 

Aluminum  hydroxide  will  also  react  with  hy- 
droxide ion  to  dissolve  according  to  the  equation 

Al(OH)3(sJ  +  OH-(aq)  — »-  Al(OH)4-(aqJ    (23) 

Reaction  (23)  shows  that  Al(OH)3  has  the  prop- 
erties of  an  acid,  since  it  reacts  with  the  base 
OH~.  A  substance  that  acts  as  an  acid  under  some 
conditions  and  as  a  base  under  other  conditions 
is  said  to  be  amphoteric.  The  electronic  situa- 
tion in  Al(OH)3  is  such  that  it  can  either  accept 
a  proton  (act  as  a  base)  or  react  with  OH~  (act 
as  an  acid).  We  will  see  in  Chapter  22  that 
several  other  hydroxides  also  show  amphoteric 
behavior. 


20-3.3    Silicon,  Phosphorus,  Sulfur, 
and  Chlorine  Oxyacids 

Both  our  original  prediction  about  the  effect  of 
ionization  energy  on  acid-base  behavior  and  the 
trend  which  we  have  observed  in  the  first  three 
elements  lead  us  to  expect  that  the  hydroxide  or 
oxide  of  silicon  should  not  be  basic,  but  perhaps 
should  be  weakly  acidic.  This  is  in  fact  observed. 
Silicon  dioxide,  Si02,  can  exist  as  a  hydrated 
solid  containing  variable   amounts   of  water, 


Si02  xH20.  This  hydrated  oxide  does  react  with 
hydroxide  ion  to  form  soluble  silicate  ions: 

Si02xH,0(solid  hydrate)  +  20H  (aq)  — >- 

SiO,->qJ+H,0     (24) 

This  reaction  shows  that  the  hydrated  oxide 
Si02  xH20  is  acidic,  since  it  reacts  with  a  base. 

As  we  mentioned  earlier,  phosphorus  can  be 
found  in  four  different  oxidation  states.  The 
"hydroxides''  of  the  +1,  +3,  and  +5  states  of 
phosphorus  are  hypophosphorous  acid,  H3P02, 
phosphorous  acid,  H3P03,  and  phosphoric  acid, 
H3P04.  Their  structures  are  shown  in  Figure 
20-4.  As  suggested  by  their  names,  these  com- 
pounds are  distinctly  acidic,  and  are  of  moderate 
strength.  The  equilibrium  constant  for  the  first 
ionization  of  each  acid  is  approximately  10-2: 

hypophosphorous  acid: 

H3P02  — >-  H+(aq)  +  H2P02~  (aq) 

K  =  1        X  10"2    (25) 
phosphorous  acid: 

H3POs  -+  H+(aq)  +  H2P03"  (aq) 

K  =  1.6    X  10-2    (26) 
phosphoric  acid: 

H3P04  — ->-  H+(aq)  +  H2P04-  (aq) 

AT  =  0.71  X  10-2     (27) 

Notice  that  this  series  differs  from  the  oxyacids 
of  chlorine,  shown  in  Figure  19-6.  In  the  chlorine 
oxyacids,  oxygen  atoms  add  successively  and  the 
first  ionization  constants  become  larger.  In  the 
phosphorus  oxyacids,  hydrogen  atoms  are  suc- 
cessively replaced  by  O — H  groups  but  the  first 
ionization  constants  change  very  little. 

The  most  common  oxyacid  of  sulfur  is  H2S04, 
sulfuric  acid.  In  dilute  aqueous  solutions  this 
substance  is  almost  completely  dissociated  into 
ions  according  to  the  equation 

H2S04  — -»-  H+(aq)  +  HS04-  (aq)  (28) 

so  sulfuric  acid  is  classified  as  a  strong  acid.  The 
bisulfate  ion,  HS04~  (also  called  the  hydrogen 
sulfate  ion),  is  also  an  acid,  since  the  equilibrium 
constant  for  the  reaction 


HS04" 


H+(aq)  +  SO^(aq)  (29) 


is  approximately  10~2.  From  these  equilibrium 
constants  it  is  clear  that  sulfur  in  the  +6  oxida- 


372 


THE    THIRD    ROW    OF    THE    PERIODIC    TABLE    I    CHAP.    20 


NAME 


FORMULA.  BALL-AND -STICK MODEL       SPACE- FILLING- MODEL 


Hypophosphorous    acid  H3P02 


Phosphorous   acid 


H3P0s 


Phosphoric   acid 


H3PO+ 


Fig.  20-4.  Presumed    structures    of    the    phosphorus 
oxyacids. 


tion  state  is  even  more  acidic  than  silicon  and 
phosphorus. 

Sulfur  in  the  +4  oxidation  state  also  forms 
an  oxyacid,  sulfurous  acid  (H2S03).  This  com- 
pound is  not  as  strong  an  acid  as  H2S04.  The 
equilibrium  constant  for  the  reaction 


H2SO3 


H+(aq)  +  HSOz(aq)         (30) 


is  approximately  10~2.  The  species  HS03~  is 
called  the  bisulfite  ion  or  the  hydrogen  sulfite 
ion.  It  too  is  a  weak  acid  and  dissociates  in  water 
to  form  the  sulfite  ion,  S03-2 

HSO3-  — *■  H+(aq)  +  SO;*(aq)  (31) 

Therefore  we  see  that  the  oxyacids  of  sulfur 
continue  the  trend  of  increasing  acidity  which 


we  have  observed  in  the  successive  third-row 
elements. 

The  oxyacids  (or,  hydroxides)  of  chlorine  were 
discussed  in  Section  19-2.4  and  an  explanation 
of  their  acid  strengths  was  given  there.  The  ar- 
guments used  to  explain  the  increase  in  acid 
strength  of  the  oxyacids  of  chlorine  with  in- 
creasing oxidation  number  connect  with  those 
used  in  this  chapter  to  explain  the  increase  of 
acidity  which  occurs  as  we  successively  consider 
the  oxygen  compounds  of  the  third-row  ele- 
ments. As  a  halogen  atom  adds  oxygen  atoms, 
electron  charge  is  drawn  away  from  the  halogen. 
This,  in  turn  draws  electrons  from  the  O — H 
bond  and  weakens  it.  Acid  strength  increases. 
As  we  move  to  the  right  in  the  periodic  table,  the 
ionization  energy  increases  and,  again,  attrac- 
tion of  electrons  by  the  central  atom  increases. 
This,  too,  draws  electrons  from  the  O — H  bond 
and  weakens  it.  Acid  strength  increases. 


SEC.    20-4    I    OCCURRENCE    AND    PREPARATION    OF    THE    THIRD-ROW    ELEMENTS 


373 


20-4     OCCURRENCE  AND  PREPARATION  OF  THE  THIRD-ROW  ELEMENTS 


Except  for  argon,  the  third-row  elements  make 
up  an  important  fraction  (about  30%)  of  the 
earth's  crust.  Silicon  and  aluminum  are  the  sec- 
ond and  third  most  abundant  elements  (oxygen 
is  the  most  abundant).  Both  the  occurrence  and 
the  mode  of  preparation  of  each  element  can  be 
understood  in  terms  of  trends  in  chemistry  dis- 
cussed earlier  in  this  chapter. 


20-4.1    Occurrence  in  Nature 

Sodium  (fifth  most  abundant  element)  is  found 
principally  as  Na+  ion  in  water  soluble  salt 
deposits,  such  as  NaCl,  and  in  salt  waters.  The 
element  reacts  rapidly  with  water  and  with  at- 
mospheric oxygen,  hence  is  not  found  in  an 
uncombined  state  in  nature. 

Magnesium  (eighth  most  abundant  element)  is 
found  principally  as  Mg+2  ion  in  salt  deposits, 
particularly  as  the  slightly  soluble  carbonate, 
MgC03,  and  also  in  sea  water.  The  element  is 
oxidized  by  atmospheric  oxygen  and  is  not  found 
in  an  uncombined  state  in  nature. 

Aluminum  (third  most  abundant  element)  is 
found  as  the  A1+3  ion  in  oxides  and  as  the  com- 
plex ion  A1F6~3.  Important  minerals  are  bauxite, 
which  is  best  described  as  a  hydrated  aluminum 
oxide,  A1203  xH20,  and  cryolite,  Na3AlF6.  The 
element  is  readily  oxidized  and  is  not  found  in 
an  uncombined  state  in  nature. 

Silicon  is  the  second  most  abundant  element 
in  the  earth's  crust.  It  occurs  in  sand  as  the 
dioxide  Si02  and  as  complex  silicate  derivatives 
arising  from  combinations  of  the  acidic  oxide 
Si02  with  various  basic  oxides  such  as  CaO, 
MgO,  and  K20.  The  clays,  micas,  and  granite, 
which  make  up  most  soils  and  rocks,  are  silicates. 
All  have  low  solubility  in  water  and  they  are 
difficult  to  dissolve,  even  in  strong  acids.  Silicon 
is  not  found  in  the  elemental  state  in  nature. 

Phosphorus  (eleventh  most  abundant  element) 
occurs  mostly  as  the  phosphate  anion,  PCX,-3,  in 
such  minerals  as  "phosphate  rock,"  which  is  a 


complex  mixture  of  Ca3(PO,)2  and  CaF2.  Most 
phosphates  have  low  solubility  in  water.  The 
element  is  not  found  uncombined  in  nature. 

Sulfur  (fourteenth  most  abundant  element)  oc- 
curs in  minerals  either  in  an  oxidized  state  as 
sulfate  anion,  S04~2,  or  in  a  reduced  state  as 
sulfide  anion,  S~2.  Gypsum,  CaS04-2H20,  with 
low  solubility  in  water,  and  Epsom  salt, 
MgS04-7H20  with  high  solubility  in  water,  are 
two  common  sulfate  minerals.  Galena,  PbS,  iron 
pyrites,  FeS2,  and  zinc  blende,  ZnS,  are  impor- 
tant sulfide  minerals.  Sulfur  occurs  as  the  free 
element  in  large  underground  deposits. 

Chlorine  (sixteenth  most  abundant  element)  is 
found  as  Cl~  in  water  soluble  salt  deposits,  such 
as  NaCl,  and  in  salt  waters.  The  element,  Cl2,  is 
not  found  in  the  atmosphere. 

Argon  is  found  only  in  the  elemental  state. 
Air  contains  about  1  %  argon. 


20-4.2    Mode  of  Preparation  of  the  Element 

Sodium  is  prepared  by  electrolysis  of  molten 
NaCl  (giving  chlorine  as  a  by-product)  or  of 
molten  NaOH. 

Magnesium  is  an  important  structural  metal. 
It  can  be  prepared  through  the  sequence  of 
steps:  precipitation  of  Mg+2  from  sea  water  by 
OH~  to  form  Mg(OH)2;  conversion  of  Mg(OH)2 
to  MgCl2;  electrolysis  of  molten  MgCl2. 

Aluminum,  though  the  third  most  abundant 
element,  was  quite  expensive  until  about  1886, 
when  a  practical  commercial  electrolysis  process 
was  developed  by  a  young  American  chemist, 
C.  M.  Hall.  Bauxite,  A1203jcH20,  is  dissolved 
at  about  1000°C  in  molten  cryolite,  Na3AlF6,  and 
electrolyzed. 

Silicon  in  the  elemental  state  has  important 
electronic  applications  as  a  semiconductor  that 
were  developed  only  during  the  last  decade.  The 
discovery  of  these  uses  was  possible  only  after 
methods  were  developed  for  preparing  silicon  of 
extremely  high  purity.  Reduction  of  Si02  with 


374 


THE    THIRD    ROW    OF    THE    PERIODIC    TABLE    ]    CHAP.    20 


carbon  in  an  electric  furnace  is  one  process  for 
manufacture  of  silicon.  Very  pure  silicon  is  made 
by  decomposing  SiCl4.  Still  further  purification 
of  the  element  is  based  upon  the  "zone-melting" 
technique  in  which  a  rod  of  silicon  is  heated  to 
melting  in  a  thin  zone.  This  molten  zone  is 
gradually  moved  along  the  length  of  the  rod. 
The  impurities  dissolve  in  the  liquid  and  move 
along  with  the  zone,  leaving  metal  of  ultra-high 
purity. 

Phosphorus  is  prepared  by  heating  a  mixture 
of  Ca3(P04)2,  sand,  and  carbon  (coke).  White 
phosphorus,  P4,  distills  out  and  can  be  cooled 
and  collected  under  water. 

Sulfur  is  pumped  out  of  natural  underground 
deposits  in  the  molten  state  after  it  is  melted 
with  water  heated  under  pressure  to  about 
170°C. 

Chlorine  is  prepared  by  the  electrolysis  of 
molten  NaCl  or  of  aqueous  NaCl. 


Argon  is  obtained  through  fractional  distilla- 
tion of  liquefied  air. 

20-4.3     Some  Properties  of  the  Second- 
and  Third-Row  Elements 

The  first  ionization  energies  of  elements  1  to  19 
are  shown  in  Table  15-1 II.  The  energies  to  re- 
move successive  electrons  from  gaseous  Na,  Mg, 
and  Al  atoms  are  shown  in  Table  20-IV. 

Trends  in  the  properties  of  AH  of  vaporization 
and  boiling  point  for  the  second-  and  third-row 
elements  are  compared  in  Table  20-V. 

Table  20-IV 

SUCCESSIVE    IONIZATION    ENERGIES 
OF    SODIUM,    MAGNESIUM, 
AND    ALUMINUM    (kcal/mole) 


ELEMENT 


(1st  e~) 


E2 
(2nd  e~) 


Ez 
(3rd  e~) 


Ea 

(4th  e~) 


Na 

118 

1091 

1653 

— 

Mg 

175 

345 

1838 

2526 

Al 

138 

434 

656 

2767 

Table  20-V. 


TRENDS    IN    PROPERTIES    OF    SECOND 

A//Vap  B.P. 

element  (kcal/mole)  (°C) 


QUESTIONS  AND  PROBLEMS 


AND    THIRD-ROW    ELEMENTS 

A//v.p  B.P. 

element  (kcal/mole)  (°C) 


Li 

32.2 

1326 

Na 

23.1 

889 

Be 

53.5 

2970 

Mg 

31.5 

1120 

B 

128.8 

~3900 

Al 

67.9 

2327 

C 

170 

~4000 

Si 

(105) 

2355 

N 

0.67 

-196 

P 

3.0 

280 

O 

0.81 

-183 

S 

2.5 

445 

F 

0.78 

-188 

CI 

4.9 

-34.1 

Ne 

0.42 

-246 

Ar 

1.6 

-186 

Make  a  graph  with  an  energy  scale  extending 
on  the  ordinate  from  zero  to  3000  kcal/mole  and 
with  the  abscissa  marked  at  equal  intervals  with 
the  labels  Na,  Mg,  and  Al.  Now  plot  and  con- 
nect with  a  solid  line  the  first  ionization  energies, 
£w  of  these  three  elements  (see  Table  20-IV). 
Plot  E2  and  connect  with  a  dashed  line,  E3  with 
a  dotted  line,  and  EA  with  a  solid  line.  Draw  a 


circle  around  each  ionization  energy  that  identi- 
fies a  valence  electron. 

2.  Plot  the  ionization  energy  of  the  first  electron 
removed  from  the  atoms  of  both  the  second-  and 
third-row  elements  against  their  atomic  number 
(abscissa).  What  regularity  do  you  observe? 

3.  Silicon  melts  at  1410°C  and  phosphorus  (white) 


QUESTIONS    AND    PROBLEMS 


375 


at  44°C.  Explain  this  very  great  difference  in 
terms  of  the  structures  of  the  solids. 

4.  Recalling  the  chemistry  of  nitrogen,  write  for- 
mulas for  phosphorus  compounds  correspond- 
ing to 

(a)  ammonia, 

(b)  hydrazine, 

(c)  ammonium  iodide. 

5.  Write  the  formula  for  the  fluoride  you  expect  to 
be  most  stable  for  each  of  the  third-row  ele- 
ments. 

6.  The  heat  of  reaction  for  the  formation  of 
MgO(s)  from  the  elements  is  —144  kcal/mole 
of  MgO(s).  How  much  heat  is  liberated  when 
magnesium  reduces  the  carbon  in  C02  to  free 
carbon?  See  Table  7-II. 

Answer.  AH  =  —  97  kcal/mole  MgO. 

7.  Magnesium  oxide  is  an  ionic  solid  that  crystal- 
lizes in  the  sodium  chloride  type  lattice. 

(a)  Explain  why  MgO  is  an  ionic  substance. 

(b)  How  many  calories  would  be  required  to 
decompose  8.06  grams  of  MgO?  (Use  the 
data  in  problem  6.) 

(c)  Draw  a  diagram  of  a  crystal  of  MgO. 

8.  Aluminum  oxide  (A1203)  is  thought  to  dissociate 
at  high  temperature  (1950°C)  according  to  the 
equation:  2Al203(sJ  — >■  4A10(gj  +  02(g).  The 
total  vapor  pressure  at  1950°C  is  about  1  X  10~6 
atm. 

(a)  Which  element  is  oxidized  and  which  is  re- 
duced in  this  reaction? 

(b)  Write  the  equation  for  the  equilibrium  con- 
stant. 

(c)  Calculate  its  value  using  partial  pressure  as 
the  unit  of  "concentration"  for  the  gases. 

9.  Explain  the  observation  that  phosphorus  acts 
both  as  a  weak  reducing  agent  and  as  a  weak 
oxidizing  agent. 

10.  (a)  What  are  the  oxidation  numbers  of  phos- 
phorus in  the  two  compounds  phosphorous 
acid,  H3PO3,  and  phosphoric  acid,  H3P04? 
(b)  From  the  £°  values  in  Appendix  3,  decide 
which  of  the  following  substances  might  be 
reduced  by  phosphorous  acid:  Fe+2;  Sn+4; 
I2;  Cr". 

H20  +  H3P03  — *■  H3PO4  +  2H+  +  2e~ 

E°  =  0.276  volt 


(c)  Balance  the  equation  for  the  reaction  be- 
tween phosphorous  acid  and  Fe+3  and  cal- 
culate E°  for  the  reaction. 

11.  Answer  the  following  in  terms  of  electron  con- 
figuration and  ionization  energy: 

(a)  Which  elements  in  the  second  and  third  rows 
are  strong 

(i)  oxidizing  agents? 
(ii)  reducing  agents? 

(b)  What  properties  do  strong  oxidizing  agents 
have? 

(c)  What  properties  do  strong  reducing  agents 
have? 

12.  Of  the  elements  Na,  Mg,  Al,  which  one  would 
you  expect  to  be  most  likely  to 

(a)  form  a  molecular  solid  with  chlorine? 

(b)  form  an  ionic  solid  with  chlorine? 

13.  One  kilogram  of  sea  water  contains  0.052  mole 
of  magnesium  ion.  What  is  the  minimum  num- 
ber of  kilograms  of  sea  water  that  would  have 
to  be  processed  in  order  to  obtain  1  kg  of 
Mg(OH)2? 

Answer.  3.3  X  102  kg  of  sea  water. 

14.  Why  is  aluminum  hydroxide  classed  as  an  am- 
photeric compound? 

15.  Some  of  the  following  common  compounds  of 
the  third-row  elements  are  named  as  hydroxides 
and  some  as  acids : 

NaOH         sodium  hydroxide 

Mg(OH)2    magnesium  hydroxide 

Al(OH)3      aluminum  hydroxide 

Si(OH)4       silicic  acid  (usually  written  H4Si04) 

P(OH)3        phosphorous  acid  (usually   written 

H3P03) 
S(OH)2        not  known 
Cl(OH)       hypochlorous  acid  (usually  written 

HOC1) 

(a)  Explain  why  these  compounds  vary  system- 
atically in  their  acid-base  behavior. 

(b)  Write  equations  that  show  the  reactions  of 
each  of  these  substances  either  as  acids,  as 
bases,  or  both. 

16.  A  solution  containing  0.20  M  H3P03,  phospho- 
rous acid,  is  tested  with  indicators  and  the 
H+(aq)  concentration  is  found  to  be  5.0  X 
10-2  M.  Calculate  the  dissociation  constant  of 
H3P03,  assuming  that  a  second  proton  cannot 
be  removed. 


376 


THE    THIRD    ROW    OF    THE    PERIODIC    TABLE    I    CHAP.    20 


17.  Elemental  phosphorus  is  prepared  by  the  reduc- 
tion of  calcium  phosphate,  Ca3(P04)j,  with  coke 
in  the  presence  of  sand,  Si02.  The  products  are 
phosphorus,  calcium  silicate,  CaSi03,  and  car- 
bon monoxide. 


(a)  Write  the  equation  for  the  reaction. 

(b)  Using  75.0  kg  of  the  ore,  calcium  phosphate, 
calculate  how  many  grams  of  P4  can  be  ob- 
tained and  how  many  grams  of  coke  (as- 
sumed to  be  pure  carbon)  will  be  used. 


CHAPTER 


21 


The  Second  Column 
of  the  Periodic  Table: 
The  Alkaline  Earths 


^i 

Us 

z 
He 

u 
Ha 

4 
Bt 

Mo 

s 
B 

6 

c 

7 

N 

8 

o 

9 

F 

JO 

Ne 

13 
At 

14 

Si 

15 
P 

16 
S 

17 

CI 

18 
Ar 

19 

X 

20 

Ca 

21 
Sc 

.22 

n 

23 
V 

24 
Cr 

25 
Mn 

26 

Fe 

27 
Co 

28 

Ni 

29 
Cu 

30 
Zn 

31 
Ga 

32 
Ge 

33 
As 

34* 
Se 

35 
Br 

36 
Kr 

Rb 

SB 

Sr 

39 

Y 

40 

Zr 

Nb 

42 

Mo 

43 
Tc 

44 
Ru 

45- 

Rh 

46 

Vd 

47 

48 
Cd 

49 
In 

so 
Sn 

SI 

sb 

52 

Te 

S3 
I 

54 

Xe 

ss 
Cs 

T6 

Sa 

17-71 

72 

Hf 

73 

Ta 

74 

w 

75 

Re 

76 
■09 

77 

Ir 

76 

Ft 

79 
Au 

80 

Ho 

61 

Tt 

62 
Fb 

83 

Si 

34 
Po 

85 

At 

86 
Rn 

67 

Ft 

66 
Ra 

69- 

iii' 

i       i       i       » 

In  the  preceding  chapter  we  looked  at  the  ele- 
ments of  the  third  row  in  the  periodic  table  to 
see  what  systematic  changes  occur  in  properties 
when  electrons  are  added  to  the  outer  orbitals 
of  the  atom.  We  saw  that  there  was  a  decided 
trend  from  metallic  behavior  to  nonmetallic, 
from  base-forming  to  acid-forming,  from  simple 
ionic  compounds  to  simple  molecular  com- 
pounds. These  trends  are  conveniently  discussed 


in  terms  of  the  ionization  energies  and  orbital 
occupancies. 

There  are  similar,  but  smaller,  trends  in  the 
properties  of  elements  in  a  column  (a  family)  of 
the  periodic  table.  Though  the  elements  in  a 
family  display  similar  chemistry,  there  are  im- 
portant and  interesting  differences  as  well.  Many 
of  these  differences  are  explainable  in  terms  of 
atomic  size. 


21-1     ELECTRON  CONFIGURATION  OF  THE  ALKALINE  EARTH  ELEMENTS 


The  elements  of  the  second  column  and  their 
electron  configurations  are  given  in  Table  21-1. 
For  each  element,  the  neutral  atom  has  two  more 


electrons  than  an  inert  gas.  We  can  expect  these 
two  electrons  to  be  easily  removed,  to  give  the 
stability  of  the  inert  gas  electron  configuration. 

377 


378      THE  SECOND  COLUMN  OF  THE  PERIODIC  TABLE:  THE   ALKALINE  EARTHS    I    CHAP.    21 


Table  21-1.     electron  configurations  of  the  alkaline  earth  elements 

ELECTRON  ARRANGEMENT 


ELEMENT 

SYMBOL 

NUCLEAR 
CHARGE 

INNER  LEVELS 

OUTER  LEVELS 

beryllium 

Be 

4 

1j» 

2* 

magnesium 

Mg 

12 

\s* 

2s*2pt 

3* 

calcium 

Ca 

20 

•••2jj2/7< 

3s23p« 

45* 

strontium 

Sr 

38 

•••3s*3/>»        3rf10 

4s*4p> 

5s2 

barium 

Ba 

56 

•••4^4/j«        4dl° 

5s*5(fi 

6s1 

radium 

Ra 

88 

■  ■  ■tpWSpfi        5d10 

6526/7« 

7s* 

EXERCISE  21-1 

On  the  basis  of  the  electron  configurations  and 
positions  in  the  periodic  table,  answer  the  follow- 
ing questions. 

(a)  Is  calcium  likely  to  be  a  metal  or  nonmetal  ? 

(b)  Is  calcium  likely  to  resemble  magnesium  or 
potassium  in  its  chemistry? 

(c)  Is  calcium  likely  to  have  a  higher  or  a  lower 
boiling  point  than  potassium?  than  scan- 
dium? 

EXERCISE  21-2 

Predict  the  chemical  formula  and  physical  state 
at  room  temperature  of  the  most  stable  com- 
pound formed  by  each  alkaline  earth  element 
with  (a)  chlorine;  (b)  oxygen;  (c)  sulfur. 


Exercises  21-1  and  21-2  pose  some  of  the 
simplest  questions  we  can  ask  about  the  alkaline 
earths.  The  periodic  table  arranges  in  a  column 
elements  having  similar  electron  configurations. 
We  can  expect  elements  on  the  left  side  of  the 
periodic  table  to  be  metals  (as  magnesium  is). 
Furthermore,  we  can  expect  that  the  elements  in 
a  given  column  will  be  more  like  each  other  than 
they  will  be  like  elements  in  adjacent  columns. 
Thus,  when  we  find  that  the  chemistry  of  mag- 
nesium is  almost  wholly  connected  with  the 
behavior  of  the  dipositive  magnesium  ion,  Mg+2, 
we  can  expect  a  similar  situation  for  calcium, 
and  for  strontium,  and  for  each  of  the  other 
alkaline  earth  elements.  This  proves  to  be  so. 

Remembering,  then,  that  the  alkaline  earths 
are  classed  as  a  family  because  of  general  simi- 
larity, we  shall  investigate  the  detailed  differences 
among  them. 


21-2    TRENDS  IN  PHYSICAL  PROPERTIES 


21-2.1    Atomic  Radii  in  Solids 

The  size  of  an  atom  is  defined  in  terms  of  the 
interatomic  distances  that  are  found  in  solids 
and  in  gaseous  molecules  containing  that  atom. 
For  an  atom  on  the  left  side  of  the  periodic  table, 
gaseous  molecules  are  obtained  only  at  very  high 
temperatures.  At  normal  temperatures,  solids 
are  found  and  there  are  two  important  types  to 
consider,  metallic  solids  and  ionic  solids.  Table 
21-11  shows  the  nearest  neighbor  distances  in  the 


pure  metals,  in  the  gaseous  oxide  molecules,  and 
in  the  solid  oxides  (which,  except  for  BeO,  have 
the  sodium  chloride  crystal  structure  pictured  in 
Figure  5-10,  p.  81).  This  table  also  shows  the 
corresponding  radii  assigned  to  each  alkaline 
earth  atom,  first,  in  the  metallic  state,  second, 
in  the  gaseous  molecule  (assuming  that  oxygen 
has  the  same  size  as  in  02)  and  third,  in  the  state 
of  a  -f-2  ion  (assuming  that  the  oxide  ion,  O-2, 
should  be  assigned  a  radius  of  1.32  A). 


SEC.    21-2    I    TRENDS    IN    PHYSICAL    PROPERTIES 


379 


Table  21-11.     trends  in  interatomic  distances 

NEAREST    NEIGHBOR    DISTANCE 

(Angstroms) 


ALKALINE   EARTH    ATOMIC   SIZE 


IN  THE  PURE 

IN  THE  GASEOUS 

IN  THE  SOLID 

METALLIC 

DOUBLE-BOND 

IONIC 

ELEMENT 

METAL 

OXIDE 

OXIDE 

RADIUS 

RADIUS 

RADIUS 

Be 

2.23 

1.33 

1.64 

1.11 

0.73 

0.32 

Mg 

3.20 

1.75 

2.10 

1.60 

1.15 

0.78 

Ca 

3.95 

1.82 

2.40 

1.97 

1.22 

1.08 

Sr 

4.30 

1.92 

2.57 

2.15 

1.32 

1.25 

Ba 

4.35 

1.94 

2.76 

2.17 

1.34 

1.44 

We  see  that,  no  matter  what  type  of  bonding 
situation  is  considered,  there  is  a  trend  in  size 
moving  downward  in  the  periodic  table.  The 
alkaline  earth  atoms  become  larger  in  the 
sequence  Be  <  Mg  <  Ca  <  Sr  <  Ba.  These 
atomic  sizes  provide  a  basis  for  explaining  trends 
in  many  properties  of  the  alkaline  earth  elements 
and  their  compounds. 

21-2.2    Ionization  Energies 

Table  21 -III  shows  the  first  three  ionization  en- 
ergies of  the  alkaline  earths. 


Table  21-111 

ionization  energies  of  the 
alkaline  earth  elements 

ionization  energy  (kcal/mole) 
element        Et  (1st  e~)        £2  (2nd  e~)       E3  (3rd  e~) 


Be 

214 

420 

3533 

Mg 

175 

345 

1838 

Ca 

140 

274 

1173 

Sr 

132 

253 

986 

Be 

120 

230 

811 

EXERCISE  21-4 

If  the  ionization  energy  £\  is  regarded  as  a 
measure  of  the  distance  between  the  electron 
and  the  nuclear  charge,  what  do  the  ionization 
energies  of  Be  and  Ba  indicate  about  the  relative 
sizes  of  the  two  atoms? 


From  Exercise  2 1  -4  we  see  that  the  decreasing 
ionization  energies  observed  for  the  alkaline 
earth  atoms  are  readily  explained  in  terms  of 
their  increasing  size  moving  down  in  the  periodic 
table.  Notice  that  the  ionization  energy  trend 
going  down  in  the  periodic  table  is  the  same  as 
the  trend  going  to  the  left  in  the  periodic  table. 


EXERCISE  21-5 

From  the  ionization  energies,  predict  which 
solid  substance  involves  bonds  having  the  most 
ionic  character:  BeCl2,  MgCl2,  CaCl2,  SrCl2, 
BaCl2.  Which  substance  is  expected  to  have  most 
covalent  character  in  its  bonds? 


EXERCISE  21-3 


For  each  of  the  alkaline  earths,  calculate  the 
ratio  Ei/E\.  Account  for  the  results  in  terms  of 
the  charges  on  the  ions  formed  in  the  two  ioniza- 
tion steps. 


21-2.3    Metallic  Properties 

Table  21 -IV  shows  some  properties  of  the  metals 
and  their  crystal  forms.  Since  different  crystal 
forms  are  involved  in  the  series,  trends  in  the 
properties  are  obscured.  Figure  21-2  shows  scale 
representations  of  the  crystal  structures  of  metal- 
lic beryllium,  calcium,  and  barium. 


380      THE  SECOND   COLUMN  OF  THE   PERIODIC   TABLE:   THE   ALKALINE  EARTHS    |    CHAP.    21 


Metal,     M 


Gaseous  oxide,     Jvf-O 


Be 


M9 


Ca 


Sr 


Ba 


' 


Metallic    radius 


-     -  ,   - 


Double   bond  radius 


1.08 


Ionic  radius 


Fig.  21-1.  Sizes  of  the  alkaline  earth  atoms  with  various  bond  types  (in  Angstroms). 


SEC.    21-3    ;    TRENDS    IN    CHEMICAL    PROPERTIES 


381 


Be 
Hexagonal 


Ca 
Face--  centered  cubic 


Ba 
Body-cen-t-ered    cubic 


Fig.  21-2.  Scale  representations  of  the  crystal  struc- 
tures of  Be,  Ca,  and  Ba. 


Table  21 -IV.     properties  of  the  alkaline  earths  in  the  metallic  state 


CRYSTAL 
STRUCTURE 


DENSITY 

(g  ml) 


MELTING 

HEAT  OF 

ELECTRICAL 

POINT 

VAPORIZATION 

CONDUCTIVITY 

(°C) 

(kcal/mole) 

(ohm-cm)-1 

Be 

hexagonal 

1.85 

1283 

54 

1.69  X  10s 

Mg 

hexagonal 

1.75 

650 

32 

2.24  X  10s 

Ca 

face-centered  cubic 

1.55 

850 

42 

2.92  X  10* 

Sr 

face-centered  cubic 

2.6 

770 

39 

0.43  X  10" 

Ba 

body-centered  cubic 

3.5 

710 

42 

0.16  X  10s 

EXERCISE  21-6 


the  metallic  state.  Compare  the  trend  in  these 
From  the  density  of  each  element,  calculate  the  molar  volumes  with  the  trend  in  the  metallic 
volume  occupied  by  one  mole  of  its  atoms  in     radii  shown  in  Table  21-11. 


21-3    TRENDS  IN  CHEMICAL  PROPERTIES 


We  have  already  observed  (in  Exercise  21-2)  that 
the  alkaline  earths  have  similar  chemistry.  As 
shown  in  Table  21-1,  they  have  similar  electron 
configurations.  Table  21 -III  shows  that  each  ele- 
ment has  two  valence  electrons.  With  these  basic 
likenesses  in  mind  we  shall  explore  the  chemical 
trends  among  these  elements. 


21-3.1    Oxidation  and  Reduction 

All  of  the  alkaline  earths  are  strong  reducing 


agents,  since  they  readily  release  electrons.  The 
values  of  E°  are  collected  in  Table  21-V. 

Table  21-V 

THE  HALF-CELL  POTENTIALS 
FOR  THE  ALKALINE  EARTHS 


Be 

Mg 
Ca 
Sr 
Ba 


Be+2  +  2e- 
Mg+2  +  2e- 
Ca+2  +  2e- 
Sr+2  +  2e- 
Bh+2    +  2c- 


E°  =  +1.85  volts 

E°  -  +2.37 

E°  =  +2.87 

E°  =  +2.89 

E>  =  +2.90 


382     THE  SECOND  COLUMN  OF  THE  PERIODIC  TABLE:   THE  ALKALINE  EARTHS   I   CHAP.    21 


EXERCISE  21-7 

The  ease  of  removal  of  an  electron  from  a  gase- 
ous atom,  the  ionization  energy,  is  one  of  the 
factors  that  is  important  in  fixing  E°.  Refer  back 
to  Table  21-111  and  predict  the  trend  in  E°  that 
this  factor  would  tend  to  cause. 


21-3.2    Acid  and  Base  Properties 

We  have  explained  the  trends  in  acid-base  char- 
acter across  the  periodic  table  by  considering 
the  increasing  ionization  energy  of  the  metal 
atom.  As  the  atom  M  in  a  structure  M — O — H 
attracts  electrons  more  and  more  strongly,  there 
is  increasing  tendency  toward  acidic  properties. 
As  the  ionization  energy  of  M  goes  down,  there 
is  increasing  tendency  toward  basic  properties. 

Moving  down  in  a  column  is  equivalent  in 
many  respects  to  moving  to  the  left  in  the  peri- 
odic table.  Since  we  find  basic  properties  pre- 
dominant at  the  left  of  the  periodic  table  in  a 
row,  we  can  expect  to  find  basic  properties 
increasing  toward  the  bottom  of  a  column.  Thus 
the  base  strength  of  the  alkaline  earth  hydroxides 
is  expected  to  be  largest  for  barium  and  stron- 
tium. The  greatest  acid  strength  is  expected  for 
beryllium  hydroxide. 

Experimentally  we  find  that  strontium  and 
barium  hydroxides  are  indeed  strong  bases.  All 
of  the  alkaline  earth  hydroxides  dissolve  readily 
in  acidic  solutions,  showing  that  they  are  all 
bases  to  some  extent: 


CaO(s)  +  H20(l) 


Ca(OH)2f8J 

A7/  =  -15.6  kcal    (7) 


Be(OH)2(sJ  +2H+(aq) 
Mg(OH)2(sJ  +  2H+(aq) 
Ca(OH)2fsJ  +  2H+(aq) 
Sr(OH)2(sJ  +2H+(aq) 
Ba(OH)2fsJ  +2H+(aq) 


Be«  +2H20  (7) 

Mg+2  +  2H20  (2) 

Ca4*  +2H20  (3) 

Sr^    +2H20  (4) 

Ba+2  +2H2Q  (5) 


Only  beryllium  hydroxide  dissolves  appreciably 
in  strong  base  solutions, 


Be(OH)2(s)  +  20H- 


BeO-2  +  2H20    (5) 


These  hydroxides  are  formed  from  the  cor- 
responding oxides.  For  example,  calcium  oxide, 
or  lime,  reacts  with  water  as  in  reaction  (7). 


The  process  is  called  "slaking"  the  lime  and  it  is 
used  by  plasterers  in  preparing  mortar,  which 
requires  Ca(OH)2.  As  water  is  added  to  lime 
there  is  a  considerable  evolution  of  heat,  as 
evidenced  by  wisps  of  steam  that  rise  from  the 
sample. 

Because  all  of  the  alkaline  earth  oxides  react 
with  water  to  form  basic  hydroxides,  they  are 
called  basic  oxides.  The  reactions  and  their 
heats  are  as  follows: 


BeO(s)  +  H20(l) 
MgO(s)  +  H20(Zj 
CaOfsJ  +  H20(l) 
StO(s)  +  H20(%) 
BaO(s)  +  HiOd) 


Be(OH)2(s) 

AH  =     -2.5  kcal      (8) 

Mg(OH)/sJ 

AH  =    -8.9  kcal      (9) 

Ca(OH)2(sJ 

AH  =  -15.6  kcal    (70) 

Sr(OH)2(sJ 

AH  =  -19.9  kcal    (77) 

Ba(OH)./sJ 

AH  =  -24.5  kcal    (72) 


Notice  the  progressively  increasing  exothermic 
reaction  heat,  moving  downward  in  the  series. 


EXERCISE  21-8 

How  much  heat  is  evolved  if  one  pound  (454 
grams)  of  lime  is  slaked  according  to  reaction 
(70)?  How  many  grams  of  water  can  be  evapo- 
rated with  this  heat?  (The  heat  of  vaporization 
of  water  is  about  10  kcal/mole.) 


21-3.3    Solubilities  of  Alkaline  Earth 
Compounds  in  Water 

We  encountered  the  solubilities  of  alkaline  earth 
salts  in  Chapter  10  and  discovered  some  interest- 
ing trends.  Before  looking  back  to  Figures  10-5 
and  10-6,  see  how  much  you  can  recall  about 
these  solubilities. 


SEC.    21-3    I    TRENDS    IN    CHEMICAL    PROPERTIES 


383 


EXERCISE  21-9 

In   your   notebook   indicate   one   of  the   four 
answers 

(i)  none  of  the  alkaline  earth  ions; 
(ii)  all  alkaline  earth  ions; 
'iii)  Be+2,  Mg+2,  and  Ca+2,  but  not  Sr+2,  Ba^2, 

or  Ra+2; 
(iv)  Sr+2,  Ba+2,  and  Ra+2,  but  not  Be+2,  Mg+2, 
or  Ca+2 

for  each  of  the  following: 

(a)  ||  form  compounds  of  low  solubilities 
with  C1-,  Br~,  and  I". 

(b)  IHfl  torni  compounds  of  low  solubilities 
with  sulfate,  S04~2. 

(c)  ||  form  compounds  of  low  solubilities 
with  sulfide,  S-2 

(d)  ||  form  compounds  of  low  solubilities 
with  hydroxide,  OH~. 

(e)  ||  form  compounds  of  low  solubilities 
with  carbonate,  C03-2. 

Now  compare  your  answer  with  Figures  10-5 
and  10-6. 


Table  21 -VI 


THE    SOLUBILITY    PRODUCTS    OF    THE 
ALKALINE    EARTH    HYDROXIDES 


COMPOUND 

Ktp 

Mg(OH)2 

8.9  X  10-" 

Ca(OH)j 

1.3  X  10-6 

Sr(OH)2 

3.2  X  10-* 

Ba(OH)2 

5.0  X  10~« 

few  ions  in  solution;  the  larger  values  correspond 
to  higher  concentration  in  a  saturated  solution — 
that  is,  higher  solubility. 


EXERCISE  21-10 

Suppose  you  have  a  solution  in  which  the  con- 
centration of  hydroxide  ion  is  1  M.  How  many 
moles  per  liter  of  the  different  alkaline  earth  ions 
listed  in  Table  21 -VI  could  you  have  (at  equilib- 
rium) in  this  solution?  If  the  concentration  of 
hydroxide  ions  were  0.5  M,  how  would  your 
answers  change? 


THE    HYDROXIDES 

When  a  hydroxide  such  as  calcium  hydroxide  is 
added  to  water  in  sufficient  amount,  we  get  a 
saturated  solution  containing  Ca+2  and  OH~  in 
equilibrium  with  excess  undissolved  solid.  The 
equilibrium 

Ca(OH)/S/)  ^±  Ca+Yaqj  +  20H~( aq)    (13) 

can  also  be  established  by  mixing  Ca+2  ions  (for 
example,  from  a  solution  of  CaCl2)  with  OH~ 
ions  (for  example,  from  a  solution  of  NaOH) 
until  a  precipitate  of  Ca(OH)2  forms.  In  either 
case,  at  equilibrium  the  concentrations  of  Ca+2 
and  OH-  ions  are  such  that  the  equilibrium 
expression  is  satisfied: 


[Ca«][OH-p  =  K„ 


(14) 


In  Table  21 -VI  the  numerical  values  of 
[M+2][OH~]2  are  listed  for  some  of  the  alkaline 
earth  hydroxides.  Small  values  indicate  relatively 


Exercise  21-10  demonstrates  that  there  is  a 
regular  trend  in  the  solubilities  of  the  alkaline 
earth  hydroxides. 

THE    CARBONATES    AND    SULFATES 

Although  the  hydroxides  of  the  alkaline  earth 
elements  become  more  soluble  in  water  as  we  go 
down  the  column,  the  opposite  trend  is  observed 
in  the  solubilities  of  the  sulfates  and  carbonates. 
For  example,  Table  21 -VI I  shows  the  solubility 
products  of  the  alkaline  earth  sulfates. 


Table  21 -VI I 

THE    SOLUBILITY    PRODUCTS    OF    THE 
ALKALINE    EARTH    SULFATES 


COMPOUND 


Kn 


MgSO« 

soluble  (K.p  »  10"1) 

CaSCX 

2.4  X  10-6 

SrSO« 

7.6  X  10-7 

BaSO* 

1.5  X  10-» 

384     THE  SECOND  COLUMN  OF  THE  PERIODIC  TABLE:   THE  ALKALINE  EARTHS    |    CHAP.    21 


The  solubility  of  calcium  carbonate  is  such 
that  in  a  saturated  solution  the  product  of 
ion  concentrations  [Ca+2]  [C03_ 2]  is  5  X  lO"9. 
Though  this  may  seem  quite  small,  it  is  large 
enough  to  be  important  to  man,  especially  if  he 
lives  in  a  region  of  the  earth  where  there  are 
extensive  limestone  deposits.  Calcium  carbonate 
can  be  dissolved  in  water,  especially  if  it  contains 
much  dissolved  C02.  This  is  objectionable  be- 
cause soap  added  to  water  which  contains  even 
traces  of  Ca+2  forms  a  precipitate  of  calcium 
stearate.  This  is  the  ring  that  is  so  difficult  to 
remove  from  the  bathtub. 

The  dissolving  of  limestone  by  ground  water 
is  another  example  of  chemical  equilibrium.  The 
behavior  of  this  system  depends  upon  the  chemi- 
cal equilibrium  between  CaC03  and  its  dissolved 
ions  and  the  equilibrium  between  carbonate  ion 
and  dissolved  C02  in  the  water.  When  CaC03 
dissolves  in  water  it  establishes  the  equilibrium 

CaC03(s)  +±  Ca+2(aq)  +  C03"  2(aq)     (15) 

The  carbonate  ion,  a  base,  can  accept  a  proton 
from  water,  an  acid, 

CO3-  2(aq)  +  H.O(l)  +± 

HC03  (aq)  +  OH(  aq)     (16) 

Thus,  solutions  of  carbonates  are  found  to  be 
basic.  Aqueous  solutions  of  carbon  dioxide  are, 
on  the  other  hand,  acidic.  The  reactions  in  this 
equilibrium  are 

C02(g)  +  H20(l)  =?=*:  H«C03(aq)  (17) 

H«C03(aq)  z^±  HCO3-  (aq)  +  H+(aq)     (18) 

The  combination  of  reaction  (18)  and  (16)  shows 
how  carbon  dioxide  enhances  the  solubility  of 
calcium  carbonate  by  removing  carbonate  ion  to 
form  bicarbonate  ion, 


H,C03(aq)  +  CO3- 2(aq)  z<±t  2HC03  (aq)     (19) 
or,  in  the  net  reaction, 

CaCCVsJ  +  CO,(g)  +  H2Of/J  q=b 

Ol+2(aq)  +  2HC03-  (aq)     (20) 

The  result  is  that  we  get  an  appreciable  concen- 
tration of  Ca+2  in  the  water,  giving  so-called 
hard  water — hard  on  the  soap  and  hard  on  the 
people  who  use  it.  Caves  in  limestone  regions  are 
formed  essentially  by  the  combination  of  the  two 
equilibria  above.  In  contrast,  the  weird  icicle-like 
projections  (stalactites)  found  hanging  from  the 
roofs  of  such  caves  are  formed  by  the  reverse  of 
these  reactions.  On  standing,  a  droplet  of  a  satu- 
rated solution  containing  Ca+2  and  HC03~  may 
lose  some  C02  and  H20  by  evaporation.  Loss  of 
C02  and  H20  from  the  equilibrium  (20)  en- 
hances the  reverse-directed  change,  resulting  in 
the  deposit  of  a  fleck  of  CaC03.  The  same  change 
occurs  when  hard  water  of  this  kind  is  boiled  in 
a  pot  or  heated  in  a  boiler.  The  white  scum  you 
may  see  forming  on  the  surface  of  boiling  water 
is  often  due  to  these  equilibria. 

Fig.   21-3.   Stalactites:   solubility   equilibria   at   work. 

Limestone  '■    CaCOj 
CaC03(s)  *  C02  +  H20  -*  Ca+2  +  2HC03~ 


;       HZ0(f) 

CaZ  *  ZHCO;^  Co. C03 (s)  +  C02  +H20 


21-4     OCCURRENCE  AND  PREPARATION  OF  THE  ALKALINE 
EARTH  ELEMENTS 


As  we  did  in  the  preceding  chapter,  we  conclude 
by  summarizing  some  information  about  the 
occurrence  in  nature  and  the  modes  of  preparing 
the  alkaline  earth  elements. 


21-4.1    Occurrence  in  Nature 

All  of  the  alkaline  earth  elements  exist  in  nature 
as  the  M+2  cations. 
Beryllium  (forty-fourth  most  abundant  ele- 


SEC.    21-4    I    OCCURRENCE   AND    PREPARATION   OF   THE   ALKALINE   EARTH    ELEMENTS       385 


ment)  is  rather  rare  and  occurs  mostly  as 
an  aluminum  beryllium  silicate  called  beryl, 
Be3Al2Si6Oi8.  Beryl  containing  traces  of  chro- 
mium has  a  beautiful  green  color  and  is  called 
emerald. 

Magnesium  (eighth  most  abundant  element)  is 
found  principally  as  Mg+2  ion  in  salt  deposits, 
particularly  as  the  slightly  soluble  carbonate, 
MgC03,  and  also  in  sea  water.  The  natural  de- 
posits of  MgC03  with  CaC03  are  called  dolo- 
mite. Magnesium  is  present  as  a  cation  in  the 
asbestos  silicates. 

Calcium  (sixth  most  abundant  element)  is 
found  in  limestone,  CaC03,  and  gypsum, 
CaS04-2H20.  Bones  are  made  up  of  calcium 
phosphate,  Ca3(P04)2. 

Strontium  (thirty-eighth  most  abundant  cle- 
ment) is  rather  rare  and  is  found  principally  as 
the  mineral  strontianite,  SrC03. 

Barium  (eighteenth  most  abundant  element) 
is  also  rather  rare;  it  occurs  as  the  mineral 
barite,  BaS04. 

Radium  is  radioactive  and  extremely  rare.  It 
occurs  in  trace  amounts  (one  part  in  1012)  in 
uranium  ores  such  as  pitchblende  (mainly  U308). 


EXERCISE  21-11 


What  property  held  in  common  by  the  following 
compounds  accounts  for  their  presence  in  natu- 
ral mineral  deposits:  MgC03,  CaC03,  SrC03, 
BaS04,  and  (in  bones)  Ca3(P04)2? 


EXERCISE  21-12 


What  property  held  in  common  by  the  alkaline 
earth  elements  accounts  for  the  fact  that  the  free 
elements  are  not  found  in  nature? 


21-4.2     Mode  of  Preparation  of  the  Element 

Only  magnesium  is  produced  in  any  substantial 
quantity  in  the  elemental  form.  The  reaction 
sequence  used  is  given  in  Section  20-4.2. 

The  general  method  for  preparing  the  alkaline 
earth  elements  is  to  convert  the  mineral  to  a 
chloride  or  a  fluoride  by  treatment  with  HC1  or 
HF.  Then  the  molten  salt  is  electrolyzed  or,  as 
in  the  case  of  BeF2,  reduced  with  a  chemical 
reducing  agent  such  as  Mg. 


Alfred  E.  Stock  was  one  of  the  greatest  inorganic  chemists 
of  the  twentieth  century.  His  research  investigations,  which 
resulted  in  over  250  publications,  were  characterized  by 
brilliant  experimental  technique  and  convincing  thorough- 
ness. Both  of  these  rich  qualities  were  needed  for  his 
hazardous  studies  of  the  hydrides  oj  boron,  an  overlooked 
area  of  chemistry  in  which  he  was  recognized  as  the  single 
world  authority  for  a  period  of  at  least  a  decade.  It  is  fitting 
that  his  name  is  perpetuated  in  the  "Stock  system"  of  in- 
organic chemical  nomenclature  (in  which  Roman  numerals 
indicate  oxidation  numbers). 

Stock  was  born  in  Danzig,  Poland,  and  his  aptitude  for 
science  was  displayed  early  in  his  boyhood  collections  of 
salamanders,  butterflies,  and  plants.  He  studied  at  the 
University  of  Berlin  where  the  chemistry  facilities  of  the 
day  were  so  limited  that  this  brilliant  experimentalist-to-be 
had  to  wait  till  his  third  semester  to  approach  a  laboratory 
bench.  He  received  the  Ph.D.  at  the  University  of  Berlin 
in  1899,  graduating  magna  cum  laude. 

Shortly  after  1900,  young  Alfred  Stock  began  his  life- 
time work:  study  of  the  chemistry  of  boron.  He  reasoned 
that  this  neighbor  of  the  versatile  carbon  atom  could  not 
possibly  have  the  dull  and  limited  chemistry  popularly  as- 
sumed at  the  time.  He  entered  this  study  stimulated  by  his 
own  desire  to  know — despite  advice  by  the  laboratory  di- 


rector to  select  another  area  because  the  chemistry  of 
boron  was  already  thoroughly  investigated.  His  persistence 
was  rewarded  by  discoveries  of  a  succession  of  hydrides 
of  boron,  such  as  diborane,  B^Hs,  tetraborane,  BaH\0, 
pentaborane,  BhH*,  and  decaborane,  BwHu.  The  structures 
and  even  the  very  existence  of  these  compounds  baffled 
chemists  for  many  years.  Even  to  the  date  oj  Stock's  death, 
theoreticians  had  no  convincing  explanation  of  the  absence 
of  the  prototype  molecule,  BH3,  and  their  discussions  of  the 
nature  of  the  bonding  in  diborane  were  based  upon  an 
assumed  structure  that  was  later  shown  to  be  incorrect. 
Stock's  amazing  exploratory  study  went  far  beyond  the 
expectations  and  predictions  of  other  inorganic  chemists 
of  his  day.  This  work,  that  culminated  in  his  book,  Hydrides 
of  Boron  and  Silicon,  presaged  the  rapidly  opening  field 
of"unusuar  inorganic  chemistry  now  so  actively  pursued. 
Alfred  E.  Stock  was  always  zealous  in  recognizing  the 
contributions  and  help  of  his  coworkers  and  subordinates 
during  a  period  in  which  this  was  an  uncommon  virtue.  He 
was  not  only  an  outstanding  scientist,  but  also  a  considerate 
and  thought  Jul  human  being  as  revealed  by  his  comment: 
"The  most  important  problem  for  the  scientific  mind  to 
solve  will  be  how  to  free  mankind  from  political,  social, 
and  economic  limitations  and  how  to  give  it  a  purer, 
broader-minded  understanding  of  humanity.  .  .  ." 


CHAPTER 


The  Fourth-Row 
Transition  Elements 


/ 
H 

2 

He 

3 

Li 

4 
Be 

5 
B 

6 

c 

7 

N 

8 
0 

9 

F 

10 
Ne 

n 

12 

Mo 

15 
At 

14 

Si 

15 
P 

16 

s 

17 

CI 

18 
Ar 

Transition.    elements                  ^ 

19 

K 

20 

Ca 

21 

Sc 

22 
Ti 

23 
V 

24 
Cr 

25 
Mn 

26 

Fe 

27 
Co 

28\  29 
Ni  1  C  u 

30 
Zn 

31 
Go. 

32 
Ge 

33 
As 

34 

Se 

35 
Br 

36 
Kr 

37 

Rb 

38 
Sr 

39 
Y 

40 
Zr 

41 

Nb 

42 

Mo 

45 

Tc 

44 

Ru 

45 

Rh 

4b 
Pd 

47 
A9 

48 

Cd 

49 
In 

so 
Sn 

51 

5b 

52 

Te 

53 
I 

54 

Xe 

55 
Cs 

56 
Ba 

57-71 

72 

Hf 

73 

Ta 

74 

W 

75 

Re 

7b 

Os 

77 

76 
Ft 

79 

Au 

80 

Ify 

81 

Tt 

82 
Pb 

33 
Bt 

04 
P0 

85 

At 

56 
Rn 

87 
Fr 

88 
Ra 

89- 

■  > 

■  •        • 

In  the  preceding  chapters  we  have  studied  the 
chemistry  of  the  elements  across  the  top  of  the 
periodic  table  and  down  the  two  sides.  Now  we 
shall  consider  the  elements  in  the  middle.  These 
are  usually  referred  to  as  the  transition  elements 
because  chemists  once  believed  that  some  ele- 
ments behaved  in  a  way  intermediate  between 
the  extremes  represented  by  the  left  and  right 


sides  of  the  periodic  table.  Today,  the  term 
"transition  element"  remains  mostly  a  useful 
way  of  designating  elements  in  this  particular 
region  of  the  periodic  table,  even  though  we  can- 
not pinpoint  a  specific  set  of  properties  and  say 
that  all  the  transition  elements  have  all  these 
properties. 


22-1    DEFINITION  OF  TRANSITION  ELEMENTS 


There  is  some  disagreement  among  chemists  as 
to  just  which  elements  should  be  called  transi- 
tion elements.  For  our  purposes,  it  will  be  con- 
venient to  include  all  the  elements  in  the  columns 
of  the  periodic  table  headed  by  scandium 
through  zinc. 

Across  the  top,  as  the  first  row  of  the  transi- 


tion region,  we  have  the  elements  scandium  (Sc), 
titanium  (Ti),  vanadium  (V),  chromium  (Cr), 
manganese  (Mn),  iron  (Fe),  cobalt  (Co),  nickel 
(Ni),  copper  (Cu),  and  zinc  (Zn).  On  the  left,  we 
have  the  scandium  column  which  includes,  be- 
sides Sc,  yttrium  (Y,  39),  lanthanum  (La,  57), 
and  actinium  (Ac,  89).  For  reasons  that  we  shall 

387 


Arbitrary 

erterqy 

scale. 


bs 


6p 

ooo  oqS©g  oogS 


sP 


5s 


ooo 


4  t> 

ooo 


3s 


2s 


Is 


3P 


2p 


4d 


18 


18 


Z    * 


t'ii;.  22-1 .  The  fourth-row  transition  elements  in  the  enersv  level  diagram. 


SEC.    22-1    I    DEFINITION    OF    TRANSITION    ELEMENTS 


389 


take  up  in  the  next  chapter,  we  group  with 
lanthanum  the  fourteen  elements  that  follow  La 
(Z  =  58  through  Z  =  71);  these  we  call  the 
lanthanide  elements.  On  the  right,  the  transition 
elements  end  with  the  zinc  column.  Besides  zinc, 
this  includes  cadmium  (Cd,  48)  and  mercury 
(Hg,  80).  It  is  strongly  advisable  during  the  dis- 
cussion that  follows  to  look  back  at  the  periodic 


Co. 


Is 
Is2 


2s 


Ss 


09     ®g®       <g> 


2s' 


V 


Js2 


table  frequently  to  see  where  each  particular 
element  is  placed. 

22-1.1     Electron  Configuration 

There  are  two  immediate  questions  we  ask  about 
the  transition  elements  once  we  know  where  they 
are  in  the  periodic  table:  (1)  Why  do  we  consider 
these  elements  together?  (2)  What  is  special 
about  their  properties?  These  questions  are 
closely  related  because  they  both  depend  upon 
the  electron  configurations  of  the  atoms.  What, 
then,  is  the  electron  configuration  we  might  ex- 
pect for  these  elements? 

To  answer  this  question,  we  need  to  review 
some  basic  ideas  on  the  electronic  buildup  of 
atoms.  We  saw  in  Chapter  15  that  as  we  progres- 
sively add  electrons  to  build  up  an  atom,  each 
added  electron  goes  into  the  lowest  energy  level 


that  is  not  already  fully  occupied.  With  this  prin- 
ciple as  a  guide,  let  us  consider  the  electron 
configurations  as  we  build  up  the  first  row  of 
transition  elements  from  scandium  through  zinc. 
Looking  at  the  periodic  table,  we  see  that  calcium 
comes  just  before  scandium.  The  twenty  elec- 
trons in  a  calcium  atom  are  distributed  as  shown 
in  the  following  arrangement: 


J>  4s 


+sz 


3d 


OOOOOOCO  o 


V 


In  element  number  21,  we  must  accommodate 
one  more  electron.  At  first  sight  we  might  predict 
that  the  21st  electron  goes  into  the  4p  orbital,  as 
the  next  higher  energy  level  after  45.  The  4p 
orbital  is  of  higher  energy  than  the  45  but,  more 
important,  there  is  a  set  of  five  3d  orbitals  in 
between.  The  21st  electron  goes  into  a  3d  orbital 
as  the  level  of  next  higher  energy.  This  is  shown 
in  Figure  22-1  (which  is  just  Figure  15-11  re- 
produced here  for  convenient  reference). 


EXERCISE  22-1 

Draw  on  one  line  a  set  of  orbitals  from  Is 
through  Ad.  Under  this  give  the  orbital  occu- 
pancy for  Al,  Sc,  and  Y.  Account  for  the  fact 
that  yttrium  is  much  more  like  scandium  than  is 
aluminum. 


Table  22-1. 


THE    ELECTRON    CONF 

GURAT 

ONS    OF    ' 

ELEM ENTS 

ATOMIC 

ELEMENT 

SYMBOL 

NUMBER,   Z 

scandium 

Sc 

21 

titanium 

Ti 

22 

vanadium 

V 

23 

chromium 

Cr 

24 

manganese 

Mn 

25 

iron 

Fe 

26 

cobalt 

Co 

27 

nickel 

Ni 

28 

copper 

Cu 

29 

zinc 

Zn 

30 

ROW  TRANSITION 


ELECTRON  CONFIGURATION 


l*2      2s2  2pf>      3s1  3pfi 

3d' 4s1 

3d*  4s2 

3d3  4s* 

3d*  4s1 

Each  fourth-row  transi- 

3d* 4s2 

tion  element  has  these 

3d*  4s* 

levels  filled. 

3d1  4s1 

3</Mj2 

3dv>4s* 

3du>4s* 

390 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    I    CHAP.    22 


There  are  five  3d  orbitals  available,  all  more 
or  less  of  the  same  energy.  Putting  a  pair  of  elec- 
trons in  each  of  these  five  orbitals  means  that  a 
total  of  ten  electrons  can  be  accommodated 
before  we  need  to  go  to  a  higher  energy  level. 
Not  only  scandium  but  the  nine  following  ele- 
ments can  be  built  up  by  adding  electrons  into 
3d  orbitals.  Not  until  we  get  to  gallium  (element 
number  31)  do  we  go  up  to  another  set  of 
orbitals. 

EXERCISE  22-2 

Again  using  Figure  22-1,  decide  which  orbital 
would  next  be  used  after  the  five  3d  orbitals  have 
been  filled.  What  orbital  would  next  be  used 
after  the  Ad  set  has  been  filled?  What  element 
does  this  correspond  to  in  the  periodic  table? 

At  this  stage,  with  the  help  of  Figure  22-1,  or 
with  an  atomic  orbital  chart,  you  should  be  able 
to  work  out  the  electronic  configuration  of  most 
of  the  transition  elements.  You  will  not  be  able 
to  deduce  them  all  correctly  because  there  are 
some  exceptions  resulting  from  special  stabilities 
when  a  set  of  orbitals  is  filled  or  half-filled.  The 
fourth-row  transition  elements  have  the  set  of 
electron  configurations  shown  in  Table  22-1. 
Notice  that  chromium  (Z  =  24)  and  copper 
(Z  =  29)  provide  interruptions  to  the  continuous 
buildup.  In  the  case  of  chromium  the  whole  atom 
has  lower  energy  if  one  of  the  4s  electrons  moves 
into  the  3d  set  to  give  a  half-filled  set  of  3d 
orbitals  and  a  half-filled  4s  orbital;  in  the  case 
of  copper,  the  atom  has  lower  energy  if  the  3d 
set  is  completely  populated  by  ten  electrons  and 
the  4s  orbital  is  half-filled,  instead  of  having  nine 
3d  electrons  and  two  in  the  4*  orbital. 

EXERCISE  22-3 

Make  an  electron  configuration  table  like  Table 
22-1  for  the  fifth-row  transition  elements — 
yttrium  (Z  =  39)  through  cadmium  (Z  =  48). 
In  elements  41  through  45,  one  of  the  5s  electrons 
moves  over  to  a  4d  orbital.  In  element  46,  two 
electrons  do  this. 


In  the  sixth-row  transition  elements  (lan- 
thanum through  mercury)  there  is  an  additional 
complication.  There  are  seven  4/  orbitals  which 
are  very  close  in  energy  to  the  5d  orbitals.  Put- 
ting electrons  into  these  4/  orbitals  means  there 
will  be  fourteen  additional  elements  in  this  row. 
These  fourteen  elements  are  almost  identical  in 
many  chemical  properties.  We  will  discuss  them 
in  the  next  chapter. 

22-1.2    General  Properties 

What  properties  do  we  actually  find  for  the 
transition  elements?  What  kinds  of  compounds 
do  they  form?  How  can  the  properties  be  inter- 
preted in  terms  of  the  electron  populations  of 
the  atoms? 

Looking  at  a  sample  of  each  transition  element 
in  the  fourth  row,  we  see  that  they  are  all 
metallic.  When  clean,  they  are  shiny  and  lus- 
trous. They  are  good  conductors  of  electricity 
and  also  of  heat :  some  of  them  (copper,  silver, 
gold)  are  quite  outstanding  in  these  respects.  One 
of  them  (mercury)  is  ordinarily  a  liquid;  all 
others  are  solids  at  room  temperature. 

So  far  as  chemical  reactivity  is  concerned,  we 
find  a  tremendous  range.  Some  of  the  transition 
elements  are  extremely  unreactive.  For  example, 
gold  and  platinum  can  be  exposed  to  air  or  water 
for  ages  without  any  change.  Others,  such  as 
iron,  can  be  polished  so  they  are  brightly  metallic 
for  a  while,  but  on  exposure  to  air  and  water  they 
slowly  corrode.  Still  others  are  vigorously  re- 
active and,  when  exposed  to  air,  produce  a 
shower  of  sparks.  Lanthanum  and  cerium,  for 
instance,  especially  when  finely  divided,  oxidize 
immediately  when  exposed  to  air.  (Some  ciga- 
rette lighters  have  flints  containing  these  metals.) 

It  is  hard  to  generalize  about  the  chemical 
reactivities  of  a  group  of  elements  since  reactivi- 
ties depend  upon  two  factors:  (A)  the  relative 
stability  of  the  specific  compounds  formed  com- 
pared with  the  reactants  used  up,  and,  (B)  the 
rate  at  which  the  reaction  occurs.  In  special 
cases  there  are  other  complications.  For  exam- 
ple, chromium  metal  (familiar  in  the  form  of 
chrome  plate)  is  highly  reactive  toward  oxygen. 
Still,  a  highly  polished  piece  of  chromium  holds 


SEC.    22-1    I    DEFINITION    OF    TRANSITION    ELEMENTS 


391 


its  luster  almost  indefinitely  when  exposed  to  air. 
The  explanation  is  that  a  very  thin,  invisible  coat 
of  oxide  quickly  forms  on  the  surface  and  pro- 
tects the  underlying  metal  from  contact  with  the 
oxygen  in  the  air.  In  other  words,  bulk  chromium 
is  unstable  with  respect  to  oxidation  by  air,  but 
the  protective  layer  of  oxide  cuts  the  rate  of 
conversion  so  much  that  no  reaction  is  observed. 
What  about  compounds  of  the  transition  ele- 
ments? Suppose  we  go  into  the  chemical  stock- 
room and  see  what  kinds  of  compounds  are  on 
the  shelf  for  a  particular  element,  say  chromium. 
First  we  might  find  a  bottle  of  green  powder 
labeled  Cr203,  chromic  oxide,  or  chromium(III) 
oxide.  Next  to  it  there  would  probably  be  a 
bottle  of  a  reddish  powder,  Cr03,  chromium(VI) 
oxide.  On  an  amply  stocked  chemical  shelf,  we 
might  also  find  some  black  powder  marked  CrO, 
chromous  oxide,  or  chromium(II)  oxide.  There 
would  probably  also  be  some  other  simple  com- 
pounds such  as  CrCl3,  chromic  chloride,  or 
chromium(III)  chloride,  a  flaky,  reddish-violet 
solid,  and  maybe  some  green  CrF2,  chromous 
fluoride  or  chromium(II)  fluoride.  Elsewhere  in 
the  stockroom  we  would  run  across  K2Cr04,  a 
bright  yellow  powder  (potassium  chromate), 
probably  next  to  a  bottle  of  orange  potassium 
dichromate,  K2Cr207.  Soon  we  would  get  the 
idea  that  the  compounds  of  chromium,  at  least 
the  common  ones,  correspond  to  oxidation  num- 
ber of  +2  (CrO  and  CrF2),  +3  (Cr2Os  and 
CrCl3),  and  +6  (Cr03,  K2Cr04,  K2Cr207). 


EXERCISE  22-4 


What  is  the  oxidation  number  of  chromium  in 
each  of  the  following  compounds:  Cr207-2, 
Cr042,  Cr(OH)3,  Cr02Cl2? 


Along  with  these  simple  compounds,  we  might 
also  find  some  rather  more  complex  substances. 
For  example,  we  might  find  next  to  CrCl3 
vials  of  several  brightly  colored  solids  labeled 
CrCl3-6NH3,  CrCl3  5NH3,  CrCl3  4NH3,  and 
CrCl3-3NH3.  Recalling  that  the  dot  in  these  for- 
mulas simply  indicates  that  a  certain  number  of 
moles  of  NH3  are  bound  to  one  mole  of  CrCl3, 
we  would  conclude  that  here  also  the  oxidation 
number  of  chromium  is  +3.  Looking  further, 
we  might  find  other  complex  compounds  such 
as  K3CrF6,  Na3Cr(CN)6,  KCr(S04)2  12H20.  In 
all  these  the  chromium  has  a  +3  oxidation  num- 
ber. As  a  result  of  our  stockroom  search,  we 
would  form  three  conclusions:  (1)  chromium 
forms  simple  and  complex  compounds;  (2)  chro- 
mium forms  a  number  of  stable  solids,  most  of 
them  colored;  (3)  chromium  may  have  different 
oxidation  numbers,  including  +2,  +3,  and  +6. 
Similar  conclusions  would  have  resulted  for  most 
of  the  other  transition  elements. 

Is  there  any  regularity  to  the  kind  of  com- 
pounds the  fourth-row  transition  elements  form? 
Table  22-11  shows  what  chemists  have  found. 


Table  22-11. 


SYMBOL 


TYPICAL    OXIDATION    NUMBERS    FOUND    FOR    FOURTH-ROW 
TRANSITION    ELEMENTS 


REPRESENTATIVE  COMPOUNDS 


COMMON  OXIDATION  NUMBERS 

(most  common  in  bold  type) 


NUMBER  OF  VALENCE 
3d,  4s  ELECTRONS 


Sc 

Sc,03 

+3 

3 

Ti 

TiO, 

Ti203. 

TiO* 

+  2,  +3,  +4 

4 

V 

vo, 

v2o3, 

vo2,    v2o= 

+  2,  +3,  +4, 

+5 

5 

Cr 

CrO, 

Cr503, 

CrO, 

+  2,  +3 

+6 

6 

Mn 

MnO 

Mn:0., 

Mn02,            KjMnCX,  KMnQ4 

+2,  +3,  +4 

+6,  +7 

7 

Fe 

FeO. 

Fe>03 

+  2.  +3 

8 

Co 

CoO, 

Co,03 

+2,  +3 

9 

Ni 

NiO. 

Ni.Q3 

+2,  +3 

10 

Cu 

CU;0.   CllO 

+1,  42 

11 

Zn 

ZnO 

+2 

12 

392 


THE    FOURTH-ROW    TRANSITION    ELEMENTS   I    CHAP.    22 


EXERCISE  22-6 


Look  through  a  handbook  of  chemistry  and  find 
one  other  compound  of  each  oxidation  state 
given  for  the  elements  in  Table  22-11. 


Several  points  should  be  noted  from  this  table: 

(1)  For  most  of  the  transition  elements,  several 
oxidation  numbers  are  possible. 

(2)  When  several  oxidation  numbers  are  found 
for  the  same  element,  they  often  differ  from 
each  other  by  jumps  of  one  unit.  For  exam- 
ple, in  the  case  of  vanadium  the  common 
oxidation  numbers  form  a  continuous  series 
from  +2  to  +3  to  +4  to  +5.  Compare  this 
with  the  halogens  (Chapter  19).  In  the  case 
of  chlorine,  for  example,  the  common  states 
are  —1,  +1,  +3,  +5,  and  +7  (jumps  of 
two  units  instead  of  one  unit). 

(3)  The  maximum  oxidation  state  observed  for 
the  elements  first  increases  and  then  de- 
creases as  we  go  across  the  transition  row. 
Thus  we  have  +3  for  scandium,  +4  for 
titanium,  -f-5  for  vanadium,  +6  for  chro- 
mium, and  +7  for  manganese.  The  +7  rep- 
resents the  highest  value  observed  for  this 
transition  row.  After  manganese,  the  maxi- 
mum value  diminishes  as  we  continue  toward 
the  end  of  the  transition  row. 


What  explanation  can  we  give  for  these  ob- 
servations? Why  does  the  combining  capacity 
vary  from  one  transition  element  to  another  in 
such  a  way  that  the  above  pattern  of  oxidation 
numbers  develops?  The  combining  capacity  of 
an  atom  depends  upon  how  many  electrons  the 
atom  uses  for  bonding  to  other  atoms.  The 
unique  feature  of  the  transition  elements  is  that 
they  have  several  electrons  in  the  outermost  d 
and  s  orbitals,  and  the  ionization  energies  of  all 
of  these  electrons  are  relatively  low.  Therefore 
it  is  possible  for  an  element  like  vanadium  to 
form  a  series  of  compounds  in  which  from  two 
to  five  of  its  electrons  are  either  lost  to,  or  shared 
with,  other  elements.  Consider,  for  example,  the 
oxides  VO  and  V203,  which  contain  the  V+2  and 
the  V+3  ions,  respectively.  While  more  energy  is 
needed  to  form  V+3  than  to  form  V+2,  the  V+3 
has,  because  of  its  higher  charge,  a  greater  attrac- 
tion for  the  O-2  ion  than  does  V+2.  This  extra 
attraction  in  V203  compensates  for  the  energy 
needed  to  form  the  V+3  ion,  and  both  oxides  (as 
well  as  V02  and  V205)  are  stable  compounds. 
Notice,  moreover,  that  the  maximum  oxidation 
number  of  the  transition  elements  never  exceeds 
the  total  number  of  s  and  d  valence  electrons. 
The  higher  oxidation  states  become  increasingly 
more  difficult  to  form  as  we  proceed  along  a  row, 
because  the  ionization  energy  of  the  d  and  5  elec- 
trons increases  with  the  atomic  number. 


22-2    COMPLEX  IONS 


The  remaining  general  point  to  be  made  about 
the  transition  elements  is  that  they  form  a  great 
variety  of  complex  ions  in  which  other  molecules 
or  ions  are  bonded  to  the  central  transition  ele- 
ment ion  to  form  more  complex  units.  These  are 
called  complex  ions.  Take  the  series  already  men- 
tioned: CrCl3-6NH3,  CrCl3-5NH3,  CrCl3-4NH3, 
and  CrCl3-3NH3.  How  can  we  account  for  the 
existence  of  such  a  series?  The  answer  can  be 
seen  if  we  consider  some  of  the  observed  facts 
about  these  complex  compounds.  For  example, 
if  we  dissolve  one  mole  of  each  in  water  and  add 


a  solution  of  silver  nitrate  in  an  attempt  to 
precipitate  the  chloride  as  AgCl, 

Ag+  +  CI"  zf±:  AgClfsJ  (2) 

we  find  that  sometimes  much  of  the  chloride 
cannot  be  precipitated.  The  observed  results  are: 


Moles  of  CI 

Moles  of  Cl~ 

Compound 

precipitated 
3  of  3 

not  precipitated 

CrCl3-6NH3 

0 

CrCl3-5NH3 

2  of  3 

1  of  3 

CrCl3-4NH3 

1  of  3 

2  of  3 

CrCl3-3NH3 

0 

3  of  3 

SEC.    22-2    I   COMPLEX    IONS 


393 


Evidently,  there  are  two  ways  in  which  chlorine 
is  bound  in  these  compounds,  one  way  which 
allows  the  Cl~  to  be  precipitated  by  Ag+  and 
another  way  which  does  not.  In  CrCl3-6NH3, 
all  of  the  chloride  can  be  precipitated;  in 
CrCl3  3NH3,  none  of  it  can  be.  Other  data  also 
indicate  different  types  of  bonding.  For  instance, 
the  freezing  point  lowering  of  an  aqueous  solu- 
tion of  CrCl3  6NH3  indicates  there  are  four 
moles  of  particles  per  mole  of  CrCl3-6NH3;  the 
solution  is  highly  conducting.*  On  the  other 
hand,. the  freezing  point  lowering  of  CrCl3-3NH3 
solution  indicates  there  is  one  mole  of  particles 
per  mole  of  CrCl3-3NH3;  furthermore,  the  solu- 
tion does  not  conduct  at  all.  The  explanation  of 
this  behavior  was  provided  in  the  early  1900's 
by  Alfred  Werner,  who  noted  that  complex  com- 
pounds of  chromium  +3  could  be  accounted 
for  by  assuming  each  chromium  is  bonded  to 
six  near  neighbors.  In  CrCl3-6NH3,  the  cation 
consists  of  a  central  Cr+3  surrounded  by  6  NH3 
molecules  at  the  corners  of  an  octahedron; 
the  three  chlorine  atoms  exist  as  anions,  CI-. 
In  CrCl3-5NH3,  the  cation  consists  of  the 
central  chromium  surrounded  by  the  five  NH3 
and  one  of  the  CI  atoms;  the  other  two  CI 
atoms  are  anions.  In  CrCl3-4NH3,  the  chro- 
mium Is  bound  to  four  NH3  and  two  CI  leaving 
one  chloride  anion.  In  CrCl3  3NH3,  all  three 
CI  atoms  and  all  three  NH3  molecules  are 
tied  to  the  central  chromium.  The  formulas 
can  be  written 

[Cr(NH3)6]Cl3 
[Cr(NH3)5Cl]Cl2 
[Cr(NH3)4Cl2]Cl 
[Cr(NH3)3Cl3] 


22-2.1    Geometry  off  Complex  Ions 

The  way  that  atoms  or  molecules  are  arranged 
in  space  around  a  central  atom  has  a  great  in- 
fluence on  whether  a  given  complex  aggregate 
is  stable  enough  to  be  observed.  What  kinds  of 
arrangements  are  found  in  complex  ions?  What 

*  From  work  on  simple  salts  such  as  NaCl  we  expect 
that  the  "particles"  are  ions,  and  the  conductivity  con- 
firms this. 


shapes  do  these  complex  ions  show?  Can  we  find 
any  regularity  in  the  transition  elements  that  will 
enable  us  to  predict  what  complex  ions  will 
form? 

First,  let  us  introduce  a  concept  useful  in 
giving  spatial  descriptions;  the  coordination 
number  is  the  number  of  near  neighbors  that 
an  atom  has.  For  example,  in  the  complex  ion 
A1F6""3  (which  is  the  anion  present  in  the  solid 
mineral  cryolite),  each  Al  atom  is  surrounded  by 
six  fluorine  atoms  at  the  corners  of  an  octa- 
hedron, as  shown  in  Figure  22-2.  We  say  that 
aluminum  has  a  coordination  number  of  6  with 
fluorine.  In  the  complex  ion  AlBr4~,  which  seems 
to  be  an  important  intermediate  when  aluminum 
bromide  acts  as  a  catalyst  for  many  organic  re- 
actions, the  bromine  atoms  are  arranged  around 
a  central  Al  at  the  corners  of  a  regular  tetra- 
hedron. Figure  22-3  shows  the  arrangement.  The 
coordination  number  of  aluminum  is  4  with 
bromine. 

If  more  than  simple  atoms  are  bound  to  a 
central  atom,  then  the  coordination  number  still 
refers  to  the  number  of  near  neighbors.  For 
example,  in  solid  potassium  chrome  alum, 
KCr(S04)2-  12H20,  and  also  in  its  fresh  aqueous 
solutions,   the  chromium-containing  cation   is 


Fig.  22-2.  An    octahedral    complex:    aluminum    with 
coordination  number  6. 


Ion  with 
3  charge. 


394 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    I    CHAP.    22 


Jon    -with. 
-1    charge 


Fig.  22-3.  A  tetrahedral  complex:  aluminum  with  co- 
ordination number  4. 


Cr(H20)^3.  It  consists  of  a  central  chromium 
joined  to  six  H20  molecules,  exactly  as  the  fluo- 
rines are  arranged  around  aluminum  in  Figure 
22-2.  The  oxygen  portion  of  each  H20  molecule 
is  turned  toward  the  central  chromium  and  the 
H  portions  point  away  from  the  center.  The 
corners  of  the  cage  that  holds  the  central  Cr 
atom  are  occupied  by  six  oxygen  atoms,  each  of 


Ion  with   +1  charge 


which  also  holds  two  H  atoms.  The  shape  of 
this  complex  ion  is  octahedral  and  we  say  that 
in  Cr(H20)t+3  chromium  shows  a  coordination 
number  of  6  to  oxygen. 

Notice  that  in  an  octahedral  complex  ion  such 
as  [Cr(NH3).iCl2]+  there  is  a  possibility  of  ob- 
serving isomers.  The  two  chlorine  atoms  may 
occupy  octahedral  positions  which  are  next  to 
each  other  on  the  same  side  of  the  metal  atom, 
or  positions  located  on  opposite  sides  of  the 
metal  atom  (see  Figure  22-4).  The  isomer  in 
which  the  two  similar  groups  are  located  on  the 
same  side  of  the  metal  atom  is  called  the  cis- 
isomer,  and  the  other  is  called  the  /ra/is-isomer. 

The  complex  ion  Fe(C204)3~3  is  formed  when 
rust  stains  are  bleached  out  with  oxalic  acid 
solution.  It  also  has  a  transition  element  showing 
coordination  number  of  6,  even  though  there  are 
only  three  groups  (QAf2  groups)  around  each 
iron  ion.  Figure  22-5  shows  the  arrangement. 
Each  C204"2,  the  oxalate  group,  uses  two  of  its 
oxygen  atoms  to  bind  onto  the  central  iron  atom. 
The  number  of  near  neighbors,  as  viewed  from 
the  iron  atom,  is  six  oxygen  atoms  at  the  corners 
of  an  octahedron.  Picturesquely,  a  group  such  as 


Fig.  22-4.  The  isomers  of  [CV(A///S)4C72]\ 


Ion  -with  +1  charge 


Cis.  isomer 


Tr, 


ans-  isomer 


SEC.    22-2    I    COMPLEX    IONS 


395 


Fig.  22-5.  The  structure  of  Fe(C£>t)»- 


oxalate,  which  can  furnish  simultaneously  two 
atoms  for  coordination,  is  said  to  be  bidentate, 
which  literally  means  double-toothed. 

In  addition  to  the  tetrahedral  and  octahedral 
complexes  mentioned  above,  there  are  two  other 
types  commonly  found— the  square  planar  and 
the  linear.  In  the  square  planar  complexes,  the 
central  atom  has  four  near  neighbors  at  the 
corners  of  a  square.  The  coordination  number 
is  4,  the  same  number  as  in  the  tetrahedral  com- 
plexes. An  example  of  a  square  planar  complex 
is  the  complex  nickel  cyanide  anion,  Ni(CN)^2. 

In  a  linear  complex,  the  coordination  number 
is  2,  corresponding  to  one  group  on  each  side  of 
the  central  atom.  The  silver-ammonia  complex, 
which  generally  forms  when  a  very  slightly  solu- 
ble silver  salt  such  as  silver  chloride  dissolves  in 
aqueous  ammonia,  is  an  example,  as  shown  in 
Figure  22-6.  Another  example  of  a  linear  com- 


Fig.  22-6.  A  linear  complex,  Ag{NH3)a\ 

Ion    yvt'th   +1  charge 


plex  is  Ag(CN)2",  which  is  formed  during  the 
leaching  of  silver  ores  with  NaCN  solution. 

22-2.2    Bonding  in  Complex  Ions 

What  holds  the  atoms  of  a  complex  ion  together? 
There  are  two  possibilities.  In  some  complexes, 
as  in  A1F6"3,  the  major  contribution  to  the  bond- 
ing comes  from  the  attraction  between  a  positive 
ion  (Al+3)  and  a  negative  ion  (F~).  The  bonding 
is  ionic.  In  other  complexes,  as  in  Fe(CN)6~3, 
there  is  thought  to  be  substantial  sharing  of  elec- 
trons between  the  central  atom  and  the  attached 
groups.  The  bonding  is  mainly  covalent.  When 
there  is  such  sharing,  an  electron  or  an  electron 
pair  from  the  attached  group  spends  part  of  its 
time  in  an  orbital  furnished  by  the  central  atom. 
In  either  type,  as  emphasized  in  Chapter  16,  the 
electron  is  attracted  to  both  atoms  in  the  bond. 

For  transition  elements  there  are  usually 
empty  d  orbitals  ready  to  accommodate  elec- 
trons from  attached  groups.  This  is  by  no  means 
always  necessary,  as  witness  the  case  of  Zn+2,  a 
good  complex-former  even  though  all  its  3d 
orbitals  are  already  occupied.  Any  vacant  or- 
bital, low  enough  in  energy  to  be  populated,  will 
serve  as  a  means  whereby  complex  formation 
can  be  accomplished. 

The  geometry  of  a  complex  ion  often  can  be 
explained  quite  reasonably  in  terms  of  the  or- 
bitals of  the  central  atom  populated  by  electrons 
from  the  attached  groups.  If  only  one  s  and  one 
p  orbital  is  used,  the  bonding  is  called  sp  bond- 
ing. We  have  already  seen  in  Section  16-4.5  that 
this  bonding  situation  gives  rise  to  a  linear  ar- 
rangement. Therefore  this  might  explain  why 
some  complexes  are  linear,  as  is  Ag(NH3)2+.  If 
one  5  and  three  p  orbitals  are  used,  the  complex 
uses  sp3  bonding.  Then  a  tetrahedral  complex 
can  be  expected,  as  observed  for  Zn(NH3)4+2. 
When  d  orbitals  are  involved,  other  geometries 
can  be  explained  (for  example,  square  planar, 
dsp2;  octahedral,  cPsp3). 

22-2.3    Significance  of  Complex  Ions 

Besides  their  occurrence  in  solid  compounds, 
complex  ions  such  as  we  have  mentioned  are 


396 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    I    CHAP.    22 


important  for  two  other  reasons:  (1)  they  may 
decide  what  species  are  present  in  aqueous  solu- 
tions; and  (2)  some  of  them  are  exceedingly  im- 
portant in  biological  processes. 

As  an  example  of  the  problem  of  species  in 
solution,  consider  the  case  of  a  solution  made 
by  dissolving  some  potassium  chrome  alum, 
KCr(SO,)r  12H20,  in  water.  On  testing,  the  solu- 
tion is  distinctly  acidic.  A  currently  accepted 
explanation  of  the  observed  acidity  is  based  upon 
the  assumption  that,  in  water  solution,  chromic 
ion  is  associated  with  six  H20  molecules  in  the 
complex  ion,  Cr(H20)c+3.  This  complex  ion  can 
act  as  a  weak  acid,  dissociating  to  give  a  proton 
(or  hydronium  ion).  Schematically,  the  dissocia- 
tion can  be  represented  as  the  transfer  of  a  pro- 
ton from  one  water  molecule  in  the  Cr(H20)^3 
complex  to  a  neighboring  H20  to  form  a  hy- 
dronium ion,  H30+.  Note  that  removal  of  a 
proton  from  an  H20  bound  to  a  Cr+3  leaves  an 
OH-  group  at  that  position.  The  reaction  is 
reversible  and  comes  to  equilibrium: 

Cr(H,0)+3  +  HO  +±:  Cr(HJ0)5OH+-  +  H30+    (3) 

We  see  that  Cr(H2O)0+3  acts  as  a  proton-donor, 
that  is,  an  acid. 


22-2.4    Amphoteric  Complexes 

Another  reason  chemists  find  the  above  complex 
ion  picture  of  aqueous  solutions  useful  is  that 
it  is  easily  extended  to  explain  amphoteric  be- 
havior. Take  the  case  of  chromium  hydroxide, 
Cr(OH)3,  a  good  example  of  an  amphoteric  hy- 
droxide. It  dissolves  very  little  in  water,  but  is 
quite  soluble  both  in  acid  and  in  base.  Presum- 
ably it  can  react  with  either.  How  can  this 
behavior  be  explained  in  terms  of  the  complex 
ion  picture? 

First,  consider  the  equilibrium  represented  by 
equation  (3)  when  NaOH  is  added  to  solution. 
Added  OH-  combines  with  the  H30+  to  form 
H20.  This  removes  one  of  the  species  shown  on 
the  right  side  of  the  equation,  so  formation  of 
Cr(H20)5OH+2  is  favored.  In  other  words,  as 
OH-  is  added  to  Cr(H2O)0+3  the  reaction  is 
favored  which  corresponds  to  pulling  a  proton 
off  Cr(H20)+3. 


What  happens  when  enough  NaOH  has 
been  added  to  remove  three  protons  from 
each  Cr(H20)G+3?  Removal  of  three  protons 
leaves  the  neutral  species  Cr(H20)3(OH)3,  or 
Cr(OH)3-3H20.  This  neutral  species  has  no 
charges  to  repel  other  molecules  of  its  own  kind 
so  it  precipitates.  However,  as  more  NaOH  is 
added  to  this  solid  phase,  one  more  proton  can 
be  removed  to  produce  Cr(H20)2(OH)4~  ;  and  the 
Cr(OH)3-3H20  dissolves.  [In  principle,  more 
protons  could  be  removed,  perhaps  eventually  to 
form  Cr(OH)6"3,  but  there  is  as  yet  no  evidence 
for  this.] 

The  following  equations  summarize  the  steps 
believed  to  occur  when  NaOH  is  slowly  added 
to  a  solution  of  chromic  ion.  Step  (4c)  corre- 
sponds to  formation  of  solid  hydrated  chromium 
hydroxide;  step  (4cl)  corresponds  to  its  dissolving 
in  excess  NaOH. 

Cr(H20)G+3  +  OH-  +±: 

Cr(H20)5OH+2  +  H20  (4a) 
Cr(H20)5OH+2  +  OH"  z^±: 

Cr(H20),(OH)2+  +  H.O  (4b) 
Cr(H20)J(OH)2+  +  OH-  ^=t 

Cr(H20)3(OH)3(sJ  +  HO  (4c) 
Cr(H20)3(OH)3(sJ  +  OH-  +±l 

Cr(H  ,0)2(OH)4-  +  H20       (4d) 

When  acid  is  added  to  a  solution  such  as  in 
equation  (4cf),  the  above  set  of  reactions  is 
progressively  reversed,  first  causing  precipitation 
of  chromium  hydroxide  by  the  reverse  of  reac- 
tion (4d)  and  then  its  subsequent  dissolving  by 
the  reverse  of  reaction  (4c). 

22-2.5    Complexes  Found  in  Nature 

Complex  ions  have  important  roles  in  certain 
physiological  processes  of  plant  and  animal 
growth.  Two  such  complexes  are  hemin,  a  part 
of  hemoglobin,  the  red  pigment  in  the  red  cor- 
puscles of  the  blood,  and  chlorophyll,  the  green 
coloring  material  in  plants.  The  first  of  these, 
hemoglobin,  contains  iron  and  properly  fits  in  a 
discussion  of  complex  compounds  of  the  transi- 
tion elements;  the  second,  chlorophyll,  is  a  com- 
plex compound  of  magnesium.  Magnesium  is 


SEC.    22-2    I    COMPLEX    IONS 


397 


not  a  transition  element,  but  chlorophyll  is  dis- 
cussed here  because  it  has  some  features  in 
common  with  hemoglobin  and  because  it  avoids 
the  impression  that  only  transition  elements  form 
complexes. 

Chlorophyll,  as  extracted  from  plants,  is  actu- 
ally made  up  of  two  closely  related  compounds, 
chlorophyll  A  and  chlorophyll  B.  These  differ 
slightly  in  molecular  structure  and  can  be  sepa- 
rated because  they  have  different  tendencies  to 
be  adsorbed  on  a  finely  divided  solid  (such  as 
powdered  sugar). 


five-membered  ring.  Consider  also  the  vast 
amount  of  knowledge  and  experimentation  that 
are  summed  up  in  the  statement  that  the  struc- 
ture of  this  complicated  molecule  is  known. 


EXERCISE  22-7 

If  a  typical  plant  leaf  yields  49.0  mg  of  chloro- 
phyll A,  how  many  milligrams  of  this  will  be 
magnesium?  The  molecular  weight  of  chloro- 
phyll A  is  893. 


EXERCISE  22-6 


If  you  wish  to  prepare  some  chlorophyll,  grind 
up  some  fresh  leaves  and  extract  with  alcohol. 
The  alcohol  dissolves  the  chlorophyll,  as  shown 
by  the  solution  color. 


To  show  the  complexity  of  this  biologically 
important  material,  the  structural  formula  of 
chlorophyll  A  is  shown  on  the  left  in  Figure  22-7. 
You  need  not  memorize  it.  The  most  obvious 
thing  to  note  is  that  it  is  a  large  organic  molecule 
with  a  magnesium  atom  in  the  center.  Right 
around  the  magnesium  atom  there  are  four  near- 
neighbor  N  atoms,  each  of  which  is  part  of  a 


Hemin  is  shown  on  the  right  in  Figure  22-7. 
It  is  shown  beside  the  model  of  chlorophyll  A  to 
emphasize  the  astonishing  similarity.  The  por- 
tions within  dotted  lines  identify  the  diiferences. 
Except  for  the  central  metal  atom,  the  differences 
are  all  on  the  periphery  of  these  cumbersome 
molecules.  We  cannot  help  wondering  how  na- 
ture managed  to  standardize  on  this  molecular 
skeleton  for  molecules  with  such  different  func- 
tions. We  cannot  avoid  a  feeling  of  impatience 
as  we  await  the  clarification  of  the  possible  rela- 
tionship, a  clarification  that  will  surely  be  pro- 
vided by  scientists  of  the  next  generation. 


Fig.  22-7.  The  structures  of  chlorophyll  A  and  hemin. 


Chlorophyll     A 


Hemin 


398 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    '    CHAP.    22 


The  most  important  function  of  hemoglobin 
in  the  blood  is  that  of  carrying  oxygen  from  the 
lungs  to  the  tissue  cells.  This  is  done  through  a 
complex  between  the  iron  atom  of  the  hemin 
part  and  an  oxygen  molecule.  Just  how  the  02 
is  bound  to  the  hemin  is  not  yet  clear,  but  it  must 
be  a  rather  loose  combination  since  the  02  is 
readily  released  to  the  cells.  The  complex  is 
bright  red,  the  characteristic  color  of  arterial 
blood.  When  the  02  is  stripped  off  the  hemin 
group,  the  color  changes  to  a  purplish  red,  the 
color  of  blood  in  the  veins. 

Not  only  02  molecules  but  also  other  groups 
can  be  bound  to  the  iron  atom  of  hemoglobin. 
Specifically,  carbon  monoxide  molecules  can  be 
so  attached  and,  in  fact,  CO  is  more  firmly 
bound  to  hemoglobin  than  is  02.  This  is  one 
detail  of  the  carbon  monoxide  poisoning  mecha- 
nism. If  we  breathe  a  mixture  of  CO  and  02 
molecules,  the  CO  molecules  are  preferentially 
picked  up  by  the  red  blood  cells.  Since  the  sites 


normally  used  to  carry  02  molecules  are  thus 
filled  by  the  CO  molecules,  the  tissue  cells  starve 
for  lack  of  oxygen.  If  caught  in  time,  carbon 
monoxide  poisoning  can  be  treated  by  raising 
the  ratio  of  02  to  CO  in  the  lungs  (in  other  words, 
administering  fresh  air  or  oxygen).  The  two  re- 
actions, 

02(g)  +  hemoglobin  -<->-  complex!       (5a) 
CO(g)  +  hemoglobin  -<->-  complex2       (5b) 

have  sufficiently  similar  tendencies  to  go  to  the 
right  that  the  first  reaction  can  be  made  to  exceed 
the  second  if  the  concentration  of  02  sufficiently 
exceeds  that  of  CO.  Another  remedial  measure 
is  to  inject  methylene  blue  directly  into  the  blood 
stream.  Carbon  monoxide  bonds  more  strongly 
to  methylene  blue  than  to  hemin.  Equilibrium 
conditions  then  favor  the  transfer  of  CO  to  the 
methylene  blue,  thus  freeing  hemoglobin  for  its 
normal  oxygen  transport  function. 


22-3    SPECIFIC  PROPERTIES  OF  FOURTH-ROW  TRANSITION  ELEMENTS 


The  preceding  discussion  of  the  transition  ele- 
ments has  been  quite  general,  with  the  implica- 
tion of  rather  wide  applicability.  Now  we  turn 
to  a  consideration  of  the  transition  elements  and 
their  compounds  as  specific  individuals. 

Table  22-111  collects  some  of  the  data  ordi- 
narily found  useful  for  the  transition  elements 
of  the  fourth  row  of  the  periodic  table.  The 
following  are  some  notes  on  regularities  ob- 
served. 

The  atomic  weight  increases  regularly  across 
the  row  except  for  the  inversion  at  cobalt  and 
nickel.  We  would  expect  the  atomic  weight  of 
Ni  to  be  higher  than  that  of  Co  because  there 
are  more  protons  (28)  in  the  Ni  nucleus  than  in 
the  Co  nucleus  (27).  The  reason  for  the  inversion 
lies  in  the  distribution  of  naturally  occurring 
isotopes.  Natural  cobalt  consists  entirely  of  the 
isotope  27C0;  natural  nickel  consists  primarily  of 
the  isotopes  ^Ni  and  asNi,  the  58-isotope  being 
about  three  times  as  abundant  as  the  60-isotope. 

Abundance  in  the  earth's  crust.  With  the  excep- 


tion of  iron,  which  is  very  abundant,  and  tita- 
nium, which  is  moderately  abundant,  all  the 
other  elements  of  the  first  transition  row  are 
relatively  scarce.  However,  some  of  them,  such 
as  copper,  are  quite  familiar.  Copper  is  one  of 
the  few  metallic  elements  found  free  in  nature. 
The  existence  of  deposits  of  metallic  copper 
undoubtedly  accounts  for  the  fact  that  man 
evolved  through  the  Bronze  Age  before  the  Iron 
Age.  Copper,  the  essential  ingredient  of  bronze, 
did  not  require  the  difficult  smelting  process 
needed  for  iron. 

Melting  point.  Except  for  zinc  at  the  end  of 
the  row,  the  melting  points  are  quite  high.  This 
is  appropriate,  since  these  elements  have  a  large 
number  of  valence  electrons  and  also  a  large 
number  of  vacant  valence  orbitals.  Toward  the 
end  of  the  row,  in  zinc,  the  3d  orbitals  become 
filled  and  the  melting  point  drops. 

Density.  There  is  a  steady  increase  in  density 
through  this  row,  with  some  leveling  off  toward 
the  right.  This  trend  is  closely  tied  to  the  almost 


SBC.    22-3    I   SPECIFIC    PROPERTIES    OF    FOURTH-ROW    TRANSITION    ELEMENTS 


399 


constant  size  of  the  atoms  so  the  main  effect 
producing  density  change  is  the  increasing  nu- 
clear mass. 

Ionization  energy.  As  ionization  energies  go, 
the  values  found  for  the  transition  elements  are 
neither  very  high  nor  very  low.  They  are  all 
rather  similar  in  magnitude.  The  sequential  in- 
crease in  nuclear  charge,  which  would  tend  to 
increase  the  ionization  energy,  seems  to  be  al- 
most offset  by  the  extra  screening  of  the  nucleus 
provided  by  the  added  electrons. 

Ionic  radius.  Ionic  radii  do  not  change  much 
in  going  across  a  transition  row.  The  reason  for 
this  is  essentially  a  balance  of  two  effects:  (1)  As 
nuclear  charge  increases  across  the  row,  the 
electrons  would  be  pulled  in,  so  the  ions  ought 


to  shrink.  (2)  As  more  3d  electrons  are  added, 
these  electrons  repel  each  other  and  the  ions 
ought  to  swell.  These  effects  just  about  cancel. 
As  expected,  the  size  of  the  +3  ion  is  smaller 
than  the  size  of  the  +2  ion  of  that  same  element. 
Keeping  nuclear  charge  constant,  removal  of  one 
additional  3d  electron  would  reduce  the  repul- 
sion between  the  3d  electrons  remaining,  thus 
allowing  them  all  to  be  pulled  closer  to  the 
nucleus. 

Color.  Many  solid  compounds  of  the  transi- 
tion metals  and  their  aqueous  solutions  are 
colored.  This  color  indicates  light  is  absorbed  in 


Fig.  22-8.  Atomic  sizes  of  the  transition  elements. 


2.5 


z.o 


1.5 

*     1.0 


0.5  - 


0 


Ck--- 


Melral 


o- 


-o o^ 


19K     20Ca    2lSc     22Ti      23V    24Cr  25Mn  26Fe     27Co  23™    29Cu   30Zn    31&a 


Etlemenrts 


400 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    !    CHAP.    22 


the  visible  part  of  the  spectral  region.  The  energy 
levels  that  account  for  this  absorption  are  rela- 
tively close  together  and  involve  unoccupied  d 
orbitals.  The  environment  of  the  ion  changes  the 
spacing  of  these  levels,  thereby  influencing  the 
color.  A  familiar  example  is  the  Cu+2(aq)  ion, 
which  changes  from  a  light  blue  to  a  deep  blue 
when  NH3  is  added.  The  formation  of  the  am- 
monia complex  alters  the  energy  level  spacing 
of  the  central  Cu+2  ion  to  produce  the  color 
change. 

E°.  The  last  row  of  the  table  gives  the  values 
of  the  oxidation  tendencies  for  these  metals. 
Except  for  scandium  (which  goes  to  a  +3  state), 
the  values  quoted  correspond  to  the  reaction 


M(s) 


U+Haq)  +  2e~ 


(6) 


As  can  be  seen  from  the  table,  all  the  elements 
except  copper  have  positive  values,  which  means 
these  metals  are  more  easily  oxidized  than  is 
hydrogen  gas,  for  which  E°  is  zero.  Thus,  man- 
ganese metal  should  dissolve  in  acid  to  liberate 
hydrogen  gas.  The  E°  for  the  overall  reaction, 

Mn(s)  +  2H+(aq)  +±  Un^(aq)  +  H2(g)    (7) 

is  +1-18  volts,  so  the  reaction  should  proceed 
spontaneously  to  the  right.  (Note  that  this  is  an 
equilibrium  consideration  and  it  tells  nothing 


about  the  rate.  The  rate  may  be  slow.)  For  cop- 
per the  reaction 

Cu(s)  +  2H+(aq)  ^=fc  Cu^(aq)  +  H2(g)     (8) 

has  a  negative  E°  (  —  0.34  volt),  so  reaction  to 
the  right  is  not  expected. 

Zinc,  at  the  end  of  the  row,  has  E°  (+0.76), 
which  is  intermediate  between  the  values  at  the 
beginning  of  the  row  and  those  toward  the  end. 
We  predict,  with  this  value  of  E°,  that  Zn  should 
reduce  Fe+2,  Co+2,  Ni+2,  and  Cu+2  to  the  corre- 
sponding metals  but  should  not  be  able  to  reduce 
Sc+3,  Ti+2,  V+2,  Cr+2,  or  Mn+2  to  the  metals. 

22-3.1    Scandium 

Scandium  has  not  yet  been  available  in  large 
enough  amounts  to  have  it  develop  interesting  or 
important  uses.  Neither  has  it  been  available  for 
much  experimental  work,  so  there  remains  much 
to  be  learned  about  this  element. 

22-3.2    Titanium 

There  is  intense  interest  in  titanium.  This  interest 
stems  from  an  unusual  combination  of  desirable 
properties  in  one  metal.  It  is  strong;  it  has  low 
density;  and  it  is  remarkably  resistant  to  corro- 


Table  22-11  I.     some  properties  of  fourth-row  transition  elements 

element  Sc  Ti  V  Cr  Mn  Fe  Co  Ni  Cu  Zn 


Atomic  number 

21 

22 

23 

24 

25 

26 

27 

28 

29 

30 

Atomic  weight 

45.0 

47.9 

51.0 

52.0 

54.9 

55.9 

58.9 

58.7 

63.5 

65.4 

Abundance* 

(%bywt.) 

0.005 

0.44 

0.015 

0.020 

0.10 

5.0 

0.0023 

0.008 

0.0007 

0.01 

Melting  point  (°C) 

1400 

1812 

1730 

1900 

1244 

1535 

1493 

1455 

1083 

419 

Boiling  point  (°C) 

3900 

3130f 

3530f 

2480| 

2087 

2800 

3520 

2800 

2582 

907 

Density  (g  cm3) 

2.4 

4.5 

6.0 

7.1 

7.2 

7.9 

8.9 

8.9 

8.9 

7.1 

First  ioniz.  energy 

(kcal/mole) 

154 

157 

155 

155 

171 

180 

180 

175 

176 

216 

+2  ion  radius  (A) 

— 

0.90 

0.88 

0.84 

0.80 

0.76 

0.74 

0.72 

0.72 

0.74 

+  3  ion  radius  (A) 

0.81 

0.76 

0.74 

0.69 

066 

0.64 

0.63 

0.62 

— 

— 

E°  (volt) 

M  — >-  M+J  +  2e~ 

2.1** 

1.6 

1.2 

0.90 

1.18 

0.44 

0.28 

0.25 

-0.34 

0.76 

*  In  the  earth's  crust. 

t  Estimated 

**  M  — >-  M+3  +  3e~ 


SEC.    22-3    |    SPECIFIC    PROPERTIES    OF    FOURTH-ROW    TRANSITION    ELEMENTS 


401 


sion.  The  difficulty  has  been  to  find  an  economi- 
cal way  of  getting  it  out  of  its  natural  minerals: 
rutile,  Ti02,  and  ilmenite,  FeTi03.  This  was 
solved  in  part  by  heating  Ti02  in  chlorine  gas  to 
convert  it  to  TiCU  and  then  reducing  the  TiCU 
with  magnesium  metal.  Two  problems  still  stand 
in  the  way  of  large-scale  use  of  this  rather  abun- 
dant element.  One  is  the  great  sensitivity  of  its 
properties  to  the  presence  of  trace  impurities 
(especially  H,  O,  C,  and  N);  the  other  is  the 
difficulty  of  forming  it  into  useful  shapes. 

22-3.3    Vanadium 

This  element  is  important  mainly  because  of  its 
use  as  an  additive  to  iron  in  the  manufacture  of 
steel.  A  few  percent  of  vanadium  stabilizes  a 
high-temperature  crystal  structure  of  iron  so  that 
it  persists  at  room  temperature.  This  form  is 
tougher,  stronger,  and  more  resistant  to  corro- 
sion than  ordinary  iron.  Automobile  springs,  for 
example,  are  often  made  of  vanadium  steel. 

Also  important  is  V205,  divanadium  pentox- 
ide,  an  orange  powder  which  is  used  as  a  catalyst 
for  many  reactions  of  commercial  significance. 
For  example,  in  the  manufacture  of  sulfuric  acid, 
V205  catalyzes  the  step  in  which  S02  is  oxidized 
to  S03.  How  it  works  is  still  in  dispute,  but  the 
general  belief  is  that  the  catalytic  action  is  de- 
pendent upon  the  ability  of  vanadium  to  show 
various  oxidation  states.  One  suggested  mecha- 
nism is  that  the  solid  V20B  absorbs  an  S02  mole- 
cule on  the  surface,  gives  it  an  oxygen  atom  to 
convert  it  to  S03,  and  is  itself  reduced  to  V204, 
divanadium  tetroxide.  The  V204  in  turn  is  re- 
stored to  V206  by  reaction  with  oxygen.  Catalytic 
reactions,  especially  those  involving  solid-gas 
interfaces,  are  not  very  well  understood  at  the 
present  time. 


some  reagents — chlorine,  for  instance.  In  air, 
however,  it  is  inactive,  probably  because  of  for- 
mation of  an  impervious  oxide  coat.  Other 
metals,  such  as  the  alkali  and  alkaline  earth 
metals,  also  form  oxide  coats  but  they  are  not 
very  effective  in  protecting  the  underlying  metal 
from  atmospheric  oxidation.  The  main  difference 
is  that  when  chromium  is  converted  to  oxide 
there  is  a  swelling  in  this  oxide  layer  that  arises 
from  the  increase  in  volume  per  chromium  atom. 
This  gives  a  nonporous  surface  coat  of  oxide. 
On  the  other  hand,  when  a  metal  such  as  calcium 
is  oxidized,  the  oxide  layer  has  a  smaller  volume 
per  Ca  atom  than  the  metal  itself.  The  result  is 
that  the  surface  layer  shrinks,  tending  to  crack 
and  open  up  fissures  through  which  oxygen  (and 
water  vapor)  can  reach  the  underlying  metal. 
Many  of  the  transition  elements  show  the  kind 
of  self-protective  action  found  in  chromium. 

Most  of  the  chromium  we  see  is  only  a  thin 
coating  on  iron  or  other  metals.  Such  a  coating 
called  chrome  plate,  is  put  on  in  an  electrolysis 
cell  in  which  the  object  to  be  plated  is  the 
cathode  of  the  cell.  The  essential  ingredients  of 
the  plating  bath  are  Cr03,  chrorruum(VI)  oxide, 
and  either  H2S04  or  Cr2(S04)3,  chromium(III) 
sulfate,  but  there  are  various  additives,  including 
such  unlikely  substances  as  glue  or  milk,  which 
are  supposed  to  give  better  coatings.  Pure  bulk 
chromium  metal  is  fairly  difficult  to  make.  It 
can  be  done  by  using  a  reaction  called  the 
Goldschmidt  reaction  in  which  aluminum  metal 
is  used  as  a  reducing  agent.  So  much  heat  is  re- 
leased when  A1203  is  formed  from  the  elements 
that  stable  oxides,  as  for  example  Cr203,  can  be 
reduced  by  Al.  A  mixture  of  aluminum  powder 
and  Cr203,  when  ignited,  gives  a  vigorous  reac- 
tion to  produce  A1203  and  chromium. 


22-3.4    Chromium 

Interest  in  this  metal  comes  from  its  remarkable 
inertness  to  atmospheric  corrosion.  Also,  it  is 
very  hard  and  thus  it  forms  an  ideal  protective 
coating.  On  the  basis  of  its  E°  (1.18  volts  higher 
than  hydrogen)  we  expect  chromium  to  be  quite 
reactive;  in  fact,  it  is  vigorously  reactive  with 


EXERCISE  22-8 

Write  the  equation  for  the  reduction  of  Cr203 
by  Al.  If  it  takes  399  kcal/mole  to  decompose 
A1203  into  the  elements  and  270  kcal/mole  to 
decompose  Cr203,  what  will  be  the  net  heat 
liberated  in  the  reaction  you  have  just  written? 


402 


THE    FOURTH-ROW    TRANSITION    ELEMENTS   I   CHAP.    22 


Metallic  chromium  is  an  ingredient  of  several 
important  alloys.  Some  forms  of  stainless  steel, 
for  example,  contain  about  12%  Cr.  Nichrome, 
which  is  commonly  used  for  heating  coils,  has 
about  15%  Cr  in  addition  to  60%  Ni  and  25% 
iron.  Both  these  alloys  are  quite  resistant  to 
chemical  oxidation. 

In  compounds,  the  important  oxidation  num- 
bers of  Cr  are  +2,  +3,  and  +6.  In  all  of  these 
states  the  chromium  ions  are  colored  and,  in 
fact,  the  element  got  its  name  from  this  property 
(chroma  is  the  Greek  word  for  color).  The  +2 
state  is  not  frequently  encountered  but  it  can  be 
made  quite  easily  as  the  beautiful  blue  chromous 
ion  in  solution  by  dripping  a  solution  containing 
Cr*  over  metallic  zinc.  Air  has  to  be  excluded 
since  02  rapidly  converts  Cr1"2  back  into  Cr*. 

The  +3  state  of  chromium  is  best  represented 
by  chromium(III)  oxide,  Cr203,  which  is  a  green, 
inert  solid  used  as  a  green  pigment.  It  can  be 
made  in  rather  spectacular  fashion  by  heating 
ammonium  dichromate.  Once  started,  the  re- 
action 

(mU)2CT2Ch(s)  ++ 

N2teJ  +  4H2Ofg;  +  Cr203(s)    (9) 

keeps  itself  going.  The  nitrogen  and  water  are 
formed  as  hot  gases  which  blow  the  light,  fluffy 
Cr203  about.  Another  way  of  getting  Cr203  is 
by  dehydrating  "chromium  trihydroxide"  ' 
heat: 


Ion    yvith 
-2    charge 


with 


2Cr(OH)3fsJ  *=£  Cr£>3(s)  +  m£>(g)     (70) 

There  is  much  argument  about  how  to  write  an 
appropriate  formula  for  "chromium  trihydrox- 
ide." A  gelatinous,  green  precipitate  does  form 
when  base  is  added  to  a  solution  containing 
chromic  ion  (Cr1"3),  but  it  includes  a  great  deal 
of  excess  H20,  so  pure  Cr(OH)3  is  never  ob- 
tained. This  is  a  common  problem  with  the 
transition  elements.  Their  hydroxides  are  not 
well-characterized,  mainly  because  of  the  con- 
siderable difficulty  of  distinguishing  between  a 
water  molecule  of  hydration  and  an  OH  group. 
For  example,  many  chemists  maintain  that  the 
formula  is  really  Cr203nH20,  and  that  n  usually 
is  found  to  equal  3.  This  agrees  with  the  empirical 
formula  Cr(OH),. 


EXERCISE  22-9 

Calculate  the  percent  chromium  in  Cr(OH)3  and 
inCr203-3H20. 


Whatever  that  green  precipitate  has  for  its 
chemical  formula,  it  is  observed  to  be  ampho- 
teric. It  dissolves  both  in  excess  acid  and  in 
excess  base,  as  explained  earlier. 


Fig.  22-9.  The  structures  of  chromate  and  dichromate 
ions. 


chrornaire    tori 


dichromate    ion 


;ec.   22-3  I  SPECIFIC   PROPERTIES  of  fourth-row  transition  elements 


403 


Probably  the  most  common  compound  of 
+  3  chromium  is  potassium  chrome  alum, 
KCr(S0,)212H20.  We  know  that  the  twelve 
water  molecules  are  distributed  equally,  six 
around  Cr+3  and  six  around  K+.  Potassium 
chrome  alum  is  just  one  example  of  the  general 
class  of  solids  called  alums  which  have  a  +1 
ion,  a  +3  ion,  two  sulfates,  and  twelve  molecules 
of  water.  In  the  dyeing  industry  chrome  alum 
is  used  for  fixing  dyes  to  fabrics. 

The  +6  state  of  chromium  is  represented  by 
the  chromates  and  dichromates.  The  chromate 
ion  is  a  tetrahedral  ion  with  Cr  at  the  center; 
dichromate  ion  may  be  visualized  as  two  such 
tetrahedra  having  one  oxygen  corner  in  common. 
Figure  22-9  shows  the  arrangements.  Chromates 
can  easily  be  converted  to  dichromates  by  addi- 
tion of  acid, 


2&0;2(aq)  +  2H+(aq) 


Cr207-2(aq)  +  H20    (//) 


The  change  can  be  followed  by  noting  the  color 
change  from  yellow  (characteristic  of  chromate) 
to  orange  (characteristic  of  dichromate).  The 
reverse  conversion  from  dichromate  to  chromate 
occurs  on  addition  of  base. 

Both  chromates  and  dichromates  are  strong 
oxidizing  agents.  One  example  of  their  use  is  in 
"cleaning  solution,"  a  mixture  of  K2Cr207  and 
concentrated  sulfuric  acid.  Laboratory  glassware 
can  be  thoroughly  cleaned  of  grease  films  by 
immersion  in  cleaning  solution. 

22-3.5    Manganese 

The  major  use  of  manganese  is  in  the  production 
of  steel,  during  which  manganese  reacts  with 
oxygen  and  keeps  the  gas  from  forming  bubbles 
when  the  iron  solidifies.  This  prevents  the  forma- 
tion of  structure-weakening  holes  in  finished 
steel.  Manganese  usually  occurs  in  nature  as  a 
mixture  of  oxides  along  with  oxides  of  iron.  The 
ore,  without  separation,  can  be  reduced  with 
carbon  in  a  high  temperature  furnace  to  give 
alloys  of  iron  and  manganese.  This  so-called 
"ferromanganese"  is  added  to  the  crude  molten 
iron  on  its  way  to  becoming  steel. 
Probably  the  most  commonly  encountered 


compound  of  Mn  is  manganese  dioxide,  Mn02, 
used  in  the  ordinary  flashlight  cell.  What  goes  on 
in  a  dry  cell  when  it  generates  electricity  is  very 
much  in  dispute.  The  reactions  are  complicated 
and  apparently  change  character  depending  upon 
the  amount  of  electric  current  being  drawn 
from  the  cell.  When  small  currents  are  involved, 
which  is  what  dry  cells  are  designed  for,  the 
reactions  are  believed  to  be  as  follows: 


AT  THE  ANODE 

(the  terminal  in- 
dicated as  the 
negative  pole  on 
commercial  cells) 

Zn(s) 
AT  THE  CATHODE 


the  zinc  container  is  oxidized 
from  the  metallic  state  to  the 
+2  state,  probably  as  some 
complex  zinc  ion,  but  writ- 
ten for  simplicity  as  Zn+2 


Zn+2  +  2e- 


U2) 


the  Mn02  picks  up,  in  a 
complicated  reaction,  an 
electron  which  reduces  the 
manganese  from  the  +4  to 
+3  state  in  the  presence  of 
NH.Cl 


2Mn02  +  2NH4+  +  2e~  — >■ 

Mn203  +  2NH3  +  H20 


(13) 


The  role  of  the  other  components  in  the  cell  is 
not  completely  understood.  Some  of  these  com- 
ponents (such  as  NH4C1,  ammonium  chloride, 
and  ZnCl2,  zinc  chloride,  in  the  center  paste)  are 
involved  in  other  reactions  that  come  into  play 
at  larger  current  drain. 

One  other  important  compound  of  manganese 
is  potassium  permanganate,  KMn04.  This  is  an 
intensely  violet-colored  material  much  used  as 
an  oxidizing  agent  in  the  laboratory.  (It  is  too 
expensive  to  use  on  a  large  scale;  in  industry, 
chlorine  is  more  likely  to  be  used.) 

22-3.6     Iron 

Iron  is  the  workhorse  of  the  metals.  It  is  quite 
abundant  (ranking  fourth  of  all  elements  and  sec- 
ond of  the  metals,  by  weight)  and  easy  to  make 
inexpensively  on  a  large  scale;  and  it  has  useful 
mechanical  properties,  especially  when  alloyed 
with  other  elements.  Steel,  one  of  our  most  use- 
ful construction   materials,   is  essentially   iron 


404 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    I    CHAP.    22 


containing  a  small  percentage  of  carbon  and 
often  small  amount  of  other  elements. 

NATURAL  OCCURRENCE  OF  IRON 

Most  of  the  accessible  iron  is  combined  either 
with  oxygen  or  sulfur.  The  oxygen  compounds 
are  the  common  minerals  hematite,  Fe203,  and 
magnetite,  Fe304,  both  of  which  are  useful  raw 
materials  for  producing  iron.  Another  mineral, 
FeS2,  is  called  iron  pyrites  or  "fool's  gold,"  but 
this  is  not  a  common  source  since  removing  all 
the  sulfur  is  difficult.  (Sulfur  impurity  in  steel 
makes  it  brittle.  The  sulfur  compounds  of  iron 
are  low  melting  and  on  cooling  they  stay  liquid 
longer  than  does  the  mass  itself  and  keep  the 
iron  from  compacting  efficiently.)  Iron  exists  in 
the  elemental  form. in  some  meteorites.  Since 
meteorites  are  believed  to  come  from  the 
break-up  of  a  planet,  the  existence  of  iron  me- 
teorites is  taken  as  support  for  the  theory  that 
the  core  of  the  earth  is  largely  iron. 

MANUFACTURE    OF    IRON 

The  production  of  iron  is  an  excellent  example 
of  chemical  reduction  on  a  massive  scale.  The 
process  is  carried  out  in  a  huge  vertical  reactor, 
called  a  blast  furnace,  which  may  be  several 
stories  tall.  Raw  materials — iron  ore,  limestone, 
and  coke — are  fed  into  the  top  of  the  furnace 
and  oxygen  is  blown  in  at  the  bottom.  The  pur- 
pose of  the  iron  ore  (let  us  assume  Fe203)  is  to 
provide  the  iron.  The  limestone  reacts  with  sand, 
Si02,  in  the  iron  ore  removing  it  as  molten 
calcium  silicate,  called  slag.  The  coke  supplies 
the  reducing  agent,  carbon,  and  as  it  reacts  the 
heat  released  maintains  the  required  high  tem- 
perature. 

Here  is  a  simplified  version  of  what  goes  on  in 
a  blast  furnace.  As  the  mixture  of  ore,  limestone, 
and  coke  falls  through  the  furance,  it  meets  the 
updraft  of  oxygen.  Carbon  monoxide  is  formed, 

2C(s)  +  02(g)  — >-  2CO(g)  +  52.8  kcal     (14) 

and  as  this  carbon  monoxide  rises  through  the 
furnace  it  progressively  reduces  the  Fe203 — first 
to  Fe304,  then  to  FeO,  and  eventually  to  Fe.  The 
successive  reactions  take  place  progressively  as 
the  solid  descends: 


CO(  g)  +  3Fe..(VsJ 
COte)  +  Fe304(sj 

CO(g)  +  FeOfsJ 


2Fe30/sj  +  CO-,(g)  (15) 
3FeO(s)  +  C02(g)  (16) 
Fe(l)  +  CQ2(g)  (17) 


Since  the  reactions  (15),  (16),  and  (17)  require 
successively  higher  temperatures,  the  blast  fur- 
nace temperature  is  kept  highest  near  the  bottom 
of  the  furnace.  Near  the  bottom,  the  temperature 
is  sufficiently  high  that  the  impure  iron — satu- 
rated with  carbon— collects  there  as  a  molten 
liquid.  The  slag,  which  is  mainly  calcium  silicate, 
CaSi03,  removes  any  sand  in  the  ore  through 
reaction  with  limestone,  CaC03. 

CaC03(sJ  — ►-  CaOfsJ  +  C02(g)         (18) 

CaO(s)  +  SiO/sJ  — ►■  CaSi03(l)         (19) 

Molten  CaSi03  is  less  dense  than  molten  iron 
and  floats  on  top  of  it.  An  average  furnace  that 
produces  about  750  tons  of  iron  per  day,  will 
also  yield  410  tons  of  slag.  The  slag  is  sometimes 
useful  in  the  manufacture  of  cement  and,  when 
it  contains  sufficient  phosphorus,  in  the  manu- 
facture of  fertilizer. 

When  this  impure  iron  cools,  the  resulting 
solid  is  called  pig  iron  or  cast  iron.  It  is  quite 
brittle  and  is  not  useful  if  high  strength  is  needed. 
The  impure  iron  is  made  into  steel  by  burning 
out  most  of  the  carbon,  sulfur,  and  phosphorus. 
Today  there  are  three  common  furnace  types  for 
making  steel — the  open-hearth  furnace  (85  %  of 
U.S.  production),  the  electric  arc  furnace  (10%), 
and  the  Bessemer  converter  (5  %).  These  furnaces 
differ  in  construction  but  the  chemistry  is  basi- 
cally similar. 

The  process  of  burning  out  the  impurities  is 
slowest  in  the  open-hearth  furnace.  This  implies 
there  is  plenty  of  time  to  analyze  the  melt  and 
add  whatever  is  needed  to  obtain  the  desired 
chemical  composition.  Manganese,  vanadium, 
and  chromium  are  frequent  additives.  The  prop- 
erties of  the  finished  steel  depend  upon  the 
amount  of  carbon  left  in  and  upon  the  identity 
and  the  quantity  of  other  added  elements.  Soft 
steel,  for  example,  contains  0.08-0.18  weight  per- 
cent carbon;  structural  steel,  0.15-0.25%;  hard 
steel  or  tool  steel,  1-1.2%. 

The  electric  arc  furnace  is  used  for  special 
purpose  steels.  Because  the  environment  can  be 


SEC.    22-3    !    SPECIFIC    PROPERTIES    OF    FOURTH-ROW    TRANSITION     llfMENTS 


405 


controlled,  electrically  heated  crucibles  avoid 
contamination  problems  caused  by  chemical 
fuels. 

The  Bessemer  converter  is  the  oldest  of  the 
three  methods  and  the  fastest  (about  15  minutes 
per  charge).  However,  the  speed  is  a  mixed  bless- 
ing because  there  is  not  sufficient  time  to  make 
analyses  and  fine  adjustments  in  the  amounts 
of  the  alloying  elements. 

RUSTING 

One  well-known  property  of  iron  is  the  way  in 
which  it  tends  to  go  back  to  oxide  from  which 
it  was  derived.  In  fact,  one  out  of  every  four  men 
in  the  steel  industry  is  concerned  essentially  with 
replacing  iron  lost  by  rusting!  This  shows  how 
important  corrosion  is.  What  is  the  chemical 
nature  of  rusting,  and  how  can  it  be  controlled? 
First,  rusting  is  a  special  case  of  corrosion  in 
which  the  metal  being  corroded  is  iron  and  the 
corroding  agent  is  oxygen.  The  observed  facts 
are  that  H20  and  02  are  necessary;  H+(aq) 
speeds  up  the  reaction;  some  metals  such  as  Zn 
hinder  the  corrosion,  other  metals  such  as  Cu 
speed  it  up;  and  strains  (as  are  produced  when 
a  nail  is  bent)  usually  accelerate  the  reaction. 

How  can  these  observations  be  interpreted? 
The  most  promising  mechanism  suggested  is  a 
many-step  process  in  which  the  following  se- 
quence of  events  occurs:  (1)  the  iron  acts  as  an 
anode  to  give  up  two  electrons  and  form  Fe+2 
(ferrous)  ion;  (2)  the  electrons  are  picked  up  by 
H+(aq)  ions  to  form  transient  neutral  H  atoms; 
(3)  the  H  atoms  are  immediately  oxidized  by  02 
to  form  H20;  (4)  the  Fe+2  is  oxidized  by  02  in 
the  presence  of  H20  to  form  rust.  Rust,  in- 
cidentally, is  not  a  simple  compound  but  seems 
to  be  an  indefinite  hydrate  of  Fe203,  so  it  is  fre- 
quently given  the  formula  Fe203«H20. 

Acids,  for  example  those  in  fruit  juices,  cata- 
lyze rust  formation  because  they  furnish  H+(aq) 
to  accept  electrons  from  the  iron,  causing  it  to 
dissolve  faster.  Oxygen  gas  is  necessary  to  oxidize 
Fe+2  to  Fe203.  The  presence  of  water  facilitates 
the  migration  of  Fe+2  from  the  reaction  site.  The 
resulting  reduction  in  Fe+2  concentration  allows 
more  to  be  formed.  Support  for  these  ideas 
comes  from  the  frequent  observation  that  when 


the  0_>  supply  is  restricted  (as  under  a  rivet  head) 
the  iron  is  eaten  away  at  one  spot  (shank  of  the 
rivet)  but  the  rust  deposits  where  the  02  is 
plentiful  (where  the  rivet  head  overlaps  a  plate). 
One  can  surmise  that  the  rivet  shank  is  dissolved 
by  the  half-reaction 


Fe(s) 


Fe  2  +  2e- 


(20) 


in  some  acidic  solution,  perhaps  rain  water  con- 
taining C02.  Then  the  Fe+2  could  have  been 
washed  out  to  the  surface  where  oxidation  con- 
verted Fe+2  to  Fe203/?H20.  A  similar  explana- 
tion would  hold  for  the  observation  that  iron 
pipes  buried  in  the  soil  near  cinders  usually  rust 
rapidly.  Cinders  generally  contain  acid-forming 
oxides,  which  could  help  speed  up  the  dissolving 
of  iron. 

PREVENTION    OF    CORROSION 

The  observed  effect  of  metals  on  the  rate  of 
rusting  also  supports  the  above  theory  and  sug- 
gests a  way  to  stop  the  corrosion.  When  zinc  is 
in  close  contact  with  iron,  the  iron  does  not 
corrode  but  the  zinc  tends  to  oxidize  away.  The 
belief  here  is  that  the  zinc,  with  a  more  positive 
£°  than  Fe,  gives  electrons  up  to  the  iron, 
effectively  preventing  Fe  from  dissolving.  This 
kind  of  protection  is  called  cathodic  protection 
and  has  a  variety  of  applications.  For  example, 
ship  hulls,  particularly  of  tankers,  are  so  pro- 
tected in  sea  water.  Magnesium  is  used  rather 
than  zinc  but  the  principle  is  the  same.  Easily 
replaceable  blocks  of  magnesium  are  bolted  to 
the  steel  hulls  and  the  magnesium  oxidizes  in- 
stead of  the  hull.  Zinc  coated  iron  ("galvanized 
iron")  furnishes  a  second  example.  The  zinc 
fortunately  does  not  oxidize  very  much  because 
when  it  reacts  with  oxygen  and  water  in  the 
presence  of  C02  it  forms  a  self-protective  coat  of 
basic  zinc  carbonate.  Thus  the  zinc  is  self- 
protective  and  at  the  same  time  gives  cathodic 
protection  to  the  underlying  iron. 

Some  metals  such  as  copper  or  tin,  when  in 
contact  with  iron,  actually  speed  up  the  rate  of 
rusting.  The  reason  for  this  is  that  on  these 
metals,  reaction  of  electrons  with  W+(aq)  is 
more  rapid  than  on  iron  itself.  Thus  the  effect  is 
to  draw  the  electrons  away  from  the  iron,  speed- 


406 


THE    FOURTH-ROW    TRANSITION    ELEMENTS   I    CHAP.    22 


ing  up  the  rate  at  which  Fe  goes  to  Fe+2.  Tin 
itself  is  inert  to  the  atmosphere  so  a  piece  of  iron 
completely  covered  with  tin  is  safe  from  rusting. 
However,  once  the  protective  tin  coat  is  punc- 
tured the  rusting  of  the  iron  will  be  faster  than 
if  the  tin  were  not  there  at  all.  This  accounts  for 
the  observation  that  "tin  cans,"  which  are  tin- 
covered  steel,  rust  very  quickly  once  they  start. 
One  of  the  easiest  ways  to  prevent  rusting  of 
iron  objects  is  to  shut  out  the  supply  of  02  and 
H20.  This  can  be  done  by  painting  the  object  or 
by  smearing  it  with  grease.  The  only  caution  here 
is  to  do  the  job  thoroughly,  since  an  exclusion 
of  02  and  H20  that  is  only  partial  can  do  as 
much  harm  as  good.  Witness  the  rivet  that  would 
have  been  saved  if  there  were  a  good  tight 
coating  of  paint  to  seal  the  lip  against  entrance 
of  the  dissolving  solution. 

22-3.7    Cobalt 

This  element  does  not  appear  in  the  headlines 
very  often  but  it  is  of  practical  importance.  Prob- 
ably its  greatest  single  use  is  in  alloys,  including 
stainless  steels.  Pure  cobalt  is  almost  as  magnetic 
as  iron  and,  when  alloyed  with  aluminum,  nickel, 
copper,  and  iron,  the  resulting  Alnico  alloy  has 
a  permanent  magnetization  far  exceeding  that  of 
iron. 


Ni2  Os 
cathode 


In  most  of  its  simple  compounds,  cobalt  has  a 
+2  oxidation  number.  This  includes  the  well 
known  cobaltous  chloride.  In  dilute  aqueous 
solution  this  salt  is  an  almost  invisible  pink;  on 
dehydration  it  changes  to  deep  blue.  The  color 
change,  which  is  ascribed  to  a  replacement  of 
some  of  the  water  molecules  surrounding  Co+2 
by  Cl~  ions  (to  form  a  complex  ion),  is  exploited 
in  "invisible  inks,"  the  writing  of  which  appears 
on  heat  application.  Another  use  is  the  simple- 
weather  forecasters  in  which  a  swatch  of  blotting 
paper  turns  pink  when  the  humidity  rises,  sug- 
gesting that  rain  is  supposed  to  be  coming. 

In  its  complex  compounds,  of  which  there  are 
many  thousands,  Co  almost  invariably  has  a  +3 
oxidation  number.  Apparently,  Co+3  ion  accom- 
panied by  six  coordinating  groups  is  particularly 
stable.  Cobalt  complexes  are  important  in  bio- 
chemistry. Some  enzyme  reactions  go  through  a 
cobalt-complexing  mechanism.  Although  only 
small  traces  are  needed,  cobalt  is  essential  to  the 
diet. 

22-3.8    Nickel 

The  five-cent  coin,  ordinarily  called  the  "nickel," 
is  actually  25%  nickel  (the  other  75%  is  copper). 

Fig.  22-10.  Cells  from  Edison  and  lead  storage  bat- 
teries {schematic) . 


Edison     Celt 


<fr 


Lead     Cell 


SEC.    22-3    |    SPECIFIC    PROPERTIES    OF    FOURTH-ROW    TRANSITION    ELEMENTS 


407 


This  familiar  metallic  object  furnishes  an  im- 
portant example  of  a  nickel  alloy.  Other  impor- 
tant nickel  alloys  include  the  nickel  steels,  which 
are  tough  and  rust-resistant,  Monel  metal  (60% 
Ni,  40%  Cu),  which  is  acid-proof,  and  Nichrome 
(60%  Ni,  25%  Fe,  15%  Cr),  mentioned  in  Sec- 
tion 22-3.4.  Finely  divided  nickel  is  used  as  a 
catalyst  for  hydrogenation,  the  addition  of  hy- 
drogen to  double-bonded  carbon  atoms.  This 
process  is  important  in  the  foods  industry  to 
convert  edible  vegetable  oils,  such  as  cottonseed 
oil,  into  solid  fats.  The  carbon  chains  of  the 
vegetable  oils  have  double  bonds  which  have  a 
high  tendency  to  become  oxidized  and  tend  to 
develop  an  unpleasant  flavor.  When  H2  is  added 
to  the  double  bond,  the  carbon  chain  becomes 
saturated  and  the  material  becomes  more  attrac- 
tive to  the  cook.  Oleomargarine  is  an  example  of 
such  a  catalytically  hydrogenated  compound. 

In  most  of  its  compounds  nickel  has  a  +2 
oxidation  number,  but  it  is  possible  to  get  a 
higher  state  by  heating  Ni(OH)2  with  hypochlo- 
rite ion  in  basic  solution.  Hypochlorite  ion, 
CIO-,  is  one  of  the  stronger  oxidizing  agents  at 
our  disposal  in  basic  solution.  There  is  consider- 
able argument  about  the  formula  of  the  black 
solid  that  is  formed,  but  we  shall  label  it  as  Ni203 
and  write  the  equation 

2Ni(OH)2(sj  +  C\0-(aq)  +±: 

Ni2CVs,)  +  2H20  +  C\-(aq)     (21) 

Nickel(III)  oxide  is  important  as  the  oxidizing 
agent  in  the  Edison  storage  cell,  shown  in  Figure 
22-10.  Table  22-IV  compares  the  Edison  battery 
and  the  more  common  "lead  storage  battery." 
In  both  batteries  the  electrode  products  are 
solids  and  they  adhere  to  the  electrodes;  the 
batteries  can  be  regenerated  by  reversing  the 
flow  of  electricity  with  some  external  device  such 
as  a  direct  current  generator.  "Charging  a  bat- 
tery" simply  means  reversing  the  half-reaction 
at  each  electrode.  It  should  be  noted  that  on 
discharge,  in  the  Edison  cell  there  is  no  net 
consumption  of  the  electrolyte,  potassium  hy- 
droxide, so  its  concentration  stays  constant.  In 
contrast,  in  the  lead  cell  the  sulfuric  acid  elec- 
trolyte is  consumed  during  use  of  the  cell  and  is 
regenerated  during  charging.  This  variability  of 


H2S04  concentration  during  use  of  a  lead  cell 
provides  the  basis  for  the  convenient  hydrometer 
test  of  the  state  of  discharge  of  an  automobile 
battery.  The  hydrometer  measures  the  density 
of  the  electrolyte  solution,  thus  indicating  how 
much  of  the  H2S04  has  been  consumed.  Obvi- 
ously, this  method  cannot  be  used  to  check  an 
Edison  cell  since  the  electrolyte  concentration  is 
constant. 


Table  22-IV 

COMPARISON    OF    EDISON    AND 
LEAD    STORAGE    BATTERIES 

EDISON  LEAD 

Oxidizing  agent  Ni203  Pb02 

Reducing  agent  Fe  Pb 

Electrolyte  KOH  H2SO< 

Voltage  (one  cell)       1.35  volts         2.0  volts 
Features  light  weight      heavy 

constant  variable  voltage  dur- 

voltage             ing  discharge 
expensive          inexpensive 
rugged              voltage  and  H2SO< 
density  indicate 
when  recharge 
needed 

EDISON  CELL  (DURING  DISCHARGE) 

Anode  reaction 

Ft(s)  +  20H-(aq)  — >■  Ft(OHh(s)  +  2e~     (22a) 
Cathode  reaction 
Ni203(sj  +  3H20  -I-  2e~  — >- 

2Ni(OH)2rs>l  +  20H-(aq)    (22b) 
Net  reaction 
Fe(s)  +  Ni2CVsj  +  3H20  — >- 

Fe(OH)jfsJ  +  2Ni(OH)2rsJ     (22c) 

LEAD  CELL  (DURING  DISCHARGE) 

Anode  reaction 

PbfsJ  +  HSOr  (aq) ►- 

PbSCVs;  +  H+(aq)  +  2e~     (23a) 
Cathode  reaction 
PbCVsJ  +  HSOr(aq)  +  3H+(aq)  +  2e~  — )- 

PbSCVsJ  +  2HsO    (23h) 
Net  reaction 

Pb(s)  +  PbCVsJ  +  2H +(aq)  +  2HSCV  — >- 
2PbSO«fsj  +  2HjO    (23c) 


408 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    I    CHAP.    22 


Edison  batteries  cost  more  than  lead  storage 
batteries,  but  they  have  (he  advantage  of  being 
lighter,  so  the  amount  of  electrical  energy  avail- 
able per  unit  weight  is  greater.  Also  they  are 
more  rugged  in  standing  up  to  mechanical  shock. 
The  difficulty  of  determining  when  recharging 
is  needed  and  the  expense  are  disadvantages. 


22-3.9    Copper 

This  element  occurs  in  nature  in  the  uncombined 
state  as  native  copper  and  in  the  combined  state 
as  various  oxides,  sulfides,  and  carbonates.  The 
chief  mineral  is  chalcopyrite,  CuFeS2,  from 
which  the  element  is  extracted  by  roasting  (heat- 
ing in  air)  followed  by  reduction.  The  roasting 
reaction  can  be  written 


4CuFeS2(s)  +  902(g)  — >- 

2Cu2SfsJ  +  2Fe2Q3(s)  +  6SQ2(g) 


(24) 


It  shows  the  formation  of  the  important  by- 
product S02,  which  may  be  converted  to  H2S04. 
It  also  shows  the  formation  of  Fe203,  which  is 
subsequently  removed  by  adding  sand  and  heat- 
ing in  a  furnace.  The  sand  furnishes  Si02  which 
combines  with  the  iron  oxide  to  form  a  low 
melting  slag  of  iron  silicate.  After  the  slag  is  run 
off,  the  Cu2S  is  heated  in  a  current  of  air  to  give 
consecutively 

2Cu2S(sJ  +  302(g)  — >-  2Cu2OfsJ  +  2S02(g)    (25) 

Cu2S(s)  +  2Cu2OfsJ  — >-  6Cu(sJ  +  S02(g)       (26) 

The  copper  obtained  from  this  process  is  about 
99%  pure,  yet  this  is  not  pure  enough  for  most 
uses,  especially  those  involving  electrical  con- 
ductivity. To  refine  the  copper  further,  it  is  made 
the  anode  of  an  electrolytic  cell  containing  cop- 
per sulfate  solution.  With  careful  control  of  the 
voltage  to  regulate  the  half-reactions  that  can 
occur,  the  copper  is  transferred  from  the  anode 
(where  it  is  about  99%  Cu)  to  the  cathode  where 
it  can  be  deposited  as  99.999%  Cu.  At  the  anode 
there  is  oxidation  of  copper, 

Cu(s)  — *■  Cu+2(aq)  +  2e~  (27) 

along  with  oxidation  of  any  other  metal  (such  as 
Fe)  which  is  more  readily  oxidized  than  copper. 
The  elements  less  readily  oxidized  than  copper 


(such  as  silver  and  gold)  simply  crumble  off  into 
a  heap  under  the  anode.  This  so-called  "anode 
sludge"  can  then  be  worked  to  recover  and  iso- 
late these  very  valuable  by-products.  At  the 
cathode  there  is  reduction  but,  with  well  regu- 
lated voltage,  only  of  copper. 


Cu+*(aq)  +  2e~ 


Cu(s) 


(28) 


Of  course,  as  time  goes  on  the  copper  sulfate 
solution  in  the  cell  has  to  be  replaced  because 
it  collects  undesirable  ions  such  as  Fe+2,  which 
have  not  been  reduced  because  the  voltage  used 
is  favorable  to  the  reduction  of  Cu+2  only. 

In  compounds,  copper  usually  has  a  +2  oxi- 
dation number,  cupric,  or  copper(II),  and  oc- 
casionally +  1,  cuprous,  or  copper(I).  The 
most  common  compound  of  cupric  copper  is 
CuS04-5H20.  The  blue  color  of  this  solid  is  due 
to  the  Cu+2  ion  hydrated  by  four  of  the  five 
H20  molecules  in  a  surrounding  square  planar 
arrangement.  This  ion  is  present  in  aqueous  solu- 
tions as  well,  but  with  two  more  distant  H20 
molecules  along  the  axis  perpendicular  to  this 
square.  When  cupric  solutions  are  treated  with 
an  excess  of  ammonia,  they  turn  a  deep  blue. 
This  is  attributed  to  formation  of  a  tetra- 
ammine  copper(Il)  ion  complex,  usually  written 
C^NHs)^2  but  probably  also  containing  two 
additional  H20  molecules  coordinated  to  the 
copper.  Cupric  ion  is  toxic  to  lower  organisms, 
so  it  is  used  to  suppress  the  growth  of  algae  in 
ponds  and  fungi  and  molds  on  vines.  Bordeaux 
mixture  used  to  spray  grapes  and  potatoes  is 
made  of  copper  sulfate  and  lime. 

The  + 1  state  of  copper  is  found  only  in  com- 
plex compounds  or  slightly  soluble  compounds. 
The  reason  for  this  is  that  in  aqueous  solution 
cuprous  ion  is  unstable  with  respect  to  dispro- 
portionation  to  copper  metal  and  cupric  ion. 
This  comes  about  because  cuprous  going  to 
cupric  is  a  stronger  reducing  agent  than  copper 
going  to  cuprous.  The  following  exercise  in  the 
use  of  E°  puts  this  on  a  more  quantitative  basis: 


Cu(s)  — >-  Cu+(aqj  +  e~ 


E°  =  -0.52  volt    (29) 


Cu+(aq)  — >-  Cu+Yaqj  +  e~ 


E°  =  -0.15  volt    (30) 


QUESTIONS    AND    PROBLEMS 


409 


Since  reaction  (30)  has  a  more  positive  E°  than 
reaction  (29),  it  can  force  reaction  (29)  to  reverse, 
thus  in  effect  transferring  an  electron  from  one 
Cu+  ion  to  another  Cu+  ion.  The  net  reaction, 
obtained  by  subtracting  reaction  (29)  from  re- 
action (30)  is 

2Cu+(aq)  +=±  Cu(s)  +  Cu+*(aq)  (31) 

which  has  an  E°  of  —0.15  minus  —0.52  volt, 
or  +0.37  volt.  Positive  £0,s  for  net  reactions 
mean  the  reaction  should  take  place  spontane- 
ously from  left  to  right. 

22-3.10    Zinc 

We  have  already  encountered  zinc  as  the  irregu- 
lar member  at  the  end  of  the  fourth  transition 
row.  We  have  also  mentioned,  in  Section  17-2.3, 
its  use  as  a  constituent  of  the  important  class  of 
alloys  called  brasses  and  its  use  in  "galvanizing" 
iron  to  protect  iron  from  rusting.  Galvanized 
iron  is  made  by  dipping  iron  into  molten  zinc  so 
as  to  give  a  thin  adhering  layer  of  Zn  over  the  Fe. 
On  prolonged  exposure  to  air  containing  C02  the 
zinc  forms  a  thin  protective  skin  of  basic  zinc 


carbonate  (zinc  hydroxycarbonate).  When  a  hole 
forms,  penetrating  into  the  iron,  the  iron  does 
not  rust  as  would  be  the  case  with  tin-coated 
iron.  On  the  contrary,  the  fresh  Zn  surface  ex- 
posed reacts  with  C02,  02,  and  H20  of  the  air 
to  form  a  plug  of  zinc  hydroxycarbonate  which 
seals  the  hole. 

One  other  interesting  and  important  com- 
pound of  zinc  is  the  sulfide.  ZnS.  It  is  the  mineral 
zinc  blende,  one  of  the  major  sources  of  zinc 
and,  also,  it  is  the  luminescent  material  on  the 
face  of  many  television  picture  tubes.  Zinc  sulfide 
is  a  semiconductor  and,  when  a  beam  of  elec- 
trons strikes  the  screen,  electrons  in  the  solid  are 
energized  so  they  can  wander  through  the  ZnS 
much  like  the  electrons  in  a  metal.  When  these 
electrons  find  an  attractive  site,  usually  in  the 
vicinity  of  a  purposely  added  impurity  atom, 
they  can  be  trapped  and  give  off  energy  as  visible 
light.  This  phenomenon,  called  fluorescence, 
makes  possible  the  conversion  of  one  frequency 
of  light  energy  to  another.  The  observed  color 
of  the  fluorescence  depends  upon  the  mode  of 
preparing  the  ZnS  and  on  the  nature  of  the 
impurity  in  the  ZnS  structure. 


QUESTIONS  AND  PROBLEMS 


1.  Why  are  the  elements  with  atomic  numbers  21  to 
30  placed  in  a  group  and  considered  together  in 
this  chapter? 

2.  Write  the  orbital  representation  for 

(a)  chromium, 

(b)  molybdenum, 

(c)  tungsten. 

3.  What  properties  of  the  transition  elements  are 
consistent  with  their  being  classified  as  metals? 

4.  Ferrous  ion,  iron(II),  forms  a  complex  with  six 
cyanide  ions,  CN~;  the  octahedral  complex  is 
called  ferrocyanide.  Ferric  ion,  iron(III),  forms 
a  complex  with  six  cyanide  ions;  the  octahedral 
complex  is  called  ferricyanide.  Write  the  struc- 
tural formulas  for  the  ferrocyanide  and  the  ferri- 
cyanide complex  ions. 


5.  Draw  the  different  structures  for  an  octahedral 
cobalt  complex  containing  four  NH3  and  two 
NO:  groups. 

6.  Draw  the  structures  of  the  compounds 

Cr(NH3)6(SCN)3 
Cr(NH3)3(SCN)3 

(SCN~  is  the  thiocyanate  ion).  Consider  the  oxi- 
dation number  of  chromium  to  be  +3  and  the 
coordination  number  to  be  6  in  both  compounds. 
Estimate 

(a)  the  solubility  of  these  compounds  in  water; 

(b)  their  relative  melting  points; 

(c)  the  relative  conductivity  of  the  liquid  phases. 

7.  Why  does  NH3  readily  form  complexes,  but 
NH4+  does  not? 

8.  Place  a  piece  of  paper  over  Figure  15-13  and 
trace  it.  Extend  the  abscissa  and  add  the  ioniza- 


410 


THE    FOURTH-ROW    TRANSITION    ELEMENTS    I    CHAP.    22 


tion  energies  of  the  transition  elements.  Com- 
plete the  row  with  the  following  ionization 
energies:  Ga,  138;  Ge,  187;  As,  242;  Se,  225; 
Br,  273;  Kr,  322;  Rb,  96  kcal/mole. 

9.  The  volume  per  mole  of  atoms  of  some  fourth- 
row  elements  (in  the  solid  state)  are  as  follows: 
K,  45.3;  Ca,  25.9;  Sc,  18.0;  Br,  23.5;  and  Kr, 
32.2  ml/mole  of  atoms.  Calculate  the  atomic 
volumes  (volume  per  mole  of  atoms)  for  each  of 
the  fourth-row  transition  metals.  Plot  these 
atomic  volumes  and  those  of  the  elements  given 
above  against  atomic  numbers. 

10.  Chromic  oxide,  Cr203,  is  used  as  a  green  pigment 
and  is  often  made  by  the  reaction  between 
Na2Cr2CVsJ  and  NH4CIM  to  give  Cr203(sj, 
NaClfsJ,  N/gJ,  and  HzO(g).  Write  a  balanced 
equation  and  calculate  how  much  pigment  can 
be  made  from  1.0  X  102  kg  of  sodium  dichro- 
mate. 

11.  Chromic  hydroxide,  Cr(OH)3,  is  a  compound 
with  low  solubility  in  water.  It  is  usually  hy- 
drated  and  does  not  have  the  definite  composi- 
tion represented  by  the  formula.  It  is  quite 
soluble  either  in  strong  acid  or  strong  base. 

(a)  Write  an  equation  showing  the  ions  pro- 
duced by  the  small  amount  of  Cr(OH)3  that 
dissolves. 

(b)  Explain,  using  Le  Chatelier's  Principle,  why 
Cr(OH)3  is  more  soluble  in  strong  acid  than 
in  water. 

(c)  What  is  the  significance  of  the  fact  that 
Cr(OH)3  dissolves  in  base,  as  well  as  in  acid  ? 

12.  What  is  the  oxidation  number  of  manganese  in 
each  of  the  following:  MnCV (aq);  Mn+2(  aq); 
Mn5CVsJ;  MnCVsj;  Mn(OH)2fsJ;  MnCl2(sj; 
MnF/sj? 

13.  Manganese(III),  Mn+3( aq),  spontaneously  dis- 
proportionates  to  Mn+i(aq)  and  Mn02(s).  Bal- 
ance the  equation  for  the  reaction. 

14.  Use  the  E°  values  in  Table  22-IH  to  predict  what 
might  happen  if  a  piece  of  iron  is  placed  in  a 
1  M  solution  of  Mn+2  and  if  a  piece  of  manga- 
nese is  placed  in  a  1  M  solution  of  Fe+2.  Balance 
the  equation  for  any  reaction  that  you  feel  would 
occur  to  an  appreciable  extent. 


15.  Iron  exists  in  one  cubic  crystalline  form  at  20°C 
(body  centered  cubic,  with  cube  edge  length 
2.86  A)  and  in  another  form  at  1100°C  (face 
centered  cubic,  with  cube  edge  length  3.63  A). 

(a)  Draw  a  picture  of  each  unit  cell,  showing  the 
nine  atoms  involved  in  a  body  centered  cubic 
cell  and  the  fourteen  atoms  involved  in  a 
face  centered  cubic  cell.  (See  Figure  21-2.) 

(b)  Decide  the  number  of  unit  cells  with  which 
each  atom  is  involved  (in  each  structure). 

(c)  How  many  atoms  are  in  each  unit  cell  if  we 
take  into  account  that  some  atoms  are  shared 
by  two  or  more  adjoining  unit  cells? 

(d)  Calculate  the  volume  of  the  unit  cell  and, 
with  your  answer  to  part  (c),  the  volume  per 
atom  (for  each  structure). 

(e)  What  conclusion  can  be  drawn  about  the 
"effective  size"  of  an  iron  atom? 

16.  One  of  the  important  cobalt  ores  is 

Co3(As04)2-8H20 

How  much  of  this  ore  is  needed  to  make  1.0  kg 
of  Co? 

17.  Nickel  carbonyl,  Ni(CO)4,  boils  at  43°C,  and 
uses  the  sp3  orbitals  of  Ni  for  bonding.  Give 
reasons  to  justify  the  following: 

(a)  it  forms  a  molecular  solid; 

(b)  the  molecule  is  tetrahedral; 

(c)  bonding  to  other  molecules  is  of  the  van  der 
Waals  type; 

(d)  the  liquid  is  a  nonconductor  of  electricity; 

(e)  it  is  not  soluble  in  water. 

18.  Write  balanced  equations  to  show  the  dissolving 
of  Cu(OH)2(sJ  on  the  addition  of  NH3(aqJ,  and 
also  the  reprecipitation  caused  by  the  addition 
of  an  acid. 

19.  Cupric  sulfide,  copper(II)  sulfide,  reacts  with  hot 
nitric  acid  to  produce  nitric  oxide  gas,  NO,  and 
elemental  sulfur.  Only  the  oxidation  numbers  of 
S  and  N  change.  Write  the  balanced  equation  for 
the  reaction. 

20.  The  solubility  of  copper(II)  iodide,  Cul2,  is  0.004 
g/liter.  Determine  the  value  of  the  solubility 
product. 


CHAPTER 


Some  Sixth-  and 

Seventh-Row 

Elements 


La\Ce\  Pr  \Xd\pm  \  Sm  \Eu  \  Gd\  Tb\Vy\  Ho  \Er  \Tm\Yb\Lu\ 
Ac\Th\Pa\l/  \Vp  \Pu  \Am\On\Bk  \Cf\Es  \Fm  \M<il        \ 


103 

£w\ 


23-1    THE  SIXTH  ROW  OF  THE  PERIODIC  TABLE 


The  fifth-row  transition  elements  have  general 
similarity  to  the  fourth-row  transition  elements. 
The  electron  structure  is  essentially  the  same 
except  that  the  Ad  orbitals  are  filling  instead  of 
the  3d  orbitals.  Near  the  beginning  of  the  sixth- 
row  transition  elements  there  is  a  change:  the/ 
orbitals  begin  to  fill  to  form  fourteen  elements 
before  the  d  orbitals  can  be  occupied  to  give  the 
lypical  transition  elements.  This  chapter  will  dis- 


cuss these  fourteen  elements  and  some  of  the 
seventh-row  elements. 

23-1.1    The  Lanthanides,  or  Rare  Earths 

The  lanthanides,  or  rare  earths,  are  lanthanum 
and  the  fourteen  chemical  elements  following. 
These  fourteen  elements,  all  of  which  are  very 
similar  to   La  in  chemical   behavior,   include 

411 


412 


SOME    SIXTH-    AND    SEVENTH-ROW    ELEMENTS    \    CHAP.    23 


cerium,  praseodymium,  neodymium,  prome- 
thium,  samarium,  europium,  gadolinium,  ter- 
bium, dysprosium,  holmium,  erbium,  thulium, 
ytterbium,  and  lutetium  (atomic  numbers  58  to 
71).  These  elements  are  called  the  rare  earth 
elements  because  they  were  extracted  from  ox- 
ides, for  which  the  ancient  name  was  "earth," 
and  because  those  oxides  were  rather  rare. 
During  the  1940's,  techniques  for  separating 
elements  were  developed  to  such  a  degree  that 
the  rare  earth  elements  are  no  longer  so  rare. 
The  most  striking  property  of  these  elements  is 
that  their  chemical  properties  are  almost  identical. 
For  example,  they  are  all  reactive  metals  (about 
like  calcium).  They  react  with  water  to  give 
a  vigorous  evolution  of  hydrogen.  They  all  form 
basic  trihydroxides  which  are  only  slightly  solu- 
ble in  water  but  readily  soluble  in  acid. 
For  example, 

2La(s)  +  6H20  — ►■ 

2La(OH)3(sj  +  m2(g)    (/) 

La(OH)3(sj  +  m+(aq)  z<=t  La+3(aq)  +  3H20    (2) 
And,  in  similar  reactions, 

2Ce(sj  +  6H.O  — >- 

2Ce(OH)3(s)  +  3H2(g)    (3) 

Ce(OH)3(sj  +  3H+(aq)  z<=±:  Ce+3(aq)  +  3H20    (4) 


2Pr(s)  +  6H.O  — ►- 

2Pr(OH)/s>)  +  mjg)     (5) 
Pr(OHh(s)  +  2H+(aq)  +±:  Pr+3f aq)  +  3H.O     (6) 

The  reason  usually  cited  for  the  great  simi- 
larity in  the  properties  of  the  lanthanides  is  that 
they  have  similar  electronic  configurations  in  the 
outermost  65  and  5d  orbitals.  This  occurs  be- 
cause, at  this  point  in  the  periodic  table,  the 
added  electrons  begin  to  enter  Af  orbitals  which 
are  fairly  deep  inside  the  atom.  These  orbitals 
are  screened  quite  well  from  the  outside  by  outer 
electrons,  so  changing  the  number  of  4/ electrons 
has  almost  no  effect  on  the  chemical  properties 
of  the  atom.  The  added  electrons  do  not  become 
valence  electrons  in  a  chemical  sense — neither 
are  they  readily  shared  nor  are  they  readily  re- 
moved. 

The  very  slight  differences  that  do  exist  among 
these  elements  are  due  to  small  changes  in  size 
brought  about  by  increase  of  nuclear  charge.  The 
separation  of  the  lanthanide  elements  from  each 
other  is  based  upon  clever  exploitation  of  these 
slight  differences  in  properties.  Table  23-1  shows 
a  comparison  of  some  of  the  properties  of  the 
various  lanthanide  elements.  As  can  be  seen, 
+3  is  the  common  oxidation  number  and  is  most 
characteristic  of  the  chemistry  of  these  elements. 
Another  thing  to  note  is  the  steady  decrease  in 


Table  23-1.     some  properties  of  lanthanum  and  the  lanthanide  elements 


OUTER  ELECTRON 

OXIDATION 

E° 

+  3  ION 

ELEMENT 

Z 

CONFIGURATION 

STATES 

M  — >-  M+3  +  3e~ 

RADIUS 

La 

57 

54' 

6s2 

+3 

2.52  volts 

1.15  A 

Ce 

58 

Af  5d 

6s2 

+3, 

+4 

2.48 

1.11 

Pr 

59 

4/s 

6s2 

+  3, 

+4 

2.47 

1.09 

Nd 

60 

Af 

6s2 

+3 

2.44 

1.08 

Pm 

61 

4f 

6s2 

+3 

2.42 

1.06 

Sm 

62 

Af* 

6s2 

+2, 

+3 

2.41 

1.04 

Eu 

63 

Af 

6s2 

+2 

+3 

2.41 

1.03 

Gd 

64 

Af  5d 

6s2 

+3 

2.40 

1.02 

Tb 

65 

Af 

6s2 

+3, 

+4 

2.39 

1.00 

Dy 

66 

Af0 

6s2 

+3 

2.35 

0.99 

Ho 

67 

Afn 

6s2 

+3 

2.32 

0.97 

Er 

68 

Af1 

6s2 

+3 

2.30 

0.96 

Tm 

69 

Af3 

6s2 

+3 

2.28 

0.95 

Yb 

70 

Af* 

6s2 

+2, 

+3 

2.27 

0.94 

Lu 

71 

Af*5d 

'6i2 

+3 

2.25 

0.93 

SEC.    23-2       THE    SEVENTH    ROW    OF    THE    PERIODIC    TABLE 


413 


the  M+3  ionic  size,  shown  in  the  last  column. 
This  decrease  is  called  the  "lanthanide  contrac- 
tion" and  is  due  to  the  fact  that  the  nuclear 
charge  increases  through  the  series  (and,  con- 
sequently, so  is  the  attraction  for  the  electrons 
increased),  while  the  added  electrons  are  not 
entering  outer  orbitals  where  they  would  tend  to 
increase  atomic  size. 


EXERCISE  23-1 

(a)  Balance  the  equation  for  the  reaction  be- 
tween neodymium  metal  and  chlorine  gas. 

(b)  At  0°C  and  one  atmosphere  pressure,  how 
many  liters  of  chlorine  gas  would  react  with 
14.4  grams  of  the  metal? 

EXERCISE  23-2 

To  a  solution  containing  "tracer"  amounts  of 
radioactive  gadolinium,   Gd~3(aq)  (concentra- 


tion less  than  10~12  M)  is  added  0.01  M  lantha- 
num chloride,  LaCl3,  and  0.1  M  hydrogen  fluo- 
ride, HF.  A  precipitate  of  lanthanum  fluoride 
forms  and  it  contains  most  of  the  radioactive 
gadolinium.  Explain  in  terms  of  the  similarity  of 
lanthanum  and  gadolinium  ions. 


23-1.2    Occurrence  and  Preparation 

The  most  important  minerals  of  the  lanthanide 
elements  are  monazite  (phosphates  of  La,  Ce, 
Pr,  Nd  and  Sm,  as  well  as  thorium  oxide)  plus 
cerite  and  gadolinite  (silicates  of  these  elements). 
Separation  is  difficult  because  of  the  chemical 
similarity  of  the  lanthanides.  Fractional  crystal- 
lization, complex  formation,  and  selective  ad- 
sorption and  elution  using  an  ion  exchange  resin 
(chromatography)  are  the  most  successful  meth- 
ods. 


23-2    THE  SEVENTH-ROW  OF  THE  PERIODIC  TABLE 


The  last  row  of  the  periodic  table  is  unique  in 
that  all  of  these  elements  have  radioactive  nuclei. 
However,  this  property  has  no  direct  effect  on 
the  chemistry  of  these  elements.  The  nuclear 
charge  and  nuclear  mass  affect  the  chemistry  of 
an  atom  but  the  fact  that  the  nucleus  might  ex- 
plode at  any  moment  (that  is,  undergo  some  sort 
of  radioactive  decay)  isn't  known  by  the  elec- 
trons surrounding  the  nucleus.  We  might  com- 
pare the  situation  to  the  packing  of  eggs  in  egg 
crates.  How  many  eggs  can  be  placed  in  an  egg 
crate  is  fixed  entirely  by  their  shape  and  size.  It 
is  immaterial  that  eggs  have  the  additional  prop- 
erty that  they  can  hatch  if  fertilized.  The  packing 
of  fertile  eggs  is  also  fixed  entirely  by  their  shape 
and  size,  so  a  crate  of  fertile  eggs  is  indistinguish- 
able from  a  crate  of  infertile  eggs  (unless  one  of 
the  fertile  eggs  happens  to  hatch  while  you  are 
looking  at  them).  In  exactly  the  same  way,  the 
chemistry  of  an  atom  has  nothing  to  do  with  the 
stability  or  instability  of  its  nucleus. 

Nevertheless,  there  is  special  interest  in  the 
chemistry  of  the  seventh-row  elements,  a  special 


interest  directly  tied  to  the  nuclear  instability.  As 
we  have  mentioned  in  Chapter  7,  the  energy 
contents  of  nuclei  are  many  orders  of  magnitude 
higher  than  chemical  heat  contents.  If  this  nu- 
clear energy  content  can  be  tapped  and  put  to 
work,  enormous  quantities  of  energy  become 
available.  Thus  far.  the  radioactive  nuclei  of 
some  of  these  seventh-row  elements  have  been 
most  useful  in  our  attempts  to  harness  nuclear 
energy. 

Nuclear  fuels,  like  chemical  fuels,  must  be 
purified  to  be  most  effective.  The  purification  of 
the  seventh-row  elements  has  presented  some 
fascinating  and  difficult  problems  of  chemistry — 
so  difficult,  in  fact,  that  chemists  have  played  as 
big  a  role  in  the  development  of  nuclear  energy 
as  have  physicists. 

23-2.1    The  Occurrence  of  the 
Seventh-Row  Elements 

Only  five  of  the  seventh-row  elements  are  found 
in  nature:  radium,  actinium,  thorium,  protac- 


414 


SOME    SIXTH-    AND    SEVENTH-ROW    ELEMENTS    I    CHAP.    23 


tinium,  and  uranium.  It  is  one  of  the  significant 
advances  of  science  in  the  first  half  of  this  cen- 
tury that  ten  additional  elements  of  this  row  have 
been  synthesized:  francium,  neptunium,  pluto- 
nium,  americium,  curium,  berkelium,  califor- 
nium, einsteinium,  fermium,  and  mendelevium. 
The  methods  for  raising  the  nuclear  charge  of  a 
nucleus  require  ingenious  applications  of  phys- 
ics. However,  the  synthesis  of  these  ten  elements 
could  never  have  been  demonstrated  if  it  had  not 
been  for  the  solution  of  some  difficult  problems 
of  chemistry,  worked  on  by  some  of  the  most 
highly  skilled  chemists  in  the  world.  At  this  time, 
elements  102  and  103  have  been  prepared  and 
the  preparation  of  elements  beyond  is  under 
study. 

23-2.2    The  Elements  Following  Actinium 

The  most  interesting  elements  of  the  seventh  row 
are  those  following  actinium.  For  some  of  these 
elements  a  large  amount  of  chemistry  is  known. 
The  first  four,  actinium,  thorium,  protactinium, 
and  uranium,  used  to  be  shown  in  the  periodic 


table  under  lanthanum,  hafnium,  tantalum,  and 
tungsten,  since  they  resemble  these  elements  in 
many  chemical  reactions.  With  the  discovery  that 
one  could  make  elements  beyond  uranium  (those 
of  higher  Z  than  92),  it  was  suggested  that  there 
might  be  a  5/ series  of  elements  analogous  to  the 
4/  series  of  rare  earths  (the  lanthanides).  As  a 
result,  it  has  become  customary  to  place  the  ele- 
ments following  actinium  under  the  rare  earths 
(see  inside  front  cover)  and  to  call  these  elements 
the  "actinides."  Table  23-11  collects  information 
concerning  the  known  oxidation  states  of  the 
elements  following  lanthanum,  those  following 
lutetium  (that  is,  beginning  with  hafnium),  and 
those  following  actinium. 

The  most  striking  feature  of  the  contrasts 
shown  in  Table  23-11  is  that  the  seventh-row 
elements  display  the  multiplicity  of  oxidation 
states  characteristic  of  transition  elements  rather 
than  the  drab  chemistry  of  the  +3  rare  earth 
ions.  Whereas  Ce+3(aq)  can  be  oxidized  to 
Ce+i(aq)  only  with  an  extremely  strong  oxidizing 
agent,  Th+YagJ  is  the  stable  ion  found  in  tho- 
rium salts  and  7h+3(aq)  is  unknown.  In  a  similar 


Table  23-11.     oxidation  numbers  found  for  some  sixth-row  and 

SEVENTH-ROW    ELEMENTS 


RARE  EARTH  ELEMENTS 

SIXTH-ROW  ELEMENTS 

SEVENTH-ROW  ELEMENTS 

FOLLOWING 

LANTHANUM 

FOLLOWING  LUTETIUM 

FOLLOWING  ACTINIUM 

Oxidation 

Oxidation 

Oxidation 

Name 

Numbers* 

Name 

Numbers 

Name 

Numbers 

lanthanum 

+  3 

lutetium 

-3 

actinium 

+3 

cerium 

+3. +4 

hafnium 

-4 

thorium 

+4 

praseodymiurr 

i 

+  3. +4 

tantalum 

+5 

protactinium 

+4,  +5 

neodymium 

+3 

tungsten 

+2,  +3,  +4,  +5, 

+6 

uranium 

+3,  +4,  +5,  +6 

promethium 

+3 

rhenium 

+  3,  ~4.  +5, 

+6,  +7 

neptunium 

+  3,  +4,  +5.  +6 

samarium 

+2, 

+3 

osmium 

+2, +3,  +4, 

+6           +8 

Plutonium 

+3,  +4. +5,  +6 

europium 

+2, 

+3 

iridium 

-3.  +4, 

+6 

americium 

+3,  +4,  +5,  +6 

gadolinium 

+  3 

platinum 

-2.          +4 

curium 

+3 

terbium 

+  3. +4 

gold 

+  1, 

-3 

berkelium 

+  3. +4 

dysprosium 

+3 

mercury 

+  1. 

+2 

californium 

+3 

holmium 

+3 

thallium 

+1, 

+3 

einsteinium 

+  3 

erbium 

+3 

lead 

+2,         +4 

fermium 

+  3 

thulium 

+3 

bismuth 

+  3.          +5 

mendelevium 

+3 

ytterbium 

+2, 

+3 

polonium 

+2,          +4, 

+6 

102 

lutetium 

+3 

astatine 

(-1) 

103 

•The  most  common  oxidation  numbers  are  in  blue  type. 


SEC.    23-2    I    THE    SEVENTH    ROW    OF    THE    PERIODIC    TABLE 


415 


contrast,  the  most  important  oxidation  state  of 
protactinium  is  +5  whereas  no  compound  in- 
volving the  +5  oxidation  state  of  praseodymium 
is  known. 

Yet  it  is  generally  accepted  today  that  the 
elements  following  actinium  involve  5/  energy 
levels,  identifying  them  as  the  seventh-row  equiv- 
alents of  the  lanthanides. 

Spectroscopic  investigations  of  the  gaseous 
atoms  and  of  ions  in  suitable  crystals  furnish 
one  of  the  bulwarks  of  this  view.  Table  23-111 
contrasts  the  actinide  electronic  configurations 
as  they  are  known  today  with  the  configurations 
of  the  corresponding  rare  earths.  Although  the 
corresponding  configurations  are  not  identical 
they  are  quite  similar.  They  offer  no  hint  of 
explanation  for  the  varied  actinide  oxidation 
states  and  the  monotonous  similarity  of  the 
lanthanides.  It  may  be  that  one  of  the  most 
significant  conclusions  to  be  drawn  from  the 
investigations  of  the  actinides  is  that  elements  of 
an  /  orbital  transition  series  do  not  necessarily 
resemble  each  other  strongly  in  chemical  be- 
havior, contrary  to  what  had  been  inferred  from 
studies  of  the  rare  earths. 


Table  23-III 

THE    ELECTRON    CONFIGURATIONS 
OF    GASEOUS    ACTINIDE    AND 
LANTHANIDE    ATOMS 

LANTHANIDES  ACTINIDES 


La 

5dl 

6s2 

Ac 

to1 

7s2 

Ce 

4/1 

Sd> 

6s2 

Th 

6cP 

7s2 

Pr 

4/3 

6s2 

Pa 

5P 

6dl 

7s2 

Nd 

V 

6s2 

U 

5P 

to1 

7s2 

Pm 

4/6 

6s2 

Np 

5f* 

&/> 

7s2 

Sm 

4/8 

6s2 

Pu 

5P 

7s2 

Eu 

4f 

6s2 

Am 

sp 

7s2 

Gd 

4f 

5dl 

6s2 

Cm 

5P 

&/> 

7s2 

EXERCISE  23-3 

Two  half-reactions  involving  neptunium  are 
Np+3  — >-  Np+4  +  e~ 

E°  =  -0.147  volt    (7) 


Np+<  +  2H.O  — >-  Np02+  +  4H+  +  e~ 

E°  =  -0.75  volt      (8) 

Find  an  oxidizing  agent  in  Appendix  3  that  could 
oxidize  Np+3(aq)  to  Np+i(aq)  but  not  to 
Np02+  (aq)., 

EXERCISE  23-4 

Balance  the  equation  for  the  reaction  between 
permanganate,  MnO^aq),  and  plutonium(III), 
Pu+3(aq),  to  form  manganous,  Mn+2(aq),  and 
plutonyl  ion,  Pu0^2faqj,  in  acid  solution. 

EXERCISE  23-5 

Plutonium(IV),  Pu+YagJ,  forms  a  complex  ion 
with  fluoride  ion,  PuF+3: 


PuF^ 


pU+4  _|_  f- 


K  =  1.6  X  10"7    (9) 


If  the  F~  concentration  is  adjusted  to  0.10  M  in 
a  solution  containing  1.0  X  10-3  M  Pu+4,  calcu- 
late the  ratio  of  the  concentration  of  Pu"1"4  to 
PuF+3.  What  is  the  equilibrium  concentration  of 
Pu+4? 


Chemical  investigation  of  the  actinides  is  made 
difficult  by  the  extreme  instability  of  the  nuclei — 
an  instability  that  increases  as  we  go  to  higher 
atomic  numbers.  This  leads  to  intense  radioac- 
tivity (which  makes  shielding  precautions  neces- 
sary) and  small  amounts  of  material  to  work 
with.  The  heavier  actinides  are  usually  made  by 
bombarding  the  lighter  actinides  with  neutrons, 
helium  nuclei  (alpha  particles),  or  even  carbon 
nuclei.  With  very  unstable  nuclei,  it  is  hard  to 
accumulate  enough  atoms  to  permit  chemical 
studies  on  a  macroscopic  scale  since  the  element 
is  disintegrating  at  the  same  time  it  is  being 
produced.  The  consequence  is  that  much  of  the 
chemical  investigation  is  done  by  tracer  tech- 
niques wherein  chemical  reactions  are  performed 
on  a  bulk  scale  with  some  other,  more  available, 
element.  Clues  to  the  chemistry  of  the  trace  ele- 
ment are  furnished  by  the  behavior  of  the 
radioactivity — whether  it  follows  along  with  the 
bulk  carrier  or  is  lost  in  the  chemical  operations. 

There  is  reason  to  believe  that  not  many  more 
new  elements  will  be  created.  The  instability  of 
the  nuclei  of  the  last  few  elements  suggests  that 


416 


SOME    SIXTH-    AND    SEVENTH-ROW    ELEMENTS    I    CHAP.    23 


most  or  all  of  the  still  higher  elements  have 
nuclei  so  unstable  that  they  will  not  survive  long 
enough  to  be  detected. 

23-2.3    Nuclear  Stability  and  Radioactivity 

The  outstanding  characteristic  of  the  actinide 
elements  is  that  their  nuclei  decay  at  a  measur- 
able rate  into  simpler  fragments.  Let  us  examine 
the  general  problem  of  nuclear  stability.  In 
Chapter  6  we  mentioned  that  nuclei  are  made  up 
of  protons  and  neutrons,  and  that  each  type  of 
nucleus  can  be  described  by  two  numbers:  its 
atomic  number  (the  number  of  protons),  and  its 
mass  number  (the  sum  of  the  number  of  neu- 
trons and  protons).  A  certain  type  of  nucleus  is 
represented  by  the  chemical  symbol  of  the  ele- 
ment, with  the  atomic  number  written  at  its 
lower  left  and  the  mass  number  written  at  its 
upper  left.  Thus  the  symbol 


represents  the  nucleus  of  the  plutonium  isotope 
which  contains  94  protons  and  239  —  94  =  145 
neutrons.  Since  the  forces  that  exist  in  the  nu- 
cleus depend  upon  the  number  of  protons  and 
neutrons  of  which  it  is  composed,  we  can  get  a 
rough  idea  of  whether  a  nucleus  is  stable  just 
by  examining  these  two  numbers. 

First  of  all,  let  us  define  what  we  mean  by 
stability.  Consider  an  initially  pure  sample  of 
292U.  Regardless  of  the  physical  or  chemical 
state  in  which  we  find  the  uranium  atoms,  some 
of  them  will  decay  each  instant  to  become 
thorium  atoms  by  the  spontaneous  reaction: 


*1  J 


42He  +  2^Th 


UO) 


Notice  that  both  the  electric  charge  and  the 
total  number  of  nuclear  particles  (nucleons)  are 
conserved  in  the  nuclear  decomposition.  Careful 
study  of  the  rate  of  this  nuclear  decay  shows  that 
in  a  given  period  of  time  a  constant  fraction  of 
the  nuclei  present  will  undergo  decomposition. 
This  observation  allows  us  to  characterize  or 
describe  the  rate  of  nuclear  decay  in  a  very  simple 
manner.  We  simply  specify  the  length  of  time  it 
takes  for  a  fixed  fraction  of  the  nuclei  initially 
present  to  decay.  Normally  we  pick  the  time  for 


half  the  nuclei  to  decay;  this  length  of  time  is 
known  as  the  half-life  of  the  nucleus.  For  ex- 
ample, measurements  show  that  after  4.5  X  10° 
years,  half  the  atoms  in  any  sample  of  2H\J  will 
decay  to  2goTh.  A  nucleus  is  considered  to  be 
stable  if  its  half-life  is  much  longer  than  the  age 
of  the  earth,  which  is  about  5  X  109  years. 
Nuclei  that  are  very  unstable  are  characterized 
by  half-lives  which  are  quite  short — in  some 
cases  only  a  small  fraction  of  a  second. 

Now  let  us  turn  to  the  problem  of  how  the  composition 
of  a  nucleus  affects  its  stability.  The  forces  that  exist 
between  the  particles  in  the  nucleus  are  very  large.  The 
most  familiar  of  the  intranuclear  forces  is  the  coulomb 
force  of  repulsion  which  the  protons  must  exert  on  one 
another.  In  order  to  appreciate  the  magnitude  of  this 
repulsive  force,  let  us  compare  the  force  between  two 
protons  when  they  are  separated  by  10-8  cm,  as  they  are 
in  the  hydrogen  molecule,  with  the  force  between  two 
protons  separated  by  10-13  cm,  as  they  are  in  a  helium 
nucleus.  In  the  first  case  we  have 


force 


(charge  on  proton  1  Xcharge  on  proton  2) 


(distance  between  protons)2 


(+gX+f)       _jl 
(10~s)2         io- 


1016e2 


un 


where  e  is  the  electric  charge  on  one  proton.   In  the 
second  case  we  have 


force  = 


(+e)(+g)  =  JL_ 
(10"13)2         10-20 


=  10-6<?2 


U2) 


By  comparing  these  two  answers  we  can  see  that  the 
repulsive  force  between  two  protons  in  the  nucleus  is 
about  ten  billion  times  as  great  as  the  repulsive  force 
between  two  protons  bound  together  in  a  hydrogen  mole- 
cule. In  order  to  overcome  these  enormous  intranuclear 
coulomb  repulsions  and  hold  the  nucleus  together  there 
must  exist  some  very  strong  attractive  forces  between  the 
nucleons.  The  nature  of  these  forces  is  not  understood 
and  remains  a  very  important  problem  in  physics. 

It  is  relatively  easy  to  summarize  how  nuclear 
stability  (and  hence  the  attractive  nuclear  forces) 
depends  upon  the  numbers  of  protons  and  neu- 
trons in  the  nucleus.  For  atoms  with  atomic 
number  less  than  20,  the  most  stable  nuclei  are 
those  in  which  there  are  equal  numbers  of  pro- 
tons and  neutrons.  For  atoms  with  atomic  num- 
bers between  20  and  83,  the  most  stable  nuclei 
have  more  neutrons  than  protons.  For  atoms  of 
atomic  number  greater  than  83,  no  nucleus  can 
be  considered  stable  by  our  definition.  These 


SEC.    23-2    I    THE    SEVENTH    ROW    OF    THE    PERIODIC    TABLE 


417 


loo 


Number    of  protons ,   P 


Fig.  23-1.  The  relation  between  number  of  neutrons 
and  protons  in  stable  nuclei.  Each  dot 
identifies  a  stable  nucleus. 


statements  are  demonstrated  in  Figure  23-1, 
where  the  number  of  neutrons  is  plotted  against 
the  number  of  protons  for  all  stable  nuclei.  We 
see  that  these  stable  nuclei  form  a  belt  which 
deviates  increasingly  from  a  neutron  to  proton 
ratio  of  unity  (the  dashed  line,  TV  =  P)  as  the 
charge  on  the  nucleus  increases.  Nuclei  whose 
neutron-proton  ratio  is  such  that  they  lie  outside 
of  this  belt  of  stability  are  radioactive. 


23-2.4    Types  of  Radioactivity 

There  are  three  common  ways  by  which  nuclei 
can  approach  the  region  of  stability:  (1)  loss  of 
alpha  particles  (a-decay);  (2)  loss  of  beta  parti- 
cles (j3-decay);  (3)  capture  of  an  orbital  electron. 
We  have  already  encountered  the  first  type  of 
radioactivity,  a-decay,  in  equation  (10).  Emis- 
sion of  a  helium  nucleus,  or  alpha  particle,  is  a 
common  form  of  radioactivity  among  nuclei 
with  charge  greater  than  82,  since  it  provides  a 
mechanism  by  which  these  nuclei  ean  be  con- 
verted to  new  nuclei  of  lower  charge  and  mass 
which  lie  in  the  belt  of  stability.  The  actinides,  in 
particular,  are  very  likely  to  decay  in  this  way. 


418 


SOME    SIXTH-    AND    SEVENTH-ROW    ELEMENTS    I    CHAP.    23 


For  example,  the  most  stable  isotope  of  element 
100,  fermium,  has  a  half-life  of  only  4.5  days: 

?$Fm  — >-  2HCf  +  *He  (13) 

Nuclei  that  have  a  neutron-proton  ratio  which 
is  so  high  that  they  lie  outside  the  belt  of  stable 
nuclei  often  decay  by  emission  of  a  negative 
electron  (a  beta  particle)  from  the  nucleus.  This 
effectively  changes  a  neutron  to  a  proton  within 
the  nucleus.  Two  examples  are 


281Cf 
?JNa 


23*Es    +  _?<? 
2^Mg  +  _°e 


(14) 
(15) 


This  type  of  radioactivity,  /3-decay,  is  found  in 
both  the  light  and  heavy  elements. 

Nuclei  that  have  too  many  protons  relative  to 
their  number  of  neutrons  correct  this  situation 
in  either  of  two  ways.  They  either  capture  one 
of  their  Is  electrons  or  they  emit  a  positron 
(a  positively  charged  particle  with  the  same  mass 
as  an  electron).  Either  process  effectively  changes 
a  proton  to  a  neutron  within  the  nucleus. 

23-2.5    Nuclear  Energy 

In  the  previous  section  we  saw  that  the  stability 
of  a  nucleus  is  affected  by  its  neutron/proton 
ratio.  Even  among  those  nuclei  that  we  consider 
stable,  however,  there  is  a  variation  in  the  forces 
which  hold  the  nucleus  together.  In  order  to 
study  this  variation  in  nuclear  binding  energy, 
let  us  consider  the  process  of  building  a  nucleus 
from  protons  and  neutrons.  For  an  example,  let 
us  look  at  the  hypothetical  reaction 

2  }H  +  2  In  — ■*■  |He  (16) 

First  we  will  compare  the  masses  of  the  reactants 
with  those  of  the  product: 


30  160 

fvfass      number 


24-0 


Fig.  23-2.  The    binding    energy    per   particle    in    the 
nucleus. 


amount  of  energy  must  be  released  (see  Section 
7-4.1).  By  the  Einstein  mass-energy  relationship, 
E  =  mc2,  we  can  calculate  that  in  the  formation 
of  1  mole  of  helium  nuclei  there  would  be 

E  =  mc1  =  (o.03  -^-\  X  (3  X  1010  —  Y 
\         mole/       \  sec/ 

=  2.7  X  1019     *cm\ 
sec2  mole 


"( 


E  =  I  2.7  X  101 


sec2  mole, 


=  6.4  X  10+11 


X     2.4  X  10-8 


calories 


calories   ' 
g  cmVsec2, 


mole 


(17) 


or  640  billion  calories  of  energy  released.  This 
amount  of  energy  can  be  considered  as  the  bind- 
ing energy  of  a  mole  of  helium  nuclei  since  this 
much  energy  must  be  supplied  to  dissociate  a 
mole  of  ^He  into  two  moles  each  of  protons  and 
neutrons. 

Similar  calculations  can  be  made  for  other 
nuclei.  A  significant  comparison  between  nu- 
clear binding  energies  can  be  made  if  we  divide 
the  total  binding  energy  of  each  nucleus  by  the 


Mass  of  2  protons    =  2  X  1.00759  -  2.01518  g  mole 
Mass  of  2  neutrons  =  2  X  1.00897  =  2.01794  g  mole 


Total  mass  of  reactants 
Mass  of  ^He 

Mass  difference  between 
products  and  reactants 

Since  there  is  a  decrease  of  0.03035  gram/mole 
of  helium  formed  in  this  reaction,  an  equivalent 


4.03312  g /mole 
4.00277  g/mole 


0.03035  g/mole 


number  of  nucleons  in  the  nucleus.  This  calcula- 
tion provides  us  with  the  binding  energy  per 


SEC.    23-2    I    THE    SEVENTH    ROW    OF    THE    PERIODIC    TABLE 


419 


particle  in  the  nucleus.  The  binding  energy  per 
particle  varies  in  a  systematic  way  as  the  mass 
number  of  the  nucleus  increases.  This  variation 
is  shown  in  Figure  23-2. 

The  nuclei  that  have  a  mass  number  of  ap- 
proximately 60  have  the  highest  binding  energy 
per  nuclear  particle,  and  are  therefore  the  most 
stable  nuclei.  This  graph  helps  us  to  understand 
the  existence  of  the  processes  of  nuclear  fission 
and  nuclear  fusion.  If  the  nuclei  of  the  heavier 
elements  such  as  uranium  and  plutonium  are 
split  into  two  smaller  fragments,  the  binding 
energy  per  nucleon  is  greater  in  the  lighter  nuclei. 
As  in  every  other  reaction  in  which  the  products 
are  more  stable  than  the  reactants,  energy  is 
evolved  by  this  process  of  nuclear  fission.  Gener- 
ally this  fission  reaction  is  induced  by  the  bom- 
bardment of  a  particular  isotope  of  uranium  or 
plutonium  with  neutrons, 

In  +  239|U  — »-  >&Ba  +  39|Kr  +  3  J/i       (18) 

The  two  nuclei  on  the  right  side  are  just  two  of 
the  many  possible  products  of  the  fission  process. 
Since  more  than  one  neutron  is  released  in  each 
process,  the  fission  reaction  is  a  self-propagating, 
or  chain  reaction.  Neutrons  released  by  one 
fission  event  may  induce  other  fissions.  When 
fission  reactions  are  run  under  controlled  condi- 
tions in  a  nuclear  reactor,  the  energy  released  by 


the  fission  process  eventually  appears  as  heat. 
The  energy  released  by  the  fission  of  one  pound 
of  *yu  is  equivalent  to  that  obtained  from  more 
than  1000  tons  of  coal. 

Figure  23-2  shows  that  when  very  light  nuclei 
such  as  J H  or  ,H  are  brought  together  to  form 
heavier  elements,  the  binding  energy  per  nucleon 
again  increases  and  energy  is  released.  The  graph 
also  shows  that  the  energy  released  per  nucleon 
(therefore  per  gram  of  reactant)  is  considerably 
greater  in  the  fusion  process  than  in  the  fission 
reaction.  By  use  of  a  set  of  reactions  in  which 
four  protons  are  converted  into  a  helium  nucleus 
and  two  electrons,  one  pound  of  hydrogen  could 
produce  energy  equivalent  to  that  obtained  from 
10,000  tons  of  coal.  For  this  reason,  and  because 
of  the  great  abundance  of  hydrogen,  fusion  re- 
actions are  potentially  sources  of  enormous 
amounts  of  energy.  Unfortunately  fusion  reac- 
tions take  place  rapidly  only  at  temperatures 
that  are  greater  than  a  million  degrees.  These 
temperatures  have  been  attained  briefly  by  use 
of  nuclear  fission  explosions.  At  present,  at- 
tempts are  being  made  to  attain  the  temperatures 
required  for  nuclear  fusion  by  less  destructive 
means,  so  that  these  reactions  can  be  used  as  an 
energy  source.  The  most  promising  processes  are 
based  upon  the  fusion  of  deuterium,  ,H,  or 
tritium,  jH,  nuclei. 


One  of  the  major  advances  of  science  in  the  first  half  of 
this  century  was  the  synthesis  of  ten  elements  beyond  ura- 
nium. Glenn  T.  Seaborg  participated  in  the  discovery  oj 
most  of  these,  a  sufficient  tribute  to  his  outstanding  ability 
as  a  scientist.  For  the  first  such  discoveries,  those  of  nep- 
tunium and  plutonium,  he  shared  with  Professor  Edwin  M. 
McMillan  the  Nobel  Prize  in  Chemistry  for  1951. 

Seaborg  rose  from  humble  surroundings.  He  was  born 
in  a  small  mining  town,  Ishpeming,  Michigan.  During  his 
boyhood,  his  Swedish- American  parents  moved  to  southern 
California.  There,  two  important  trails  began  to  become 
apparent — his  high  intellect  and  his  unlimited  willingness 
to  work.  Attending  the  University  of  California  at  Los 
Angeles,  he  worked  his  way  through,  first  as  a  stevedore, 
then  picking  apricots,  then  as  a  linotype  apprentice.  By  his 
sophomore  year  Seaborg  had  distinguished  himselj  enough 
to  be  appointed  a  laboratory  assistant.  This  employment 
continued  through  his  undergraduate  career  and  it  providea 
him  his  first  opportunity  to  engage  in  research. 

For  graduate  study,  Seaborg  moved  to  the  Berkeley 
campus  of  the  University  of  California.  He  received  the 
Ph.D.  in  1937  and,  after  two  years  of  work  with  G.  N. 
Lewis,  began  work  on  the  attempted  preparation  of  elements 
beyond  uranium.  This  research,  following  earlier  work  oj 
E.  M.  McMillan,  culminated  rather  rapidly  in  the  discovery 
of  plutonium.  During  the  war  years,  Seaborg  had  a  key 
role  in  working  out  the  chemical  processes  used  to  extract 
and  purify  plutonium.  Then,  after  the  war,  he  relumed  with 
full  force  to  the  discovery  of  new  elements.  There  grew 
under  his  direction  a  large  portion  of  the  now-famous 
University  of  California  Radiation  Laboratory.  With  Sea- 
borgs  stimulation  and  guidance,  highly  skilled  research 
teams  prepared,  many  for  the  first  time,  americium, 
curium,  berkelium,  californium,  einsteinium,  fermium.  and 
mendelevium.  More  elements  are  to  come  from  this  excit- 
ing laboratory. 

In  1958,  Glenn  T.  Seaborg  was  asked  to  become  Chan- 
cellor of  the  Berkeley  campus  of  the  University  of  Cali- 
fornia, a  world  center  of  learning.  Two  years  later  his 
talents  were  demanded  for  national  service:  Seaborg  was 
appointed  Chairman  of  the  U.S.  Atomic  Energy  Commis- 
sion. Thus  he  brings  his  deep  knowledge  of  science  and  of 
scientists  to  the  service  of  society  in  an  age  when  society  has 
desperate  need  for  guidance  by  scholars  oj  this  preeminent 
stature. 


243;    [244) 

COMPANY 


CHAPTER 


Some  Aspects 
of  Biochemistry: 
An  Application 
of  Chemistry 


•  •  •  the  elucidation  of  the  structure  of  biochemicals  •  •  •  vv/7/  neces- 
sarily lead  to  a  deeper  understanding  of  their  junction,  and  thus  finally 
to  the  understanding  of  the  mechanism  of  life  itself. 

ALBERTE    PULLMAN    AND    BERNARD    PULLMAN,    1962 


Living  organisms — bacteria,  fungi,  mosses,  al- 
gae, plants,  animals — are  highly  organized  sys- 
tems of  chemical  compounds.  All  organisms 
derive  the  energy  for  their  activities,  and  produce 
the  substances  of  which  they  are  built,  by  means 
of  chemical  reactions. 

A  century  and  a  half  ago  men  regarded  the 
chemistry  of  living  organisms  as  something  quite 
distinct  from  the  chemistry  of  rocks,  minerals, 
and  other  nonliving  things.  Indeed,  there  was  in 
their  minds  at  that  time  the  inclination  to  believe 
that  living  things  were  imbued  with  some  mys- 
terious "vital  force"  that  was  beyond  the  power 
of  men  to  define  and  understand. 

As  time  went  on,  it  became  apparent  that  the 


mystery  in  the  chemistry  of  living  things  was  due 
to  ignorance  of  the  details  of  what  went  on,  and 
with  an  increased  understanding  of  chemical 
principles,  the  mystery  gradually  began  to  dis- 
appear. Compounds  that  were  earlier  known 
only  as  the  products  o!  plants  and  animals  were 
produced  in  the  laboratory  from  ordinary  in- 
organic substances.  By  the  middle  of  the  nine- 
teenth century  the  superstitious  belief  in  a  chemi- 
cal "vital  force"  had  disappeared,  and  now  there 
are  few  chemists  who  believe  that  the  chemistry 
of  living  organisms  is  beyond  the  power  of  men 
to  understand. 

We  still,  however,  mark  off  a  large  area  of 
chemical  study  by  the  term  "biochemistry."  This 

421 


422 


SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY    I    CHAP.    24 


is  not  because  biochemistry  is  fundamentally 
different  from  chemistry  in  general.  It  is  because 
in  order  for  a  chemist  to  use  his  talents  effectively 
to  solve  certain  kinds  of  problems  he  must  devote 
special  (but  not  exclusive)  attention  to  what  is 
known  about  a  particular  field  of  knowledge. 
Biochemists  are  chiefly  concerned  with  the 
chemical  processes  that  go  on  in  living  organ- 


isms. These  scientists  must  use  information  from 
all  branches  of  chemistry  to  answer  the  questions 
they  ask,  but  their  questions  are  usually  some- 
thing like,  "What  kind  of  molecules  make  up 
living  systems?"  "How  does  a  living  system 
produce  the  energy  it  needs?"  or  "What  struc- 
tures do  biochemicals  have?" 


24-1     MOLECULAR  COMPOSITION  OF  LIVING  SYSTEMS 


The  chemical  system  of  even  the  smallest  plant 
or  animal  is  one  of  extreme  complexity.  It  has  a 
multitude  of  compounds,  many  of  polymeric 
nature,  existing  in  hundreds  of  interlocking  equi- 
librium reactions  whose  rates  are  influenced  by 
a  number  of  specific  catalysts.  We  will  not  try 
to  study  such  a  system.  Instead  we  will  show 
some  parts  of  it,  some  examples  that  have  been 
well  studied  and  which  illustrate  the  applicability 
of  chemical  principles.  All  of  our  knowledge  of 
biochemistry  has  come  through  use  of  the  same 
basic  ideas  and  the  same  experimental  method  you 
have  learned  in  this  course. 

We  shall  consider  four  classes  of  compounds 
that  have  great  importance  in  biochemistry. 
Sugars,  fats,  and  proteins  occur  in  most  animals 
and  plants,  while  cellulose  is  more  common  in 
plants.  These  are  all  discussed  in  the  following 
sections. 


24-1.1    Sugars 

SIMPLE    SUGARS 

The  word  "sugar"  brings  to  mind  the  sweet, 
white,  crystalline  grains  found  on  any  dinner 
table.  The  chemist  calls  this  substance  sucrose 
and  knows  it  as  just  one  of  many  "sugars" 
which  are  classed  together  because  they  have 
related  composition  and  similar  chemical  re- 
actions. Sugars  are  part  of  the  larger  family  of 
carbohydrates,  a  name  given  because  many  such 
compounds  have  the  empirical  formula  CH20. 


EXERCISE  24-1 


Glucose,  a  sugar  simpler  than  sucrose,  has  a 
molecular  weight  of  180  and  empirical  formula 
CH20.  What  is  its  molecular  formula? 


The  structure  of  the  glucose  molecule  was 
deduced  by  a  series  of  steps  somewhat  like  those 
described  in  Chapter  18  for  ethanol.  Glucose  was 

/        / 

found  to  contain  one  aldehyde  group    — C 

I    \. 

and  five  hydroxyl  groups  ( — OH).  These  func- 
tional groups  show  their  typical  chemistry.  The 
aldehyde  part  can  be  oxidized  to  an  acid  group. 
This  reaction  is  like  equation  (18-19)  (p.  336). 
If  a  mild  oxidizing  agent  (such  as  the  hypobro- 
mite  ion  in  bromine  water)  is  used,  the  aldehyde 
group  can  be  oxidized  without  oxidizing  the 
hydroxyl  groups. 


EXERCISE  24-2 


Write 


the 
O 


equation    for    the    oxidation    of 


/ 


R — C         by  hypobromite  ion,  BrO-,  to  pro- 

\ 
H 

duce  Br-.  What  are  the  oxidation  numbers  of 

carbon  and  bromine  before  and  after  reaction? 

Which  element  is  oxidized?  Which  is  reduced? 


SEC.    24-1    I    MOLECULAR    COMPOSITION    OF    LIVING    SYSTEMS 


423 


If  all  the  oxygen  containing  groups  are  re- 
duced, n-hexane  results.  This  test  helps  establish 
that  the  glucose  molecule  has  a  chain  structure. 
One  representation  of  the  structural  formula  of 
glucose,  C6Hi206,  is 

H    O 

V 

I 

CHOH 

I 
CHOH 

I 
CHOH 

I 
CHOH 

I 
CHOH 

Using  a  delicate  reduction  method,  the  alde- 
hyde group  can  be  converted  to  a  sixth  hydroxyl 
group,  giving  the  substance  called  sorbitol.  This 
compound  shows  the  typical  behavior  of  an 
alcohol.  For  example,  it  forms  esters  with  acids: 


CH2OH 

I 
CHOH 

I 
CHOH 

+  6H3CCOOH 

CHOH  acetic 

acid 

CHOH 

I 
CHoOH 


H  O 

I  II 

H— C— O— C— CH3 

O 

II 
H— C— O— C— CH3 


And  so  on  for  all  six  carbons. 
This  is  a  hexa-acetate. 


Another  naturally  occurring  sugar  is  fructose, 
also  C6Hi206.  It  is  an  isomer  of  glucose  but  the 


There  is  another  aspect  of  the  structure  of 
glucose  and  fructose.  They,  like  other  simple 
sugars,  can  exist  as  a  straight  chain  but  this  form 
is  in  equilibrium  with  a  cyclic  structure.  In  solu- 
tions the  latter  form  prevails.  Reaction  (2)  shows 
both  forms  of  glucose. 


H    .0 

C 

H 

CHOH 
1 
CHOH 

H    H-?-°H  0      H 
\   /•      \   / 

1 

CHOH 
1 

/  \?H       9/  \ 

'        C c   x 

CHOH 

1 

HO     1        1      OH 
H      OH 

(2) 


CH,OH 


The  ring  form  can  be  written  in  a  simpler  way, 
showing  the  hydrogen  atoms  attached  to  carbon 
atoms  by  lines  only  and  omitting  the  symbols  for 
the  ring  carbons: 


—  C— OH 


HO 


OH 


carbon  of  the      C=0  group  is  at  the  second 


EXERCISE  24-4 

At  equilibrium  in  a  0.1  M  solution  of  glucose  in 
position  in  the'carbon  chain  instead  of  at  the     water,  only  1  %  of  the  glucose  is  in  the  straight 
end.  This  makes  fructose  a  ketone  (see  Section     chain  form.  What  is  K  for  reaction  (2)? 
18-3.2).  . 


EXERCISE  24-3 


Draw  a  structural  formula  for  the  fructose  mole- 
cule (remember  that  fructose  is  an  isomer  of 
glucose).  Explain  why  fructose  cannot  be  oxi- 
dized to  a  six-carbon  acid. 


DISACCHARIDES 

The  two  sugars  we  have  discussed  are  mono- 
saccharides—they  have  a  single,  simple  sugar 
unit  in  each  molecule.  The  sugar  on  your  table 
is  a  disaccharide— it  has  two  units.  One  molecule 
of  sucrose  contains  one  molecule  of  glucose 
and  one  of  fructose  hooked  together  (losing  a 


424 


SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY       CHAP.    24 


molecule  of  water  in  the  joining  reaction).  Fruc-  what  with  the  sugar).  High  solubility  in  water 

tose  has  a  slightly  different  ring  structure  because  is  readily  explained  because  sugars  have  many 

\  functional  groups  that  can  form  hydrogen  bonds. 

the      C=0  group  is  not  on  the  end  carbon.  From  your  previous  study  of  hydrogen  bonding 

'  (Section    17-2.6)  you  might  recall  that  about 

The  formation  of  sucrose  is  shown  in  equa-  5  kca,  ar£  rdeased  per  mde  of  hydrogen  bonds 

tion  (5).  


C — OH 


HO 


OH 


OH  + 


glucose. 


HO 


C  — OH 


HO 


H20 


C  — OH 


sitcrose. 


PROPERTIES 

Sugars  occur  in  many  plants.  Major  commercial 
sources  are  sugar  cane  (a  large,  specialized  grass 
which  stores  sucrose  in  the  stem)  and  sugar  beet 
(as  much  as  15%  of  the  root  is  sucrose).  In 
addition,  fruits,  some  vegetables,  and  honey  con- 
tain sugars.  On  the  average  every  American  eats 
almost  100  pounds  of  sugar  per  year.  The  nation 
requires  about  2  X  1010  pounds  per  year  of  which 
about  one-fourth  is  grown  in  the  United  States. 
Sugar,  at  about  10  cents  a  pound  retail,  is  one 
of  the  cheapest  pure  chemicals  produced. 

Sugars  are  fairly  soluble  in  water,  about  5 
moles  dissolving  per  liter  (solubility  varies  some- 


formed.  This  energy  can  then  make  up  part  of 
that  needed  to  disrupt  the  structure  of  the 
crystal. 

Sugars  are  easily  oxidized.  One  oxidation  re- 
action that  shows  this  involves  cupric  hydroxide, 
O 


R— C 


y 


+  2Cu(OH); 


H 


R— C 


• 


+  Cu2Q(s)  +  2H20    (4) 


OH 


The  reaction  is  more  complicated  than  shown. 
The  Cu(OH)2  is  not  very  soluble  in  the  basic 


SEC.    24-1    I    MOLECULAR    COMPOSITION    OF    LIVING    SYSTEMS 


425 


solution  used,  so  tartaric  acid  is  added  to  form  a 
complex  ion.  The  Cu20(s)  is  a  red  solid  which 
does  not  form  such  a  complex  and  thus  precipi- 
tates from  solution.  The  reaction  is  characteristic 
and  is  used  as  a  qualitative  test  for  simple  alde- 
hydes and  for  sugars. 

An  important  metabolic  reaction  of  disaccha- 
rides  is  the  reverse  of  (3).  Water,  in  the  presence 
of  H+(aq),  reacts  with  sucrose  to  give  glucose 
and  fructose.  This  process  is  called  hydrolysis, 
meaning  "reaction  with  water." 


24-1.2    Cellulose  and  Starch 

Cellulose  is  an  important  part  of  woody  plants, 
occurring  in  cell  walls  and  making  up  part  of  the 
structural  material  of  stems  and  trunks.  Cotton 
and  flax  are  almost  pure  cellulose.  Chemically, 
cellulose  is  a  polysaccharide — a  polymer  made 
by  successive  reaction  of  many  glucose  molecules 
giving  a  high  molecular  weight  (molecular  weight 
~600,000).  This  polymer  is  not  basically  dif- 
ferent from  the  polymers  that  were  discussed  in 
Section  18-6: 


24-1.3     Fats 

Fats,  as  well  as  animal  and  plant  oils,  are  esters. 
Actually  they  are  triple  esters  of  glycerol  (1,2,3- 
propanetriol): 


H 

I 
H C- 

I 
O 


H 

I 
C 

I 

o 


H 


H 

I 
— C- 

I 

o 

V    \ 

H        H 


H    or    C3H803 


When  carboxylic  acids,  similar  to  those  you 
studied  in  Section  18-3.2,  react  with  glycerol  OH 
groups,  a  fat  is  formed.  In  natural  fats  the  acids 
usually  have  twelve  to  twenty  carbon  atoms, 
Ci6  or  Ci8  acids  being  most  common. 


EXERCISE  24-6 

Write  the  formula  for  glycerol  tributyrate,  and 
then  write  the  formula  of  the  fat  made  from 
glycerol  and  one  molecule  each  of  stearic 
(C17H35COOH),  palmitic  (C15H3iCOOH),  and 
myristic  (C13H27COOH)  acids.  How  many  iso- 


—  C— OH 


I 
—  C  — OH 


(5) 


—  C— OH 


OH 


Starch  is  a  mixture  of  glucose  polymers,  some 
of  which  are  water-soluble.  This  soluble  portion 
consists  of  comparatively  short  chains  (molecular 
weight  ~4000).  The  portion  of  low  solubility 
involves  much  longer  chains  and  the  polymer 
chain  is  branched. 


EXERCISE  24-5 

The  monomer  unit  in  starch  and  that  in  cellulose 
each  has  the  empirical  formula  C6Hi0O5.  These 
units  are  about  5.0  A  long.  Approximately  how 
many  units  occur  and  how  long  are  the  molecules 
of  cellulose  and  of  the  soluble  starch? 


OH 


OH 


mers  are  possible  for  the  last  fat?  How  many 
would  be  possible  if  all  possible  combinations  of 
the  three  acids  were  used?  Compare  your  answer 
with  that  for  Exercise   18-15,  p.   349. 


Common  fats  (butter,  tallow)  and  oils  (olive, 
palm,  and  peanut)  are  mixed  esters;  each  mole- 
cule has  most  often  three,  sometimes  two,  or, 
rarely,  one  kind  of  acid  combined  with  a  single 
glycerol.  There  are  so  many  such  combinations 
in  a  given  sample  that  fats  and  oils  do  not  have 
sharp  melting  or  boiling  points.  Ranges  are 
found  instead. 


426  SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY    |    CHAP.    24 


An  important  reaction  of  fats  is  the  reverse  of 
ester  formation.  They  hydrolyze,  or  react  with 
water,  just  as  disaccharides  do.  Usually  hydroly- 
sis is  carried  out  in  aqueous  Ca(OH)2,  NaOH,  or 
KOH  solution.  Because  of  long  use  in  the  prepa- 
ration of  soap  from  fats,  the  alkaline  hydrolysis 
reaction  (6)  is  called  saponification. 


The  metal  salts  of  natural  carboxylic  acids,  like 
sodium  stearate,  are  called  soaps. 

Fats  make  up  as  much  as  half  the  diet  of  many 
people.  Fats  are  a  good  source  of  energy  because 
when  they  are  completely  "burned"  in  the  body 
they  supply  twice  as  much  energy  per  gram  as 
do  proteins  or  carbohydrates. 


o 
I  n 

—  C  — O  — C  —  C,,H 


nll3S 


1  II 

—  C  —  O  —  C  —  Cl7H35     +    3Na+(a,)    +   30H~(aq) 

I  ° 

II 

C  — O  — C— C„H 


13  "27 


fat 


—  C — OH 


—    — C  — OH 


—  C— OH 


glycerol 


0 
2     Na_0-C-C17H35 
sodium,    stea.ra.ire. 

O 

II 
Na-  O-C—  Ci3H27 

sodium,   myristate. 


24-2  ENERGY  SOURCES  IN  NATURE 


24-2.1    Some  Fundamental  Biochemical 
Processes 

An  animal  (such  as  man)  expends  energy  con- 
tinuously, to  maintain  body  temperature  and  to 
perform  such  activities  as  breathing,  circulating 
blood,  and  moving  about.  What  chemical  proc- 
esses supply  this  energy? 

The  chief  source  of  such  energy  is  the  combus- 
tion of  carbon  compounds  to  C02.  You  know 
that  man  exhales  more  carbon  dioxide  than  he 
inhales  in  the  air  he  breathes.  This  extra  carbon 
dioxide  is  one  of  the  products  of  the  oxidation 
processes  by  which  food  is  oxidized  and  energy 
is  liberated. 

One  of  the  important  foods  of  animal  organ- 
isms is  sugar.  Man  eats  sugar  in  various  forms: 
as  sucrose,  as  glucose,  and  as  starch,  the  form 
in  which  sugar  is  stored  in  many  plant  tissues, 
such  as  the  potato.  Cellulose,  although  a  glucose 
polymer,  is  not  a  good  food  for  humans  because 
their  digestive  chemistry  cannot  hydrolyze  it  to 
sugar  rapidly  enough.  Termites,  however,  can 
hydrolyze  cellulose  and  they  find  wood  products 
quite  palatable. 

We  can  regard  sucrose  and  starch  as  sources 
of  glucose,  for  these  react  with  water  to  form 
glucose  in  the  body: 


C12H220„  +  H20 


C6H1206  +  C6H1206    (7a) 

glucose  fructose 


HO(C6H10O5)„H  +  (n  -  1)H20 

starch 


/jC6H1206    (7b) 

glucose 


Glucose  is  one  of  the  most  important  sources 
of  energy  for  living  creatures  of  all  kinds.  We 
can  illustrate  this  with  the  fermentation  of  sugar. 
Yeast,  a  plant,  uses  glucose  in  a  chemical  reac- 
tion 

QH1206  — »-  2C2H5OH(*J  +  2C02(gj       (8) 

In  this  transformation  of  glucose  into  alcohol 
and  carbon  dioxide,  energy  is  liberated,  and  this 
energy  is  used  by  the  yeast  plants.  Thus  glucose 
is  used  as  a  fuel  by  the  growing  organism  to 
furnish  the  energy  needed  for  growth. 

The  process  by  which  yeast  breaks  down  glu- 
cose has  been  carefully  studied  by  biochemists 
and  the  way  in  which  this  transformation  occurs 
is  now  known  in  considerable  detail.  One  of  the 
reasons  this  process  is  so  interesting  is  that  a 
nearly  identical  process  takes  place  in  human 
muscle,  in  this  case  to  furnish  energy  needed  for 
muscular  activity. 

Both  yeast  and  muscle  break  glucose  down 
into  an  acid  called  pyruvic  acid, 


SEC.  24-2  I  ENERGY  SOURCES  IN  NATURE 


427 


O 

/ 

CH3— C  O 

V 

\ 

O— H 

The  process  requires  eight  separate  steps,  all  of  which  are  carried  out  with  the  aid  of  biological 
catalysts  called  enzymes.  We  can  picture  the  series  of  reactions  in  the  following,  much  sim- 
plified manner: 


0 

V 

CH.OH 

CHOH 

c=o 

1 

CHOH 

is 

CHOH 

and 

1 

converted 

1 

broken 

CHOH 

1 

to 

CHOH 

1 

down 
into 

CHOH 

CHOH 

CH2OH 

CH2OH 

glucose 
(C.HiiOi) 

fructose 
(CiHuO.) 

H 


CH:OH 
CHOH 


V 


which  reacts 

with  water 

to  give 


glyceraldehyde 
2(CiH.0i) 


CH2OH" 

I 
CHOH 

I 
CHOH 

I 
OH 

an  intermediate 


(9) 


CH2OH" 

I 
CHOH 

I 
CHOH 

I 
OH 


reacts  with  an  enzyme 
(E)  to  produce 


CH2OH 

I 
CHOH 

C 

/  \ 
HO  O 

glyceric  acid 


+  reduced  enzyme  (£H2)         (10) 


HO 


CH2OH 

CH2 

then 

II 

CHOH      loses  a 

C— OH 

water 

C            molecule 

C 

\          to  give 

/  V 

+  H20 


O 


HO 


O 


(11) 


CH2 

II 

C— OH 


HO 


C 


O 


rearranges 
into 


U2) 


CH3 

I 

c=o 

I 

c 

/  \ 

HO  O 

pyruvic  acid 


To  this  stage,  yeast  and  muscle  reactions  are  the  same.  Now,  however,  they  proceed  differently: 


In  Yeast 

CH3 

CH3 
C=0  is  | 

decomposed  C        +  C02 

C  to  /  \ 

/  \  HO 

HO  O  acetaldehyde 

U3) 


CH3 

I 
C 

/  \ 
H  O 


reduced 


enzyme 

•om  (10 
above, 


"  from  (W),  (EH2) 


CH3 

+  enzyme  (E) 
CH.OH 

etbanol 

(14) 


428 


SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY    I    CHAP.    24 


You  will  notice  that  in  equation  (9)  each  mole- 
cule of  six-carbon  sugar  gave  two  molecules  of 
the  three-carbon  compound,  glyceraldehyde. 
Thus,  each  molecule  of  glucose  gives  two  C02 
and  two  ethanol  molecules. 


In  Muscle 
CH3 


c=o 


+  reduced  enzyme  (EH2) 


HO 


V 


O 


-f  enzyme  (E)    (75) 


CH3 

CHOH 

I 
C 

/  \ 
HO  O 

lactic  acid 


Lactic  acid  is  commonly  produced  when  sugar 
is  broken  down  by  living  cells.  Lactic  acid  is  so 
named  because  it  is  produced  when  milk  sours. 

What  happens  to  the  energy  from  the  oxida- 
tion of  glucose?  A  study  of  the  breakdown  of 
glucose  in  the  absence  of  oxygen  shows  that  about 
20  kcal  are  liberated  per  mole  of  glucose  con- 
sumed : 

C6H1206fsj  — >-  2C2HbOH(l)  +  2C02(g) 

AH  =  -20  kcal    (76) 

This  energy  is  used  by  the  organism  to  synthesize 
other  very  reactive  chemical  compounds  which 
are  not  shown  in  our  simplified  scheme.  These 
reactive  molecules  then  take  part  in  other  proc- 
esses (such  as  muscle  action)  in  which  the  energy 
is  released.  Part  of  the  energy  of  glucose  is  stored 
as  heat  content  (or  "chemical  energy")  in  the 
reactive  compounds.  This  "storage"  of  energy 
in  compounds  enables  living  organisms  to  make 
efficient  use  of  the  energy  of  oxidation. 


QHuOefsJ  +  602fgj  — ►■ 

6C02(gj  +  6H20(l)  +  673  kcal    (77) 

Since  673  kcal/mole  could  be  released  by  com- 
plete oxidation,  we  might  wonder  why  the  yeast 
cells  (and  muscle)  extract  only  20  kcal/mole  and 
leave  so  much  of  the  potentially  available  energy 
untouched.  This  extra  energy  is  there  in  ethanol 
and  lactic  acid  and  could  be  released  if  these 
compounds  were  oxidized  further  to  C02. 

Fermentation  in  the  absence  of  oxygen,  equa- 
tion (76),  with  the  liberation  of  only  a  small  frac- 
tion of  the  total  available  energy  of  glucose,  is 
not  the  usual  chemistry  employed  by  living  or- 
ganisms. It  is  a  "reserve"  mechanism,  useful  in 
tiding  over  yeast  or  muscle  during  times  of  oxy- 
gen shortage.  In  the  case  of  muscle,  it  provides  a 
temporary  source  of  energy  when  excessive  de- 
mands require  the  performance  of  work  at  a 
faster  rate  than  oxygen  can  be  supplied  by  the 
circulation  of  blood  to  the  tissue. 

Ordinarily,  oxygen  is  used  and  glucose  is  oxi- 
dized all  the  way  to  C02  and  H20.  Most  living 
creatures  exist  in  contact  with  a  supply  of  oxy- 
gen, either  in  the  air  or  dissolved  in  water.  Hence, 
most  of  the  metabolic  activities  of  the  living 
world  occur  in  the  presence  of  oxygen.  Under 
these  conditions,  the  breakdown  of  the  glucose 
molecule  proceeds  without  the  use  of  oxygen 
only  to  the  point  at  which  pyruvic  acid  is  formed. 
Then  the  pyruvic  acid,  instead  of  being  reduced 
to  lactic  acid  or  ethanol  and  C02,  is  oxidized  by 
oxygen  to  C02  and  water.  It  is  during  this  oxida- 
tion that  most  of  the  energy,  originally  available 
in  glucose,  is  liberated. 


EXERCISE  24-7 

Write  a  balanced  equation  for  the  oxidation  of 
pyruvic  acid  to  C02  and  H20. 


24-2.2    Oxidative  Metabolism 

If  a  mole  of  glucose  is  completely  burned  in  a 
calorimeter,  a  great  deal  of  energy  is  liberated: 


Little  of  this  energy  is  directly  given  off  as 
heat,  as  it  would  be  if  we  burned  pyruvic  acid 
with  a  match.  Most  of  the  energy  is  stored  in  new 


SEC.  24-2  I  ENERGY  SOURCES  IN  NATURE 


429 


chemical  compounds  that  can  undergo  reactions 
leading  to  the  synthesis  of  fats  and  proteins  and 
the  other  substances  of  which  living  matter  is 
made. 

Let  us  look  at  a  very  simple  part  of  the  overall 
process.  The  first  thing  that  happens  is  the  break- 
down of  pyruvic  acid  into  acetic  acid  and  C02 
(the  "oxidation"  process  has  started!): 


O 

/ 

CH3C  O 

V 


o 


oxidation  <y 

— >     CH3C 


OH 


OH 


+    CQ2(g) 


U8) 


The  acetic  acid  then  enters  a  cycle  of  reactions 


in  which  it  is  the  fuel  that  keeps  the  cycle  run- 
ning, and  C02,  water  and  energy  are  taken  off 
along  the  course  of  the  cycle.  This  cycle  is  shown 
in  Figure  24-1  and  its  steps  are  represented  by 
equations  (19a)  to  (19f). 

Each  "turn"  around  this  cycle  uses  up  one 
molecule  of  acetic  acid  and  produces  two  of  C02 
and  two  of  water.  The  stages  at  which  oxidation 
occurs  are  not  shown  in  detail,  for  this  is  a  rather 
complex  subject  and  is  beyond  our  present  ability 
to  consider  in  detail.  However,  it  can  be  seen 
that  the  oxidation  does  occur  (because  acetic 
acid  and  oxygen  are  fed  in  and  C02  and  H20 
are  discharged),  and  the  energy  of  "burning" 
sugar,  the  source  of  the  acetic  acid,  is  being 
released. 


Fig.  24-1.  The  cycle  by  which  acetic  acid  is  burned  as  an  energy  source. 


I 


CH,COOH 


(l9a) 


Wergy  +  K2  0   -« 


H20 


/Fumaric)      CHCOOH 
j      acid. 


CHCOOH 


1 


'  C-COOH 
I 
CH2COOH 

M9f) 


CH,COOH 
I 


HO-C-COOH     (C£%) 


+   1  °2    ♦* 


HO-CHCOOH 
I 
CH2COOH 

k(l9e) 


I 
CH,COOH 


\  / 

\  / 

\  / 


(19b) 


O. 


-►  COz  -h  H20  + 

eneray 


Oxygen 
supply 


C-COOH 

I 
CH, 

I 
CH2COOH 


/ 


\ 


/ 


Vz  02    in 


CHjCOOH 

I 

(19  d)  CH,COOH 


+  <T°2  iM- 


O,  ■*■   enerc 


(19  c  J 


430  SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY    j    CHAP.    24 


^ 


C-COOH 


CH,COOH 


CH,COOH 


+     CH*COOH     ■►    HO-C—  COOH 

I 
CH2C00H 


(19  a) 


CH2C00K 


HO  —  C  —  COOH      +         -j  Oz 


CHZCOOH 


^ 


C—COOH 
I 
CH2  +   COz     +     }izO     +     energy         (l9  b) 

CHzCOOH 


^ 


C—COOH 

I 
CH2 

I 
CH,COOH 


H 


CH,COOH 


CH2COOH 


+      COz        +      energy  (l9c) 


CH,COOH 

I 
CH2C00H 


+      fo2 


CHCOOH 


CHCOOH 


+        HzO       +      energy  (*9d) 


CHCOOH 


CHCOOH 


+         H,0 


HO  — CHCOOH 

I 
CH2COOH 


(l9e) 


HO  — CHCOOH 
I 
CH2COOH 


+         z02 


'** 


C-COOH 


CH2  COOH 


+       HzO       r       energy  (l9f) 


^ 


C-COOH 


CH,COOH 


+      CH3COOH        *-       cycle    begins    again 


(19  a) 


24-2.3    Photosynthesis  Green  plants  reverse  the  process  of  sugar 

What  is  the  source  of  the  vast  amount  of  energy  breakdown  by  synthesizing  sugars  from  C02  and 

consumed    by   our   mechanized   society?   The  water: 

largest  of  all  of  our  energy  sources  is  the  sun,  f£X^g)  +  6HjQ(l)  +  m  kcal  _^. 

C6HI206  +  60/gJ    (20) 


and  energy  from  the  sun  is  stored  in  our  fuels 
(wood,  coal,  petroleum)  as  a  result  of  the  photo- 
synthesis process. 


We  cannot  specify  exactly  how  this  energy  is 


SEC.    24-3    I    MOLECULAR    STRUCTURES    IN    BIOCHEMISTRY 


431 


used  by  the  plant.  We  know  that  the  reaction 
requires  a  special  compound,  chlorophyll,  which 
is  the  green  compound  that  gives  green  plants 
their  color.  Chlorophyll  does  not  appear  as  a 
reactant  in  equation  (20)  because  it  is  a  catalyst. 
We  also  know  that  the  process  of  incorporat- 
ing C02  into  complex  molecules  (that  result 
finally  in  the  synthesis  of  sugar)  is  very  similar 
to  the  reverse  of  the  process  of  sugar  breakdown. 
These  reactions  are  complicated,  however,  and 
will  not  be  discussed  here. 


EXERCISE  24-8 

The  sun  provides  about  0.50  calorie  on  each 
square  centimeter  of  the  earth's  surface  every 
minute.  How  long  would  it  take  ten  leaves  to 
make  1 .8  grams  of  glucose  if  the  area  of  each 
leaf  is  10  cm2  and  if  only  10%  of  the  energy  is 
used  in  the  reaction? 


EXERCISE  24-9 

Normally  about  0.03  %  of  the  molecules  in  air 
are  carbon  dioxide  molecules.  How  many  liters 
of  air  (at  STP)  are  needed  to  provide  enough 
C02  to  form  1.8  grams  of  glucose  in  reaction 
(20)1 

EXERCISE  24-10 

Suppose  red  light  of  wavelength  6700  A  is  ab- 
sorbed by  chlorophyll. 

(a)  Show  that  the  frequency  of  this  light  is 
4.5  X  1014  cycles  per  second. 

(b)  How  much  energy  is  absorbed  per  mole  of 
photons  absorbed?  (See  Section  15-1.1; 
h  =  9.5  X  10""  kcal  sec/mole.) 

(c)  How  many  moles  of  photons  would  be 
needed  to  provide  enough  energy  to  produce 
one  mole  of  glucose  by  reaction  (20)  if  all 
of  the  energy  were  provided  by  red  light? 


24-3    MOLECULAR  STRUCTURES  IN  BIOCHEMISTRY 


Some  of  the  most  exciting  recent  advances  in 
biochemistry  have  come  from  recognition  of  the 
importance  of  the  structural  arrangement  of 
molecular  parts.  You  saw  in  Chapter  18  that 
the  chemistry  of  a  C2H60  compound  depends 
upon  structure.  Thus  an  ether,  CH3 — O — CH3, 
behaves  quite  differently  from  an  isomeric  alco- 
hol, CH3CH2OH.  You  also  learned  how  inter- 
actions between  molecules  can  influence  the 
properties  of  water  (Sections  17-2.5,  17-2.6)  and 
arrange  the  molecules  in  preferred  positions 
around  an  ion  (Figure  17-13).  This  kind  of  struc- 
ture also  influences  the  observed  properties. 
Both  the  covalent  bond  arrangement  and  inter- 
molecular  interactions  are  involved  in  fixing  the 
structure  of  biochemical  substances.  A  few  ex- 
amples are  discussed  below. 

24-3.1    The  Structure  of  Starch  and  Cellulose 

A  striking  example  of  the  effect  of  structure  is 
shown  by  cellulose  and  water-soluble  starch. 
Both  contain  the  same  monomer  since  hydroly- 


sis gives  only  glucose  in  each  case.  But  the 
glucose  ring  differs  slightly  in  the  arrangement 
of  the  OH  groups.  This  results  in  two  different 
polymers.  Let  us  represent  the  ring  structure  of 
equation  (2)  by  this  simplified  symbol: 


HO 


cc  -form 
(21a) 


There  is  another  isomer,  identical  in  all  parts, 
except  for  the  placement  of  the  right-hand  OH 
group.  Its  symbol  is 


/?  -form 

(21b) 

If  we  connect  a  string  of  the  a-form  and  allow 
for  the  normal  105°  angle  of  bonds  to  oxygen  we 
get  the  polymer  called  starch: 


432 


SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY       CHAP.    24 


OH 


—in 


starch 


(22a) 


On  the  other  hand,  a  chain  of  the  /3-form  of 
glucose  gives  the  polymer  called  cellulose: 


cetitilose 


the  coiled  form  of  the  proteins  it  contains.  A  few 
moments  of  thought  concerning  the  profound 


v0II 

n 
(22b) 


The  very  different  geometry  of  the  ether  link- 
ages in  starch  and  cellulose  causes  these  two 
polymers  to  have  different  chemical  properties. 

24-3.2    Proteins 

In  Section  18-6.3  the  composition  of  proteins 
was  given.  They  are  large,  amide-linked  poly- 
mers of  amino  acids.  However,  the  long  chain 
formula  (Figure  18-14,  p.  348)  does  not  represent 
all  that  is  known  about  the  structure  of  proteins. 
It  shows  the  covalent  structure  properly  but  does 
not  indicate  the  relative  positions  of  the  atoms 
in  space. 

The  use  of  X-ray  diffraction  (Section  14-3.2) 
and  the  principles  that  describe  hydrogen  bond- 
ing have  led  to  the  recognition  of  a  coiled  form  of 
the  chain  in  natural  proteins.  This  model  is  con- 
sistent with  other  tests  also  and  has  received 
general  acceptance.  It  is  shown  in  Figure  24-2. 
This  form  has  a  great  deal  of  regularity — it  is  not 
at  all  a  random  shape.  Order  is  not  achieved 
without  some  energy  factor  to  maintain  it,  and 
this  order  results  in  large  part  from  hydrogen 
bonds.  These  are  shown  in  the  figure  as  dotted 
lines,  just  as  they  were  in  Section  17-2.6.  When 
the  hydrogen  bonds  are  broken  (by  heating  or 
putting  the  protein  in  alcohol),  the  order  dis- 
appears and  the  coiled  form  loses  its  shape.  Often 
this  damage  cannot  be  repaired  and  the  coil  is 
permanently  deformed.  Cooking  an  egg  destroys 


differences  between  the  physical  form  and  be- 
tween the  chemical  potentialities  of  an  egg  before 
and  after  it  is  cooked  will  suggest  the  very  great 
importance  of  molecular  structure  in  biochem- 
istry. 

24-3.3     Enzymes 

All  of  the  biochemical  reactions  we  have  just 
been  discussing  proceed  at  ordinary  tempera- 
tures and  pressures.  Most  biochemical  reactions 
(especially  those  in  the  human  body)  take  place 
at  about  37°C  (98°F),  and  proceed  at  a  rate  ade- 
quate for  the  role  they  play,  which  is  to  make 
life,  growth,  and  reproduction  possible.  Most  of 
them  would  not  proceed  at  a  measurable  rate  at 
this  temperature  outside  of  living  organisms. 
Glucose,  pyruvic  acid,  acetic  acid,  etc.,  are  very 
stable  compounds  and  can  remain  in  contact 
with  oxygen  without  apparent  change.  This  is  so, 
even  though  their  oxidation  to  carbon  dioxide 
and  water  releases  large  amounts  of  energy.  To 
make  these  reactions  proceed,  nature  uses  cata- 
lysts to  provide  new  paths  with  lower  activation 
energy  hills  over  which  the  systems  can  pass  so 
that  measurable  reaction  rates  are  achieved. 

Biological  catalysts  are  called  enzymes.  Nearly 
every  step  of  the  breakdown  of  a  complex  mole- 
cule to  a  series  of  smaller  ones,  within  living 
cells,  is  catalyzed  by  specific  enzymes.  For  in- 
stance, when  acetaldehyde  is  reduced  in  yeast 


SEC.    24-3    I    MOLECULAR    STRUCTURES    IN    BIOCHEMISTRY 


433 


cells  to  ethanol,  in  a  process  we  can  represent  as 
follows, 

O  OH 

/  / 

CH3— C         +  2[H]  — *■  CH3— CH2 


(23) 


\ 


H 


the  reaction  takes  place  in  the  presence  of  a 
specific  enzyme  called  "alcohol  dehydrogenase." 
You  can  see  that  the  hydrogenaiion  of  acetalde- 
hyde  is  the  reverse  of  the  dehyclrogenation  of 
ethanol.  The  enzyme  is  named  for  the  latter 


Fig.  24-2.  The  coiled  or  helix  form  of  a  protein 
molecule. 


reaction,  but  of  course  it  catalyzes  the  reaction 
in  either  direction  (see  Section  9-1.4).  Con- 
ditions at  equilibrium  are  not  affected  by  the 
enzyme,  but  the  rate  at  which  the  reacting  sub- 
stances reach  the  equilibrium  state  is  affected  by 
the  enzyme  (as  with  any  catalyst). 

Enzymes  are  protein  molecules.  While  all  en- 
zymes are  proteins,  we  do  not  imply  that  all 
proteins  can  act  as  enzymes.  The  protein  mole- 
cules of  enzymes  are  very  large,  with  molecular 
weights  of  the  order  of  100,000.*  In  contrast,  the 
substance  upon  which  the  enzyme  acts  (called  a 
substrate)  is  very  small  in  comparison  with  the 
enzyme.  This  creates  a  picture  of  the  reaction  in 
which  the  small  substrate  molecule  becomes  at- 
tached to  the  surface  of  the  large  protein  mole- 
cule, at  which  point  the  reaction  occurs.  The 
products  of  the  reaction  then  dissociate  from  the 
enzyme  surface  and  a  new  substrate  molecule 
attaches  to  the  enzyme  and  the  reaction  is  re- 
peated. We  can  write  the  following  sequence: 

enzyme  +  substrate  — >- 

enzyme-substrate  complex        (24) 


enzyme-substrate  complex  — >- 

enzyme  +  reaction  products 


(25) 


Adding  these  equations  and  cancelling  gives 

substrate  — >■  reaction  products  (26) 

Despite  the  large  size  of  an  enzyme  molecule, 
there  is  reason  to  believe  that  there  are  only  one 
or  a  few  spots  on  its  surface  at  which  reaction 
can  occur.  These  are  usually  referred  to  as 
"active  centers."  The  evidence  for  this  view  of 
enzyme  reactions  comes  from  many  kinds  of 
observations.  One  of  these  is  that  we  can  often 
stop  or  slow  down  enzyme  reactions  by  adding 
only  a  small  amount  of  a  "false"  substrate.  A 
false  substrate  is  a  molecule  that  is  so  similar  to 
the  real  substrate  that  it  can  attach  itself  to  the 
active  center,  but  sufficiently  different  that  no 
reaction  and  consequently  no  release  occurs. 
Thus,  the  active  center  is  "blocked"  by  the  false 
substrate. 


*  Since  so  many  enzymes  are  known,  this  number  is 
given  only  to  offer  a  rough  idea  of  size,  because  actual 
molecular  weights  may  range  from  considerably  less  than 
100,000  to  considerably  more. 


434 


SOME    ASPECTS    OF    BIOCHEMISTRY:    AN    APPLICATION    OF    CHEMISTRY    I    CHAP.    24 


SPECIFICITY    OF    ENZYMES 

Most  enzymes  are  quite  specific  for  a  given  sub- 
strate. For  example,  the  enzyme  "urease"  that 
catalyzes  the  reaction 


NH2 


o=c 


+  H20  +=£  C02  +  2NH,    (27) 


NH2 


is  specific  for  urea.  If  we  try  to  use  urease  to 
catalyze  the  analogous  reaction  of  a  very  similar 
molecule,  W-methylurea,  no  catalysis  is  ob- 
served : 

NHCHj 


o=c 

\ 
NH2 

AT-methylurea 


+  H20  ++  C02  +  NH3  +  CH3NH2 

(28) 


This  suggests  that  on  the  surface  of  the  enzyme 
there  is  a  special  arrangement  of  atoms  (belong- 
ing to  the  amino  acids  of  which  the  protein  is 
constructed)  that  is  just  right  for  attachment  of 
the  urea  molecule  but  upon  which  the  methyl 
urea  will  not  "fit." 

Specificity  is  not  always  perfect.  Sometimes  an 
enzyme  will  work  with  any  member  of  a  class  of 
compounds.  For  example,  some  esterases  (en- 
zymes that  catalyze  the  reaction  of  esters  with 
water)  will  work  with  numerous  esters  of  similar, 
but  different,  structures.  Usually,  in  cases  of  this 
kind,  one  of  the  members  of  the  substrate  class 
will  react  faster  than  the  others,  so  the  rates  will 
vary  from  one  substrate  to  another. 

A   PRACTICAL    APPLICATION   OF   ENZYME 
INHIBITION   BY    A    "FALSE"    SUBSTRATE 

It  is  now  believed  that  many  of  our  useful  drugs 
exert  their  beneficial  action  by  the  inhibition  of 
enzyme  activity  in  bacteria.  Some  bacteria,  such 
as  staphylococcus,  require  for  their  growth  the 
simple   organic   compound  paraaminobenzoic 


acid  and  can  grow  and  multiply  in  the  human 
body  because  sufficient  amounts  of  this  com- 
pound occur  in  blood  and  the  tissues.  The  con- 
trol of  many  diseases  caused  by  these  (and  other) 
bacteria  was  one  of  the  first  triumphs  of  chemo- 
therapy,* and  the  first  compound  found  to  be 
an  effective  drug  of  this  type  was  sulfanilamide: 

NH2  NH2 


(29) 


0=C— OH 

0=S— NH 

II 
O 

paraaminobenzoic  acid 

sulfanilamide 

It  seems  reasonable  that  an  enzyme  which 
used  /xzraaminobenzoic  acid  as  a  substrate 
might  be  deceived  by  sulfanilamide.  The  two 
compounds  are  very  similar  in  size  and  shape 
and  in  many  chemical  properties.  To  explain  the 
success  of  sulfanilamide,  it  is  proposed  that  the 
amide  can  form  an  enzyme-substrate  complex 
that  uses  up  the  active  centers  normally  occupied 
by  the  natural  substrate. 

Usually  fairly  high  concentrations  of  such  a 
drug  are  needed  for  effective  control  of  an  in- 
fection because  the  inhibitor  (the  false  substrate) 
should  occupy  as  many  active  centers  as  possible, 
and  also  because  the  natural  substrate  will  prob- 
ably have  a  greater  affinity  for  the  enzyme.  Thus 
the  equilibrium  must  be  influenced  and,  by  using 
a  high  concentration  of  the  false  substrate,  the 
false  substrate-enzyme  complex  can  be  made  to 
predominate.  The  bacteria,  deprived  of  a  normal 
metabolic  process,  cannot  grow  and  multiply. 
Now  the  body's  defense  mechanisms  can  take 
over  and  destroy  them. 

*  Chemotherapy  is  the  control  and  treatment  of  disease 
by  synthetic  drugs.  Most  of  these  are  organic  compounds, 
often  of  remarkably  simple  structure.  Sulfanilamide  is  one 
example  of  an  organic  compound  synthesized  by  chemists 
for  the  treatment  of  bacterial  infections. 


Robert  Woodward  is  surely  one  of  the  outstanding  Ameri- 
can synthetic  organic  chemists  of  all  time.  His  astonishing 
record  of  successful  syntheses  of  biologically  important 
substances  has  won  him  ten  honorary  doctorate  degrees 
and  at  least  as  many  major  awards  here  and  abroad. 

Woodward  was  born  in  Quincy,  Massachusetts.  His  in- 
terest in  chemistry  developed  at  an  early  age  and  it  seemed 
to  grow  without  need  for  stimulation  nor  urging.  By  the 
time  he  entered  Massachusetts  Institute  of  Technology  at 
the  age  of  sixteen,  he  knew  as  much  organic  chemistry  as 
the  average  graduating  senior.  M.I.T.  recognized  his  capa- 
bilities and  opened  the  laboratory  to  him.  He  passed  course 
examinations  at  a  rate  and  with  a  performance  that  brought 
him  the  Bachelor  s  Degree  in  three  years  and,  in  only  one 
additional  year,  the  Ph.D.  Professor  J.  F.  Norris,  then 
director  of  the  M.I.T.  laboratory  announced,  "We  saw  we 
had  a  person  who  possessed  a  very  unusual  mind.  ■  ■  ■  We 
think  lie  will  make  a  name  for  himself  in  the  scientific 
world." 

How  richly  this  prophecy  has  come  true  is  read  in  the 
scientific  literature  of  organic  chemistry.  His  first  major 
success  was  the  synthesis  of  quinine,  a  problem  he  began 
worrying  as  a  high  school  student.  This  compound  was 
to  be  typical  of  the  difficult  molecules  he  has  successfully 


synthesized.  Quinine  has  the  formula  CxHuOtN*,  and  it 
includes  two  benzene  rings,  a  tricyclic  structure,  a  double 
bond,  an  OH  group,  and  a  methyl  ether  linkage,  all  com- 
bined in  a  very  particular  geometrical  configuration.  A  host 
of  other,  comparably  difficult  syntheses  have  been  achieved 
by  Woodward  and  his  large  group  of  students,  including 
some  natural  products  whose  importance  has  made  their 
names  familiar  household  words:  cholesterol,  cortisone, 
and  chlorophyll.  His  contributions  to  the  structure  deter- 
minations of  antibiotics  reads  like  a  doctor's  chemical 
shelf,  including  penicillin,  terramycin,  and  aureomycin.  He 
has  added  to  our  knowledge  of  the  polymerization  processes 
by  which  amino  acids  link  into  proteins,  and  some  of  his 
synthetic  protein-like  polymers  have  physical  properties 
quite  comparable  to  those  of  silk  and  wool  fibers. 

All  of  these  accomplishments  bespeak  a  dedication  to 
chemistry  and  a  capacity  for  work  that  are  commensurate 
with  his  intellectual  capability.  Woodward  can  be  found 
in  his  office  or  in  the  laboratory  after  midnight  many  days 
a  week.  However,  the  rewards  for  such  intensive  effort  are 
proportionate.  To  Robert  Woodward,  as  to  most  chemists, 
chemistry  is  an  exciting  adventure  for  which  there  just 
isn't  enough  time. 


CHAPTER 


The  Chemistry 
of  Earth, 
the  Planets, 
and  the  Stars 


There  may,  of  course,  be  types  of  life  with  a  wholly  different  chemical 
basis  to  our  own,  for  example,  a  low  temperature  life  on  the  outer  planets 
which  is  based  on  reactions  in  liquid  ammonia. 

J.    B.    S.    HALDANE,    1960 


The  advent  of  space  exploration  quickens  the 
pulse  of  every  scientist  including,  as  much  as  any, 
that  of  the  chemist.  Chemists  are  playing  many 
crucial  roles:  preparing  new  fuels,  new  metals, 
and  new  plastics  to  cope  with  a  new  range  of 
performance  needs;  anticipating  the  environ- 
mental chemistry  that  will  face  the  first  space 
explorers,  devising  ways  to  permit  survival  under 
conditions  so  extreme  that  they  are  not  even 
present  on  our  planet,  collecting  and  interpreting 
data  that  will  reflect  onto  and  illuminate  age-old 
questions  about  the  origin  of  the  earth,  the  solar 
system,  and  life  itself. 

This  age  is  heavily  laden  with  terrestrial  prob- 
lems for  the  chemist,  as  well.  Our  planet  seems 
to  be  shrinking  under  man's  fantastic  success  in 
curbing  nature's  whim  and  in  bending  it  to  his 
436 


will.  The  magnitude  of  this  success  is  measured 
in  an  awesome  and  exponential  population 
growth.  This  growth  portends  problems  of  food 
production,  fuel  consumption,  and  even  of  living 
space  that  dwarf  those  of  the  past.  Suddenly 
immense  urban  areas  are  threatened  by  air  pol- 
lution problems  that  were  completely  unknown 
thirty  years  ago.  Power  consumption  is  expand- 
ing so  rapidly  that  some  scientists  anticipate  the 
depletion  of  fossil  fuel  supplies  and  they  urge 
haste  in  developing  nuclear  fuels  and  in  harness- 
ing more  completely  the  vast  energy  of  the  sun. 
It  is  an  exciting  age — challenging,  yes,  but 
exciting.  It  is  an  age  in  which  we  must  understand 
our  planet,  Earth,  and  if  we  do,  we  can  begin  to 
venture  toward  our  neighbor  planets  and  beyond 
them  toward  the  stars. 


SEC.  25-1  !  THE  CHEMISTRY  OF  OUR  PLANET,  EARTH 


437 


25-1    THE  CHEMISTRY  OF  OUR  PLANET,  EARTH 


The  earth  is  the  source  of  all  the  substances 
directly  available  to  us.  Energy  comes  to  us  from 
outside  the  earth— from  our  sun  and  to  a  small 
degree  from  the  other  stars.  The  earth,  in  turn, 
radiates  energy  into  outer  space.  If  the  amount 
of  energy  radiated  is  more  than  the  amount 
received,  the  earth  will  grow  colder— if  it  is  less, 
the  earth  will  warm.  Some  of  the  solar  energy  is 
stored  as  chemical  heat  content  when  new  sub- 
stances, particularly  organic  compounds,  are 
formed.  For  short  periods  of  time  (as  measured 
in  terms  of  geologic  ages)  we  can  use  the  energy 
stored  in  fossil  fuels  such  as  coal  and  oil  or  we 
can  draw  on  nuclear  fuels.  In  the  long  run,  we 
shall  have  to  depend  upon  the  sun  as  our  pri- 
mary source  of  energy. 

The  substances  we  can  use  come  primarily 
from  the  earth.  In  its  movement  around  the  sun, 
the  earth  sweeps  through  space  and  collects 
material  from  meteors  and  some  cosmic  dust, 
but  the  amount  of  gathered  material  is  small 
compared  with  the  amount  present  in  the  earth.* 
We  shall  consider  the  material  of  the  earth  and 
see  how  it  is  put  to  use  by  mankind. 

25-1.1    The  Parts  of  the  Earth 

A  discussion  of  the  chemistry  of  the  earth  is 
conveniently  broken  into  three  parts,  each  of 
which  corresponds  to  one  of  the  phases  solid, 
liquid,  or  gas. 

The  lithosphere  is  the  solid  portion  of  the 
earth.  We  shall  use  the  term  to  include  the 
central  core,  though  there  remains  controversy 
as  to  whether  the  core  is  solid  or  liquid.  The 
lithosphere  is  a  sphere  of  solid  material  about 
4000  miles  in  radius.  We  have  direct  access  to 
only  a  minute  fraction  of  this  immense  ball.  The 
deepest  mine  penetrates  only  two  or  three  miles. 
The  deepest  oil  wells  are  about  five  miles  deep. 


*  It  has  been  estimated  that  about  five  tons  of  material 
are  gathered  per  day  as  the  earth  sweeps  through  space. 
Yet  the  amount  collected  in  a  billion  years  would  form 
a  dust  layer  only  a  few  millimeters  thick  if  spread  evenly 
over  I  he  earth's  surface. 


This  relatively  thin  shell  that  we  can  study 
directly  is  called  the  earth's  crust.  In  view  of 
seismic  observations,  we  consider  the  thickness 
of  this  crust  to  be  about  20  miles.  The  remainder 
we  shall  call  the  inner  lithosphere,  which  includes 
the  central  part  called  the  core. 

About  80%  of  the  earth's  surface  is  covered 
with  aqueous  solution.  This  liquid  layer,  the 
oceans,  is  called  the  hydrosphere.  The  average 
depth  of  the  hydrosphere  is  about  three  miles 
but  at  ocean  "deeps"  or  "trenches,"  it  changes 
precipitously  to  depths  over  twice  that. 

Surrounding  the  earth  is  the  third  phase,  a 
gas.  The  gas  mixture  surrounding  the  earth  is 
called  the  atmosphere.  Over  98%  of  this  gas 
(air)  is  less  than  40  miles  above  the  earth's 
surface. 

25-1.2    Composition  and  Properties 
of  the  Atmosphere 

The  composition  of  the  earth's  atmosphere  dif- 
fers from  day  to  day,  from  altitude  to  altitude, 
and  from  place  to  place.  The  largest  variation  is 
in  the  concentration  of  water  vapor.  Water 
evaporates  continually  from  the  hydrosphere, 
from  the  soil,  from  leaves,  from  clothes  drying, 
etc.  At  intervals,  parts  of  the  atmosphere  become 
chilled  until  the  dew  point  or  frost  point  is 
reached  and  then  any  vapor  in  excess  of  the 
saturation  amount  is  precipitated  as  rain  or 
snow. 

Since  the  concentration  of  the  water  vapor 
varies  so  much,  geochemists  usually  report  the 
composition  of  "dry  air"— that  is,  air  from 
which  all  the  water  vapor  has  been  removed. 
The  composition  of  a  sample  of  dry  air  is  shown 
in  Table  25-1.  Notice  the  low  concentration  of 
hydrogen  and  helium  in  the  air.  The  earth  is  a 
rather  small  object  in  the  universe  and  exerts  a 
relatively  low  gravitational  attraction  on  the 
gases  above  it.  Hence,  most  of  the  hydrogen  and 
helium  originally  associated  with  the  matter  of 
the  earth  could  be  lost  rather  easily.  Notice  also 
that  nitrogen  is  more  abundant  than  oxygen  in 


438 


THE    CHEMISTRY    OF    EARTH,    THE    PLANETS,    AND    THE    STARS    |    CHAP.    25 


Earth  '&    c ru.S-t- 

20    miles 


Fig.  25-1.  The  parts  of  the  earth. 


the  air,  even  though  oxygen  is  much  more  abun- 
dant in  the  hydrosphere  and  lithosphere.  Notice 
also  that,  except  for  water  and  carbon  dioxide, 
the  major  components  of  air  are  elements. 

Table  25-1 

COMPOSITION    OF    A    CERTAIN 
SAMPLE    OF    DRY    AIR* 


SUBSTANCE 

PERCENT  OF 

Name 

Formula 

MOLECULES 

nitrogen 

N,. 

78.09 

oxygen 

o, 

20.95 

argon 

Ar 

0.93 

carbon  dioxide 

CO, 

0.03 

neon 

Ne 

0.0018 

helium 

He 

0.00052 

krypton 

Kr 

0.0001 

hydrogen 

H2 

0.00005 

xenon 

Xe 

0.000008 

♦Traces  of  other  compounds  (less  than 
0.0002%  of  the  molecules)  are  also  known  to  be 
present. 

With  the  aid  of  the  gas  laws  we  can  calculate 
the  relative  concentrations  of  the  components  in 
moist  air.  Suppose  that  on  a  certain  day  the 


Atm.  osph  ere 
200    miles 


Hydro  sph  ere 
~  3    miles 

atmospheric  pressure  is  750  mm  and  that  after 
drying  a  sample  of  this  air  we  find  the  pressure 
to  be  738  mm.  Then  the  partial  pressure  of  the 
water  vapor  was  750  mm  -  738  mm  =  12  mm. 
Since  the  partial  pressure  varies  directly  with  the 
number  of  molecules,  we  find  that  the  fraction 
of  the  air  molecules  that  are  water  molecules  in 
the  moist  gas  is  12  mm/750  mm  =  0.016.  In 
this  rather  moist  gas,  1.6%  of  the  molecules  are 
water  molecules. 

The  gravitational  force  on  a  heavy  molecule 
exceeds  that  on  a  light  molecule.  Consequently, 
there  is  a  tendency  for  "sedimentation"  of  the 
molecules  of  high  molecular  weight  relative  to 
the  gas  molecules  of  low  molecular  weight.  This 
is  opposed  by  the  tendency  toward  maximum 
randomness,  which  tends  to  keep  the  atmosphere 
thoroughly  mixed.  The  net  result  is  a  slight 
change  of  composition  with  altitude.  Dry  air  at 
sea  level  contains  about  78%  nitrogen  molecules 
and  21%  oxygen  molecules,  but  at  60,000  feet 
a  sample  of  dry  air  has  about  80%  nitrogen 
molecules  and  only  19%  oxygen  molecules. 

Apart  from  this  gravitational  effect,  there  are 
composition  changes  due  to  chemical  reactions 
induced  by  light.  These  are  caused  by  absorption 
of  ultraviolet  light  in  the  upper  atmosphere.  For 
example,  oxygen  absorbs  ultraviolet  light  and 
the  energy  taken  up  by  the  molecule  exceeds  the 


sec.  25' 


THE  CHEMISTRY  OF  OUR  PLANET,  EARTH 


439 


bond  energy.  The  bond  breaks,  to  give  two  oxy- 
gen atoms: 

o,(g)  +  hv  — ►■  20(g)  (/) 

Of  course  the  oxygen  atoms  so  produced  are 
very  reactive.  One  fate  of  these  atoms  is  to  com- 
bine with  another  oxygen  molecule,  02,  to  form 
ozone,  03: 

O(g)  +  0,(g)  — *-  03(g)  (2) 

Ozone  is  a  highly  reactive  form  of  the  element 
oxygen,  though  by  no  means  as  reactive  as  oxy- 
gen atoms.  It  is  produced  in  the  atmosphere  only 
at  high  altitudes  because  the  ultraviolet  light  of 
the  frequency  required  in  reaction  (7)  is  so  com- 
pletely absorbed  that  it  does  not  reach  lower 
altitudes.  Balloon  flights  have  shown  that  the 
ozone  concentration  is  negligible  at  sea  level  but 
that  it  rises  to  a  maximum  at  a  height  of  about 
15  miles. 

This  small  amount  of  ozone  15  miles  above 
the  earth's  surface  absorbs  most  of  the  ultra- 
violet light  not  absorbed  by  02.  Thus,  02  and  03 
together  make  the  atmosphere  opaque  in  most 
of  the  ultraviolet  spectral  region.  Presumably  the 
chemistry  of  life  on  this  planet  would  have 
evolved  quite  differently  if  this  ultraviolet  light 
reached  the  earth's  surface.  As  a  single  example, 
reflect  that  photosynthesis  would  have  photons 
of  much  higher  energy  with  which  to  operate  if 
the  atmosphere  were  transparent  in  the  ultra- 
violet region. 

EXERCISE  25-1 

Suppose  that  the  photosynthesis  reaction  (20b) 
in  Chapter  24  (p.  430)  could  be  based  upon  light 
of  wavelength  2400  A  (this  light  is  absorbed 
heavily  by  ozone).  How  many  moles  of  these 
photons  would  provide  the  673  kcal  of  energy 
needed  to  produce  one  mole  of  glucose?  (Re- 
member, E  =  hv,  and  h  =  9.5  X  10~14  kcal 
sec/mole).  Compare  your  answer  with  that  of 
Exercise  24-10. 


At  the  opposite  end  of  the  spectrum,  the  infra- 
red, again  the  atmosphere  becomes  virtually 
opaque.  This  is  due  mainly  to  absorption  by 


gaseous  water  and  carbon  dioxide.  Thus  we  find 
that  the  air,  normally  regarded  as  transparent, 
actually  serves  to  filter  the  sun's  rays  striking 
the  earth.  The  very  high  energy  photons  (in  the 
ultraviolet)  and  the  very  low  energy  photons  (in 
the  infrared)  are  removed  and  the  spectral  region 
between  is  transmitted. 


EXERCISE  2S-2 

On  this  planet,  of  what  value  would  be  an  eye 
that  is  sensitive  only  to  light  in  the  ultraviolet 
spectral  region?  Discuss  the  evolutionary  sig- 
nificance of  the  facts  that  the  human  eye  and  the 
photosynthesis  process  are  both  dependent  upon 
light  in  the  part  of  the  spectrum  called  the 
"visible." 


25-1.3    Composition  of  the  Hydrosphere 

Waters  exposed  to  air  dissolve  some  of  it.  In 
water,  oxygen  is  twice  as  soluble  as  nitrogen  but 
since  nitrogen  is  four  times  more  abundant  than 
oxygen  in  air,  more  dissolved  nitrogen  is  present 
than  dissolved  oxygen.  It  is  the  dissolved  ele- 
mentary oxygen  that  is  used  by  living  organisms 
for  their  oxidative  processes.  The  concentration 
of  dissolved  carbon  dioxide  is  low  because  its 
concentration  in  the  air  is  low.  But  dissolved 
carbon  dioxide  is  necessary  for  photosynthesis 
in  marine  plants.  Dissolved  carbon  dioxide  is 
responsible  for  part  of  the  pleasant  taste  of 
water.  Boiled  water  has  lost  almost  all  of  the 
dissolved  gases.  It  tastes  "flat." 

Ocean  water  also  contains  dissolved  molecules 
from  the  gases  of  the  air.  These  can  be  removed 
by  boiling,  but  other  solutes  remain.  When  a 
kilogram  of  average  ocean  water  is  distilled,  967 
grams  of  water  can  be  collected  and  33  grams  of 
solids  (primarily  salts)  remain  behind.  Thus,  we 
may  say  that  3.3%  of  the  weight  of  the  ocean 
water  is  due  to  dissolved  salts.  Actually,  more 
than  forty  of  the  elements  have  been  identified 
as  being  present  in  ocean  water  but  half  of  these 
are  present  in  very  small  concentrations  "less 
than  1  gram  per  billion,  109,  grams  of  water." 


440 


THE    CHEMISTRY    OF    EARTH,    THE    PLANETS,    AND    THE    STARS    I    CHAP.    25 


Table  25-11.     an  average  composition  of  ocean  water   (disregarding 

DISSOLVED    GASES) 


ELEM 

:NT 

PREDOMINANT 

NUMBER  OF  MOLES 

Name 

Symbol 

SPECIES 

PER 

KILOGRAM 

hydrogen 
oxygen 

"I 

OJ 

H20 

53.7 

chlorine 

CI 

C\-(aq) 

0.535 

sodium 

Na 

Na+(aq) 

0.460 

magnesium 

Mg 

Mg+2(aq) 

0.052 

sulfur 

S 

SOT2(aq) 

0.028 

calcium 

Ca 

Ca+2(aq) 

0.010 

potassium 

K 

K+(aq) 

0.010 

bromine 

Br 

Kr-(aq) 

0.008 

In  Table  25-11  are  shown  the  concentrations  (in 
number  of  moles  per  1000  grams  of  ocean  water) 
of  water  and  of  the  most  abundant  ions. 

Several  facts  become  apparent.  There  are 
fewer  Na+  ions  than  Cl~  ions;  other  positively 
charged  ions — Mg+2,  Ca+2,  and  K+ — are  also 
present.  Sulfate  ions,  S04~2,  and  bromide  ions, 
Br~,  are  other  negatively  charged  ions  present  in 
the  water.  Thus,  ocean  water  is  more  than  a 
solution  of  sodium  chloride.  Another  fact  is  that 
K+  ions  are  much  less  plentiful  than  Na+  ions 
(Na+/K+  is  about  46)  even  though  K+  ions  are 
rather  plentiful  in  the  earth  (Na+/K+  is  about  2). 

25-1.4    Composition  and  Properties 
off  the  Lithosphere 

We  know  very  much  about  the  outermost  por- 
tion of  the  lithosphere  because  it  is  available  for 
direct  study.  In  contrast,  we  know  almost  noth- 
ing about  the  inner  lithosphere,  though  it  con- 
stitutes over  99.5%  of  the  mass  of  the  earth. 

THE    INNER    LITHOSPHERE 

Seismic  observations  furnish  our  only  probe  of 
the  inner  lithosphere.  The  shock  waves  initiated 
by  an  earthquake  travel  through  the  interior  of 
the  earth  in  paths  that  are  bent  in  accordance 
with  the  elastic  properties  and  density  of  the 
medium  they  penetrate.  From  these  paths,  seis- 
mologists have  been  able  to  determine  the  ex- 
istence of  zones  within  the  lithosphere.   The 


outermost  portion,  or  mantle,  is  approximately 
2000  miles  in  depth  and  it  is  generally  thought 
to  be  solid.  This  solid  has  a  density  of  about  3 
grams  per  milliliter  near  the  crust  and  it  in- 
creases to  about  5  grams  per  milliliter  at  the 
bottom  of  the  mantle.  This  higher  density  is 
caused  by  increasing  pressure  in  the  depths  of 
the  earth.  For  a  comparison,  the  pressure  at  this 
depth  is  thought  to  be  over  a  million  atmos- 
pheres, two  or  three  times  greater  than  the 
highest  pressures  recorded  in  static  laboratory 
experiments. 

The  inside  of  the  lithosphere  is  called  the  core. 
Still  higher  pressures  are  expected  and  the  den- 
sity may  rise  as  high  as  18  grams  per  milliliter 
at  the  center  of  the  earth.  Some  of  the  core  may 
be  liquid  but  the  evidence  is  not  decisive. 

The  composition  of  the  mantle  is  probably 
rock-like,  meaning  it  is  made  up  of  various 
silicates.  These  minerals  have  densities,  com- 
pressibilities, and  rigidities  that  match  those 
indicated  by  the  seismic  studies.  The  core  has 
long  been  thought  to  be  largely  iron,  a  view  sug- 
gested by  the  composition  of  meteorites.  These 
are  solid  objects  that  plummet  through  the  at- 
mosphere from  space  and  they  may  be  pieces  of 
exploded  planets  resembling  earth.  Hence  their 
composition  furnishes  a  possible  clue  to  the 
composition  of  the  inner  lithosphere.  Current 
speculation  ranges  from  iron  to  a  high  density 
rock  but  new  evidence  is  needed. 

The  temperature  of  the  center  of  the  earth  is 


SEC.    25-1    I    THE    CHEMISTRY    OF    OUR    PLANET,    EARTH 


441 


thought  to  be  a  few  thousand  degrees.  Though 
this  would  melt  rock  at  the  earth's  surface,  solids 
can  remain  stable  at  the  exceedingly  high  pres- 
sures thought  to  exist  in  the  core. 

Needless  to  say,  very  much  remains  to  be 
learned  about  the  chemistry  of  the  inner  litho- 
sphere.  It  is  a  high  temperature  and  high  pressure 
laboratory  whose  door  has  not  yet  been  opened. 

THE    EARTH'S    CRUST 

Oxygen  and  silicon  are  the  most  abundant  ele- 
ments in  the  earth's  crust.  Table  25-1 II  shows 
that  60%  of  the  atoms  are  oxygen  atoms  and 
20%  are  silicon  atoms.  If  our  sample  included 
the  oceans,  hydrogen  would  move  into  the  third 
place  ahead  of  aluminum  (remember  that  water 
contains  two  hydrogen  atoms  for  every  oxygen 
atom).  If  the  sample  included  the  central  core 


Table  25-111 

ABUNDANCE    OF    ELEMENTS 

IN    THE    EARTH'S    CRUST* 

NO.  OF  ATOMS 

ATOMIC 

PER 

RANK               ELEMENT               NUMBER 

10,000  ATOMS 

1 

oxygen 

8 

6050 

2 

silicon 

14 

2045 

3 

aluminum 

13 

625 

4 

hydrogen 

1 

270 

5 

sodium 

11 

258 

6 

calcium 

20 

189 

7 

iron 

26 

187 

8 

magnesium 

12 

179 

9 

potassium 

19 

138 

10 

titanium 

22 

27 

11 

phosphorus 

15 

8.6 

12 

carbon 

6 

5.5 

13 

manganese 

25 

3.8 

14 

sulfur 

16 

3.4 

15 

fluorine 

9 

3.3 

16 

chlorine 

17 

2.8 

17 

chromium 

24 

1.5 

18 

barium 

56 

0.75 

*  From  calculations  of  I.  Asimov  (/.  Chem.  Ed.)  31, 
70  (1954)  on  the  data  of  B.  Gutenberg  (Editor)  "Internal 
Constitution  of  the  Earth"  2nd  Ed.  Dover  Publications, 
New  York,  1951,  p.  87.  Reprinted  with  permission  of  the 
publisher. 


of  the  earth,  iron  would  probably  move  into 
second  place  ahead  of  silicon,  and  magnesium 
would  be  fourth.  Thus,  the  exact  order  is  changed 
by  the  sample  (part  of  earth)  chosen.  In  any  of 
the  lists  of  elements,  the  most  abundant  atoms  are 
those  of  elements  having  low  atomic  numbers,  26 
or  less.  All  of  the  elements  beyond  iron  (element 
number  26)  account  for  less  than  0.2%  of  the 
weight  of  the  earth's  crust. 

25-1.5    Availability  off  Elements 

In  our  daily  life  most  of  us  are  more  concerned 
with  the  availability  of  the  elements  than  with 
their  general  abundance  in  earth.  The  air  is  all 
around  us  and  equally  accessible  to  all.  Water  is 
somewhat  more  restricted.  Some  regions  have 
a  surplus  of  water,  whereas  other  regions  are 
deficient.  Even  in  regions  of  abundant  rainfall, 
the  use  of  water  may  be  so  great  that  the  reserves 
become  depleted  gradually.  Thus,  as  the  earth's 
population  increases  and  more  and  more  water 
is  being  used,  water  is  becoming  a  natural  re- 
source to  be  treated  with  respect. 

Many  of  the  metals  used  by  ancient  man — 
copper  (cuprum,  Cu),  silver  (argentum,  Ag), 
gold  (aurum,  Au),  tin  (stannum,  Sn),  and  lead 
(plumbum,  Pb) — are  in  relatively  short  supply. 
Ancient  man  found  deposits  of  the  first  three 
occurring  as  the  elementary  metals.  These  three 
may  also  be  separated  from  their  ores  by  rela- 
tively simple  chemical  processes.  On  the  other 
hand,  aluminum  and  titanium,  though  abundant, 
are  much  more  difficult  to  prepare  from  their 
ores.  Fluorine  is  more  abundant  in  the  earth  than 
chlorine  but  chlorine  and  its  compounds  are 
much  more  common — they  are  easier  to  prepare 
and  easier  to  handle.  However,  as  the  best 
sources  of  the  elements  now  common  to  us  be- 
come depleted,  we  will  have  to  turn  to  the  ele- 
ments that  are  now  little  used. 

During  geological  time,  a  number  of  separat- 
ing and  sorting  processes — melting,  crystalliza- 
tion, solution,  precipitation — have  concentrated 
various  elements  in  local  deposits.  In  these,  the 
elements  tend  to  be  grouped  together  in  rather 
stable  compounds.  These  are  called  minerals. 
Many  of  the  minerals  have  compositions  similar 


442 


THE  CHEMISTRY  OF  EARTH,  THE  PLANETS,  AND  THE  STARS  ■  CHAP.  25 


to  compounds  we  can  make  in  the  laboratory 
but  most  are  not  so  pure.  There  are  large  de- 
posits of  sodium  chloride,  for  example,  appar- 
ently formed  when  ancient  seas  evaporated  in 
locales  where  the  solid  deposits  were  later  pro- 
tected from  the  dissolving  action  of  water.  The 
ocean  itself  is  an  enormous  source  of  sodium 
chloride.  In  contrast,  potassium  salts  have  not 
been  concentrated  in  a  similar  way.  Many  of  the 
metallic  elements  are  concentrated  as  sulfide 
minerals  (for  example,  lead,  PbS;  molybdenum, 
MoS2;  zinc,  ZnS).  Other  elements  occur  in  rather 
concentrated  oxide  deposits  (for  example,  iron, 
Fe203;  manganese,  Mn02).  Large  deposits  of 
carbonates  (for  example,  zinc,  ZnC03;  calcium, 
CaC03)  and  sulfates  (for  example,  barium, 
BaS04)  of  some  metallic  elements  are  known. 
The  minerals  sufficiently  concentrated  to  act  as 
commercial  sources  of  desired  elements  are  called 
ores. 


25-1.6    The  Air  as  a  Source  of  Elements 

We  are  so  used  to  free  air  that  we  do  not  think 
of  oxygen  as  an  important  chemical.  For  ex- 
ample, we  buy  natural  gas  (mostly  methane)  as 
a  fuel  and  burn  it  in  air  to  furnish  heat  to  us. 
If  methane  were  free  in  the  air  and  oxygen  were 
scarce  so  that  we  had  to  purchase  it  we  would 
consider  oxygen  to  be  the  "fuel."  In  either  case, 
the  amount  of  heat  released  would  be  that  for 
the  reaction  represented  by  the  following  equa- 
tion. 

CH<(g)  +  2Q2(g)  — > 

C02(g)  +  2H20(l)  -|-  213  kcal    (5) 

Despite  the  general  availability  of  unlimited 
quantities  of  oxygen  in  the  air,  tremendous  quan- 
tities of  the  pure  gas  are  prepared  annually  for 
industrial  and  medical  use.  Billions  of  cubic  feet 
of  oxygen  gas  are  manufactured  every  year,  by 
liquefaction  of  air  followed  by  fractional  distilla- 
tion to  separate  it  from  nitrogen. 

Ores  of  nitrogen  are  relatively  rare.  The  best 
mineral  is  sodium  nitrate,  NaN03,  found  in  large 
deposits  in  Chile.  We  now  prepare  the  nitrogen 
compounds  we  desire  from  the  nitrogen  of  the 


air.  Thus,  the  air  is  our  best  source  of  oxygen 
and  nitrogen — two  very  important  elements. 


EXERCISE  25-3 

Explain  in  terms  of  energy  the  significance  of  the 
fact  that  a  piece  of  wood  is  stable  in  air  at  room 
temperature  but  if  the  temperature  is  raised,  it 
burns  and  releases  heat. 


25-1.7    The  Age  of  the  Earth 

Part  of  looking  ahead  is  understanding  the  past. 
One  of  the  most  interesting  questions  man  has 
asked  about  the  past  is,  "How  old  is  the  earth?" 
Of  course  we  are  not  sure  there  is  an  answer.  We 
shall  see,  however,  that  ingenious  methods  have 
been  developed  that  date  the  earth's  crust. 
Scientists  generally  accept  the  proposal  that  the 
earth's  crust,  as  we  now  find  it,  has  a  finite  age. 

The  most  reliable  methods  for  establishing  the 
age  of  a  long-lasting  object  (such  as  a  mountain) 
depend  upon  the  presence  of  natural  radioac- 
tivity. The  decay  of  the  radioactive  elements  can 
be  likened  to  a  clock  that  is  partially  unwound. 
By  studying  the  extent  to  which  the  clock  has 
unwound,  we  cannot  tell  the  age  of  the  clock  but 
we  can  measure  how  long  ago  it  was  wound. 

For  example,  consider  the  chemical  composi- 
tion of  a  very  old  crystal  of  pitchblende,  U308. 
We  may  presume  that  this  crystal  was  formed 
at  a  time  when  chemical  conditions  for  its  for- 
mation were  favorable.  For  example,  it  may  have 
precipitated  from  molten  rock  during  cooling. 
The  resulting  crystals  tend  to  exclude  impurities. 
Yet,  careful  analysis  shows  that  every  deposit  of 
pitchblende  contains  a  small  amount  of  lead. 
This  lead  has  accumulated  in  the  crystal,  be- 
ginning at  the  moment  the  pure  crystal  was 
formed,  due  to  the  radioactive  decay  of  the 
uranium. 

The  sequences  of  radioactive  decays  that  lead 
to  lead  are  well-known  and  the  rates  of  decay 
have  been  carefully  measured.  We  shall  consider 
the  sequence  based  upon  the  relatively  slow  de- 
composition of  the  most  abundant  uranium 
isotope,  mass  238  (natural  abundance,  99%): 


SEC.  25-1  |  THE  CHEMISTRY  OF  OUR  PLANET,  EARTH 


443 


"|U  — ►■  2^Th  +  JHe     a-decay,  half-life, 

/1/2  =  4.6  X  109  years    (4) 

The  products  are  an  a-particle  (a  helium  nu- 
cleus), and  a  thorium  isotope  that  is  unstable  and 
that  rapidly  decays  by  emitting  successively  two 
electrons: 

-'&Th  — >-  23,?Pa  +  _?«? 

/3-decay  tm  =  24.1  days  (5) 

23JPa  — >  23,<U  +  _oe 

/3-decay  ty2  =1.14  minutes    (6) 

Thus  we  have  returned  to  an  isotope  of  uranium, 
234U,  but  one  of  half-life  very  much  shorter  than 
that  of  238U.  This  isotope  begins  a  succession  of 
a-decays,  each  moving  the  product  upward  in 
the  periodic  table: 

2S$U    — *■  2$Th  +  £He 

a-decay  tm  =  2.7  X  105  years      (7) 

a  +  *2He 

a-decay  tm  =  8.3  X  104  years      (8) 

(9) 

(70) 


22JRa 


>Rn 


L8Po 


a-decay  til2  =  1.6  X  103  years 
2!?Po  +  4He 

a-decay  tu2  =  3.8  days 
2^Pb  +  4He 

a-decay  txt2  =  3.1  minutes 


(//) 


We  have  at  last  reached  lead — but  2l4Pb  is  itself 
radioactive!  This  isotope  decays  in  a  succession 
of  /3-decays : 


;pb 


|Bi  + 


/3-decay  tm  =  27  minutes    (72) 

2^Bi  — >-  2^Po  +  _°e 

/3-decay  /i/2  =  20  minutes    (13) 

Again  a-decay  occurs,  returning  to  the  element 
lead,  but  to  another  isotope  that  decays  by 
emitting  successively  two  /3-particles: 

2MPo  _^  2,npb  +  4He 

a-decay  /,/2  =  1.5  X  10~4  seconds    (14) 

2|§Pb  — +-  2L°Bi  +  _?e 

/3-decay  tm  =  22  years  (75) 


£Bi  — >-  2i°Po  +  _°e 

/3-decay  tm  =  5  days 


(16) 


This  isotope  of  polonium,  210Po,  again  decays  by 
a-decay,  but  this  time  giving  an  isotope  that  does 
not  decay  further: 

2.oPo  _^  2oepb  +  4He 

a-decay  tU2  =  140  days    (77) 

2^Pb  — >-  not  radioactive        tV2  infinite  (18) 

The  products  in  this  long  sequence  of  reactions 
accumulate  in  the  stable  isotope  of  lead,  206Pb. 
The  amount  of  206Pb  present  depends  upon  how 
long  the  deposit  of  uranium  has  decayed  since  the 
crystal  (/308  was  formed. 

There  is,  fortunately,  a  rather  simple  verifica- 
tion of  the  presumption  that  all  of  the  lead  in 
the  U:iOs  came  from  this  tedious  sequence  of 
nuclear  reactions,  (4)  to  (77).  Lead  ores  that  do 
not  contain  uranium  include  several  isotopes — 
206Pb  makes  up  about  26  %  of  the  total  and  the 
rest  is  204Pb  ( 1 .4  %),  207Pb  (2 1  %),  and  208Pb  (52  %). 
Of  these,  two  other  isotopes  can  be  formed 
through  radioactive  decay  of  some  other  ura- 
nium or  thorium  isotope  by  a  sequence  like  that 
shown  for  238U.  Of  the  four  stable  lead  isotopes, 
only  one  is  not  derived  from  radioactive  decay, 
204Pb.  Hence,  the  ratio  of  the  amount  of  this 
isotope  to  that  of  206Pb  measures  the  amount  of 
206Pb  present  in  excess  of  the  natural  abundance. 
This  excess  must  have  come  from  decay  of  238U. 
If  there  is  no  204Pb  present,  then  all  of  the  206Pb 
came  from  238U. 

Thus,  analysis  of  uranium  minerals  with  the 
aid  of  the  mass  spectrograph  gives  information 
on  the  age  of  the  mineral.  Though  many  different 
half-lives  are  involved  in  forming  the  lead,  only 
the  longest  half-life  (the  rate-determining  step)  is 
of  importance.  Combining  the  lead  content  with 
the  mU  half-life  provides  estimates  of  mineral 
ages  in  the  range  of  five  billion  years. 

What  have  we  learned  in  this  estimate?  Surely 
we  can  say  the  age  of  the  earth  cannot  be  shorter 
than  5  X  109  years.  That  was  when  the  uranium 
mineral  clock  was  wound — but  the  clock  could 
be  much  older.  To  evaluate  this  number  further, 
we  must  look  for  other  types  of  data. 

Fortunately,  there  are  other  radioactive  ele- 
ments in  nature  that  give  similar  bases  for  esti- 
mates. As  a  second  example,  potassium  in  nature 


444 


THE    CHEMISTRY    OF    EARTH,    THE    PLANETS,    AND    THE    STARS    |    CHAP.    25 


includes  one  radioactive  isotope,  ™K,  which  de- 
cays by  capturing  an  electron  into  its  nucleus. 
The  product  is  ^Ar,  a  stable  isotope: 

fsK.  +  -?e  — ►•  JsAr    electron  capture 

tm  =  1.5  X  1010  years    (79) 

Once  again,  the  ratio  of  the  abundance  in  a 
crystal  of  the  two  isotopes,  ^Ar/^K,  provides  a 
clue  to  the  age  of  the  crystal.  Mica  is  a  mineral 
that  has  been  much  studied  in  this  type  of  min- 


eral age  estimation.  Such  estimates  also  tend  to 
date  the  minerals  as  a  few  billion  years  old. 
Similar  age  figures  are  obtained  from  the  natural 
radioactivity  of  rubidium. 

In  conclusion,  the  agreement  of  all  of  these 
methods  based  upon  radioactive  decay  furnishes 
a  strong  clue  that  the  earth's  crust  as  we  know 
it  today  was  formed  about  five  billion  years  ago. 
What  preceded  is  a  subject  of  intense  interest  and 
monumental  disagreement. 


25-2    THE  CHEMISTRY  OF  THE  PLANETS 


Our  solar  system  consists  of  the  Sun,  the  planets 
and  their  moon  satellites,  asteroids  (small  plan- 
ets), comets,  and  meteorites.  The  planets  are 
generally  divided  into  two  categories:  Earth-like 
(terrestrial)  planets — Mercury,  Venus,  Earth, 
and  Mars;  and  Giant  planets — Jupiter,  Saturn, 
Uranus,  and  Neptune.  Little  is  known  about 
Pluto,  the  most  remote  planet  from  Earth. 

As  space  exploration  begins,  we  can  look  for- 
ward to  a  vast  multiplication  of  our  present 
knowledge  of  the  planets.  Conceivably  we  shall 
be  analyzing  samples  of  the  moon  within  this 
decade.  The  distances  to  the  other  planets  are 
such  that  voyages  of  the  order  of  a  few  months 
suffice  to  reach  them.  Again  information  will 
accumulate  rapidly. 

Contrast  our  position  ten  years  ago.  There  lay 
the  planets — ours  to  see  but  not  to  touch.  And 
not  to  see  well,  either.  No,  we  must  view  the 
planets  through  the  gauze  curtain  of  the  at- 
mosphere. Small  wonder  we  know  rather  little 
about  our  nearest  neighbors,  despite  our  eager- 
ness to  pry.  Yet  we  do  know  enough  about  the 
other  planets  to  say  that  their  surface  environ- 
ments are  totally  unlike  that  on  earth.  Though 
much  of  what  is  today  thought  to  be  true  will  be 
known  to  be  false  tomorrow,  it  provides  stimu- 
lating reading. 

Table  25-IV  begins  this  survey  with  a  com- 
parison of  mass,  radius,  and  density  of  the 
planets  and  the  sun.  These  data  are  probably  the 
most  reliable  facts  known  about  the  planets  since 


they  are  deduced  from  the  orbital  movements  in 
the  solar  system. 


Table  25-IV 


DATA    ON    THE    SOLAR    SYSTEM 


RELATIVE 
MASS 


RADIUS 

(kilometers) 


DENSITY 

(g/ml) 


Sun 

3.32  X  106 

695 

X  103 

1.41 

Mercury 

0.05 

2.5 

X  10* 

5.1 

Venus 

0.81 

6.2 

X  10s 

5.0 

Earth 

(1.00) 

6.371 

X  103 

5.52 

Mars 

0.11 

3.4 

X  10* 

3.9 

Jupiter 

3.18  X  102 

71 

X  103 

1.33 

Saturn 

95 

57 

X  10* 

0.71 

Uranus 

14.6 

25.8 

X  io» 

1.27 

Neptune 

17.3 

22.3 

X  103 

2.22 

Pluto 

0.03? 

2.9 

X  103 

2? 

25-2.1    Meteorites 

We  do  have  some  direct  evidence  concerning  the 
composition  of  solid  matter  outside  the  earth's 
atmosphere.  Occasionally  a  piece  of  solid,  a 
meteorite,  falls  through  the  atmosphere  to  give 
us  a  real  sample  for  analysis.  Such  a  piece  of 
solid  may  become  hot  because  of  its  rapid  move- 
ment through  the  air  and  therefore  glow  brightly. 
Many  meteors  burn  up  or  evaporate  but  some 
are  large  enough  to  survive  and  reach  the  earth's 
surface.  These  we  can  examine  and  analyze. 


SEC.    25-2    I    THE    CHEMISTRY    OF    THE    PLANETS 


445 


Meteorites  are  of  two  kinds:  stony  meteorites 
that  are  rock-like  in  character,  and  metallic 
meteorites  that  consist  of  metallic  elements.  The 
kinds  of  substances  in  the  stony  meteorites  are 
very  much  like  the  substances  in  the  crust  of  the 
earth,  if  we  allow  for  the  fact  that  the  meteors 
could  not  bring  gases  or  liquids  with  them.  We 
feel  that  the  other  type,  the  metallic  meteors,  give 
valuable  clues  about  the  nature  of  the  earth's 
central  core.  Experts  have  long  believed  that 
these  meteorites  are  fragments  from  exploded 
planets  that,  perhaps,  resembled  the  earth. 

Whether  this  is  true  or  not,  meteorites  give  us 
one  definite  piece  of  information.  Isotopic  analy- 
sis shows  that  each  element  in  a  meteorite  has 
the  same  isotopes  in  the  same  percentages  that 
this  same  element  has  on  earth.  The  accepted 
explanation  for  this  fact  is  that  meteorites  and 
the  earth  share  a  common  origin  and  that  they 
became  separated  after  the  elements  were  cre- 
ated. 


these  are  the  substances  that  can  be  detected  and 
other  gases  are  undoubtedly  present  as  well. 

There  appears  to  be  a  correlation  between  the 
mass  of  the  planets  and  the  mass  and  composi- 
tion of  their  atmospheres.  Generally,  only  those 
planets  of  high  mass  were  able  to  retain  much 
of  their  atmospheres.  Nitrogen,  hydrogen,  and 
helium  are  probably  abundant,  though  not  yet 
detected,  on  the  heavier  planets.  Table  25-V  also 
reveals  a  considerable  range  in  the  surface  tem- 
peratures of  the  planets.  The  higher  tempera- 
tures on  the  terrestrial  planets  also  contributed 
to  the  loss  of  their  atmospheres. 


EXERCISE  25-4 

Nitrogen  is  considered  to  be  a  likely  constituent 
of  the  atmosphere  of  Jupiter,  though  it  is  un- 
detected as  yet.  As  a  chemist,  would  you  expect 
oxygen  also  to  be  an  important  constituent  of 
Jupiter's  CH4-NH3  atmosphere? 


25-2.2    The  Planetary  Atmospheres 

Through  spectroscopic  observations  and  some- 
times tenuous  deductions  there  has  accumulated 
a  significant  picture  of  the  makeup  of  the  plane- 
tary atmospheres.  Doubt  pervades  much  of  this 
picture,  yet  it  represents  our  starting  place  in 
knowledge  as  we  venture  outside  our  own  atmos- 
phere for  the  first  time.  Table  25-V  summarizes 
a  part  of  this  information — the  maximum  surface 
temperatures  and  the  chemical  compositions. 
Naturally,  these  compositions  are  incomplete: 


Table  25-V 

THE  PLANETARY  ATMOSPHERES 


MAX.  SURFACE 

SOME  OF  THE 

PLANET 

TEMPERATURE 

(°C) 

GASES  PRESENT 

Venus 

430 

co2 

Mercury 

350 

none 

Earth 

60 

02,  N2,  H20,  etc. 

Mars 

30 

N2,  C02,  HsO 

Jupiter 

-138 

CHt,  NH3 

Saturn 

-153 

CH4,  NH3 

Uranus 

-184 

CH4 

Neptune 

-200 

CH4 

The  composition  of  the  planetary  atmospheres 
is  fairly  constant.  This  is  indeed  surprising  in 
view  of  the  fact  that  molecules  such  as  methane, 
ammonia,  and  carbon  dioxide  are  easily  decom- 
posed by  the  ultraviolet  radiation  from  the  sun. 
Presumably  other  reactions  regenerate  those  sub- 
stances that  are  light  sensitive. 

The  atmosphere  of  Venus  is  chiefly  carbon 
dioxide  in  a  concentration  much  higher  than 
that  found  on  Earth.  Surprisingly,  no  evidence 
has  been  found  for  carbon  monoxide,  though 
ultraviolet  light  decomposes  C02  to  form  CO. 
The  atmosphere  of  Mars  is  thought  to  be  largely 
nitrogen  (around  °8  %)  and  some  carbon  dioxide. 

Recent  space-probe  and  earth-based  spectro- 
scopic studies  of  the  planet  Venus  suggest  how 
much  remains  to  be  learned  about  the  other 
planets.  Earlier  estimates  of  the  surface  tempera- 
ture of  Venus  placed  it  near  60°C.  The  more 
detailed  studies  show,  however,  that  two  charac- 
teristic temperatures  can  be  identified,  —  40°C 
and  430°C.  The  lower  temperature  is  attributed 
to  light  emitted  from  high  altitude  cloud  tops. 
The  higher  temperature  is  likely  to  be  the  average 
surface  temperature. 


446 


THE    CHEMISTRY    OF    EARTH,    THE    PLANETS,    AND    THE    STARS    |    CHAP.    25 


CHjf ,  NH3 ,  H^,  He    atmosphere 

Liquid  CH4 

Ice  and  solid  N'Hj 
Rock  and  mef-al 


v  %  ■ 


1. 


'--■ 


Liquid  (?) 

Solid   hydrogen. 

Ice,  high  pressure  form 
Solid  hydrogen,  high 

pressure  form 


"•:■  :• 


. 


These  latest  findings  provide  an  estimate  that 
the  atmospheric  pressure  at  the  ground  surface 
is  10  ±  2.5  atmospheres.  This  is  the  pressure  felt 
by  a  deep-sea  diver  at  a  depth  of  300  feet,  near 
the  limit  of  human  endurance.  The  atmospheric 
temperature  at  the  surface  averages  around 
430°C,  rising  in  some  locales  possibly  as  high  as 
550°C  (near  the  softening  temperature  of  ordi- 
nary glass)  whereas  in  other  regions,  cool  breezes 
may  blow  at  temperatures  below  350°C  (near 
the  boiling  point  of  mercury  and  the  melting 
point  of  lead).  Compare  these  extremes  to  the 
narrow  range  in  which  humans  can  survive  com- 
fortably. If  you  have  ever  experienced  a  desert 
temperature  as  high  as  43°C  (1 10°F)  or  a  winter 
as  cold  as  -35°C  (-30°F),  you  will  realize  that 
Venus  will  not  be  plagued  by  too  many  tourists 
from  Earth. 

The  giant  planets  possess  low  surface  tempera- 
tures and  have  atmospheres  that  extend  several 
thousand  miles.  The  markings  on  Jupiter,  the 
largest  planet,  consist  of  cloud  formations  com- 
posed of  methane  containing  a  small  amount  of 
ammonia.  The  atmosphere  of  Jupiter  absorbs  the 
extreme  red  and  infrared  portions  of  the  spec- 
trum. These  absorptions  correspond  to  the  ab- 
sorption spectra  of  ammonia  and  methane, 
suggesting  the  presence  of  these  gases  in  Jupiter's 


Fig.  25-2.  Two  proposed  structures  of  Jupiter. 


atmosphere.  Free  hydrogen  and  helium  are  also 
thought  to  be  present  but  this  is  difficult  to 
verify.  Estimates  of  the  average  molecular  weight 
of  the  gases  in  Jupiter's  atmosphere  are  around 
3.  The  atmosphere  of  Jupiter  includes  belts  or 
bands  that  seem  to  be  indicative  of  climate 
variations  similar  to  our  own  equatorial,  tem- 
perate, and  polar  climates. 

Jupiter  has  a  single  permanent  marking  called 
the  Great  Red  Spot.  This  spot  is  oval  in  shape, 
about  30,000  miles  long  and  about  7,000  miles 
wide.  The  coloring  is  thought  to  result  from  light 
reflected  from  the  different  layers  in  the  planet's 
atmosphere.  Theories  concerning  the  origin  of 
this  spot  and  of  Jupiter's  brightly  colored  at- 
mospheric belts  are  imaginative,  numerous,  and 
in  general  disagreement. 

25-2.3    The  Planetary  Lithospheres 

Needless  to  say,  the  extreme  difficulty  we  experi- 
ence in  probing  the  composition  of  the  earth 
beneath  us  suggests  that  little  is  known  about 
the  inner  composition  of  the  planets.  The  evi- 
dence available  is  indirect  (average  density,  sur- 


SEC.    25-3    I    THE    STARS 


447 


.-■'\bH 

¥# 

£       NH     CaH 

;  OH     YO 

•  AlO    ZrO 
TiO 

SH 

CH 
Cx 

CN 

NH 
CN 

r,'J 

.-' 

^$H 

AlO 

SiH 

CaH 

Tio 

Yo 

Zro 

face  composition,  etc.)  and  the  interpretations 
are  conflicting.  Nevertheless,  we  present  in  Fig- 
ure 25-2  two  proposals  that  have  been  offered  as 
possible  structures  for  the  planet  Jupiter.  A  va- 
riety of  responses  are  evoked  by  these  startling 
proposals:  dismay— that  the  available  data  are  so 
inconclusive;  discouragement— that  our  knowl- 
edge is  so  incomplete;  anticipation — for  the  near 
future  when  some  of  the  uncertainties  will  be 
removed ;  sympathy— for  the  poor  astronaut  who 
is  to  step  out  of  his  space  vehicle  to  plant  his 
flag  in  an  unfriendly  sea  of,  alas,  liquid  methane. 

25-2.4    The  Sun 

The  surface  temperature  of  the  sun  is  about 
5500°C.  Moving  inward  from  the  surface,  the 
temperature  rises,  probably  above  one  million 
degrees.  At  these  large  temperatures  the  competi- 
tion between  opposing  tendencies  toward  lowest 
energy  (favoring  molecules)  and  toward  highest 
randomness  (favoring  atoms)  is  dominated  by 
the  randomness  factor.  As  a  result,  only  the 
simplest  molecules  are  to  be  expected  there. 


Fig.  25-3.  Some  molecules  detected  in  the  solar  at- 
mosphere and  the  elements  they  contain. 

The  solar  spectrum  is,  of  course,  as  well 
studied  as  our  planetary  atmosphere  will  permit. 
More  information  will  be  forthcoming  as  spectra 
from  man-made  satellites  are  recorded  above  the 
atmosphere.  At  this  time,  the  spectra  of  many 
diatomic  molecules  have  been  detected.  These 
are  not  the  familiar,  chemically  stable  molecules 
we  find  on  the  stockroom  shelf.  These  are  the 
molecules  that  are  stable  on  a  solar  stockroom 
shelf.  Figure  25-3  shows  some  of  these  and  the 
location  in  the  periodic  table  of  the  elements 
represented. 

Inside  the  sun,  thermal  energies  are  sufficient 
to  destroy  all  molecules  and  to  ionize  the  atoms. 
These  ions  emit  their  characteristic  line  spectra 
and  tens  of  thousands  of  lines  are  observed.  The 
lines  that  have  been  analyzed  show  the  existence 
of  atoms  ionized  as  far  as  0+5,  Mn+12,  and  Fe+13. 
At  this  time,  over  sixty  of  the  elements  have  been 
detected  in  the  sun  through  their  spectral  emis- 
sions and  absorptions. 


25-3    THE  STARS 

Our  knowledge  of  the  stars  and  of  space  is  en- 
tirely obtained  through  spectroscopy  and  will  be 
for  the  forseeable  future.  That  is  not  at  all  to  say 
we  know  and  shall  know  little  of  the  other  gal- 
axies. It  is  to  say  that  our  information  will  be 
incomplete.  But  man  is  opportunistic  and  clever 
— small  pieces  of  a  spectroscopic  jawbone  may 


permit  us  to  build  quite  a  reasonable  likeness  of 
the  Universe. 

As  evidence  that  this  is  so,  consider  that  the 
element  helium  was  detected  in  the  sun  before 
it  was  found  on  earth!  Though  oxygen  contains 
0.2%  of  the  oxygen- 18  isotope  on  earth,  it,  too, 
was  first  detected  in  a  solar  spectrum.  Two 


148 


THE    CHEMISTRY    OF    EARTH,    THE    PLANETS,    AND    THE    STARS    I    CHAP.    25 


chemical  species  whose  spectra  were  first  pro- 
vided through  photographs  of  comets  are  CO+ 
and  C3.  Let  us  look  briefly  then,  and  with  respect, 
at  the  astronomers'  knowledge  of  stellar  chem- 
istry. 

25-3.1    Stellar  Atmospheres 

Our  sun  is,  of  course,  a  star.  It  is  a  relatively  cool 
star  and,  as  such,  contains  a  number  of  diatomic 
molecules  (see  Figure  25-3).  There  are  many 
stars,  however,  with  still  lower  surface  tempera- 
tures and  these  contain  chemical  species  whose 
presence  can  be  understood  in  terms  of  the 
temperatures  and  the  usual  chemical  equilibrium 
principles.  For  example,  as  the  star  temperature 
drops,  the  spectral  lines  attributed  to  CN  and 
CH  become  more  prominent.  At  lower  tempera- 
tures, TiO  becomes  an  important  species  along 
with  the  hydrides  MgH,  SiH,  and  A1H,  and 
oxides  ZrO,  ScO,  YO,  CrO,  AlO,  and  BO. 

Detailed  consideration  of  the  chemical  equi- 
libria among  these  species  provides  evidence  of 
the  presence  of  other  molecules  that  cannot  be 
observed  directly.  The  chemical  properties  of  the 
molecules  mentioned  provide  a  firm  basis  for 
predicting  the  presence  and  concentrations  of 
such  important  molecules  as  H2,  CO,  02,  N2, 
and  NO.  Thus  the  faint  light  by  which  we  view 
the  distant  stars  is  rich  with  information.  We 
need  but  learn  how  to  read  it. 

25-3.2     Interstellar  Space 

In  addition  to  the  stars,  with  their  characteristic 


light  emissions,  the  space  between  them  is  part 
of  the  astronomical  spectroscopy  laboratory. 
Light  from  a  distant  star  must  traverse  fantastic 
distances  to  reach  our  telescopes  and  the  absorp- 
tion of  this  light  by  the  most  minute  concentra- 
tions of  atoms  and  molecules  in  space  becomes 
important  and  detectable.  Absorption  spectra 
have  shown  that  diatomic  molecules  such  as  CH, 
CN,  and  CH+  are  present  at  an  average  concen- 
tration of  about  one  molecule  per  1000  liters. 
These  molecules  are  probably  concentrated  in 
"clouds"  with  one  molecule  per  100  liters. 


EXERCISE  2S-S 

Calculate  the  volume  in  liters  of  a  sphere  of 
radius  6400  kilometers  (the  radius  of  the  earth). 
How  many  grams  of  oxygen  would  be  needed 
to  fill  this  volume  to  a  concentration  of  one 
molecule  per  1000  liters? 


In  addition  to  these  molecules,  atoms  are 
present,  as  shown  by  absorptions  of  Ca,  Na,  K, 
Fe,  and  other  atoms.  There  are  some  absorptions 
that  have  not  been  identified  but  these  may  be 
due  to  small  solid  particles.  How  these  particular 
molecules  and  atoms  happen  to  be  present  in 
these  almost  nonexistent  "clouds"  and  what 
other  molecules  and  atoms  are  there,  yet  to  be 
detected,  is  a  matter  for  wondering.  But  wonder- 
ing is  at  once  the  pleasure  and  the  driving  force 
of  science. 


APPENDIX 


l 


A  DESCRIPTION  OF  A 
BURNING  CANDLE 


A  drawing  of  a  burning  candle  is  shown1  in  Figure 
Al-1.  The  candle  is  cylindrical  in  shape2  and  has  a 
diameter3  of  about  f  inch.  The  length  of  the  candle 
was  initially  about  eight  inches4  and  it  changed 
slowly5  during  observation,  decreasing  about  half 
an  inch  in  one  hour6.  The  candle  is  made  of  a 
translucent7,  white8  solid9  which  has  a  slight  odor10 
and  no  taste11.  It  is  soft  enough  to  be  scratched  with 
the  fingernail12.  There  is  a  wick13  which  extends  from 
top  to  bottom14  of  the  candle  along  its  central  axis15 
and  protrudes  about  half  an  inch  above  the  top  of 
the  candle16.  The  wick  is  made  of  three  strands  of 
string  braided  together17. 

A  candle  is  lit  by  holding  a  source  of  flame  close 
to  the  wick  for  a  few  seconds.  Thereafter  the  source 
of  flame  can  be  removed  and  the  flame  sustains  itself 
at  the  wick18.  The  burning  candle  makes  no  sound19. 
While  burning,  the  body  of  the  candle  remains  cool 
to  the  touch20  except  near  the  top.  Within  about 
half  an  inch  from  the  top  the  candle  is  warm21  (but 
not  hot)  and  sufficiently  soft  to  mold  easily22.  The 
flame  flickers  in  response  to  air  currents23  and  tends 
to  become  quite  smoky  while  flickering24.  In  the 
absence  of  air  currents,  the  flame  is  of  the  form 
shown  in  Figure  Al-1,  though  it  retains  some  move- 
ment at  all  times23.  The  flame  begins  about  £  inch 
above  the  top  of  the  candle26  and  at  its  base  the  flame 
has  a  blue  tint27.  Immediately  around  the  wick  in  a 
region  about  \  inch  wide  and  extending  about  \  inch 
above  the  top  of  the  wick28  the  flame  is  dark29.  This 
dark  region  is  roughly  conical  in  shape30.  Around 
this  zone  and  extending  about  half  an  inch  above 
the  dark  zone  is  a  region  which  emits  yellow  light31, 
bright   but  not   blinding32.   The  flame   has   rather 


JDasiA 


Ccr&r^e&uL,  Mf^ou^—f^^ 


T3supfut  yt&trus- 


&Ma>. 


CUnnot  /£  ^vcJi 


'S;M^m 


^ 


Fig.  Al-1.  A  burning  candle. 


sharply  defined  sides33,  but  a  ragged  top34.  The  wick 
is  white  where  it  emerges  from  the  candle35,  but  from 
the  base  of  the  flame  to  the  end  of  the  wick36  it  is 
black,  appearing  burnt,  except  for  the  last  r\  inch 
where  it  glows  red37.  The  wick  curls  over  about  J 

449 


450 


A    DESCRIPTION    OF    A    BURNING    CANDLE    I    APP.    1 


inch  from  its  end38.  As  the  candle  becomes  shorter, 
the  wick  shortens  too,  so  as  to  extend  roughly  a 
constant  length  above  the  top  of  the  candle39.  Heat  is 
emitted  by  the  flame40,  enough  so  that  it  becomes 
uncomfortable  in  ten  or  twenty  seconds  if  one  holds 
his  finger  \  inch  to  the  side  of  the  quiet  flame41  or 
three  or  four  inches  above  the  flame42. 

The  top  of  a  quietly  burning  candle  becomes  wet 
with  a  colorless  liquid43  and  becomes  bowl  shaped44. 
If  the  flame  is  blown,  one  side  of  this  bowl-shaped 
top  may  become  liquid,  and  the  liquid  trapped  in  the 
bowl  may  drain  down  the  candle's  side45.  As  it 
courses  down,  the  colorless  liquid  cools46,  becomes 
translucent47,  and  gradually  solidifies  from  the  out- 
side48, attaching  itself  to  the  side  of  the  candle49.  In 
the  absence  of  a  draft,  the  candle  can  burn  for  hours 
without  such  dripping60.  Under  these  conditions,  a 
stable  pool  of  clear  liquid  remains  in  the  bowl-shaped 
top  of  the  candle61.  The  liquid  rises  slightly  around 
the  wick62,  wetting  the  base  of  the  wick  as  high  as 
the  base  of  the  flame63. 


Several  aspects  of  this  description  deserve  specific 
mention.  Compare  your  own  description  in  each  of 
the  following  characteristics. 

(1)  The  description  is  comprehensive  in  qualitative 
terms.  Did  you  include  mention  of  appearance? 


smell?  taste?  feel?  sound?  (Note:  A  chemist 
quickly  becomes  reluctant  to  taste  or  smell  an 
unknown  chemical.  A  chemical  should  be  con- 
sidered to  be  poisonous  unless  it  is  known  not 
to  be!) 

(2)  Wherever  possible,  the  description  is  stated 
quantitatively.  This  means  the  question  "How 
much?"  is  answered  (the  quantity  is  specified). 
The  remark  that  the  flame  emits  yellow  light  is 
made  more  meaningful  by  the  "how  much"  ex- 
pression, "bright  but  not  blinding."  The  state- 
ment that  heat  is  emitted  might  lead  a  cautious 
investigator  who  is  lighting  a  candle  for  the  first 
time  to  stand  in  a  concrete  blockhouse  one  hun- 
dred yards  away.  The  few  words  telling  him 
"how  much"  heat  would  save  him  this  overpre- 
caution. 

(3)  The  description  does  not  presume  the  importance 
of  an  observation.  Thus  the  observation  that  a 
burning  candle  does  not  emit  sound  deserves  to 
be  mentioned  just  as  much  as  the  observation 
that  it  does  emit  light. 

(4)  The  description  does  not  confuse  observations 
with  interpretations.  It  is  an  observation  that  the 
top  of  the  burning  candle  is  wet  with  a  colorless 
liquid.  It  would  be  an  interpretation  to  state  the 
presumed  composition  of  this  liquid. 


APPENDIX 


2 


RELATIVE  STRENGTHS  OF 
ACIDS 

IN  AQUEOUS   SOLUTION   AT  ROOM   TEMPERATURE 

All  ions  are  aquated 


ACID 


STRENGTH 


REACTION 


KA 


perchloric  acid 

hydriodic  acid 

hydrobromic  acid 

hydrochloric  acid 

nitric  acid 

sulfuric  acid 

oxalic  acid 

sulfurous  acid  (S02  +  H20) 

hydrogen  sulfate  ion 

phosphoric  acid 

ferric  ion 

hydrogen  telluride 

hydrofluoric  acid 

nitrous  acid 

hydrogen  selenide 

chromic  ion 

benzoic  acid 

hydrogen  oxalate  ion 

acetic  acid 

aluminum  ion 

carbonic  acid  (C02  +  H20) 

hydrogen  sulfide 

dihydrogen  phosphate  ion 

hydrogen  sulfite  ion 

ammonium  ion 

hydrogen  carbonate  ion 

hydrogen  telluride  ion 

hydrogen  peroxide 

monohydrogen  phosphate  ion 

hydrogen  sulfide  ion 

water 

hydroxide  ion 

ammonia 


very  strong 


::yery  strong 


strong 


I 
J 

weak 


weak 


weak 

I 
» 

very  weak 


very  weak 


HCIO4 

HI 

HBr 

HC1 

HNO3 

H2SO4 

HOOCCOOH 

H2SO3 

HSO4- 

H3PO4 

Fe(H20)6+3 

H2Te 

HF 

HN02 

H2Se 

Cr(H20)6+3 

C«H5COOH 

HOOCCOO- 

CH3COOH 

A1(H20)6+3 

H2C03 

H2S 

H2P04- 

HSO3- 

NH4+ 

HCO3- 

HTe" 

H202 

HPO4-2 

HS" 

H20 

OH- 

NH3 


H+  + 
H+  + 
H+  + 
H+  + 

H+  + 
H+  + 


H* 
FT 
H^ 
H- 
H"» 
W 
H- 
H- 


H+  + 


H" 
H- 
H^ 
H^ 
H^ 

W 

H^ 

H^ 
HJ 
HJ 
H* 
1-T 
H* 
H' 


CIO4- 

I- 

Br- 

ci- 

NO3- 

HSO4- 

HOOCCOO- 

HSO3- 

SO4-2 

H2P04" 

Fe(H20)5(OH)+2 

HTe" 

F" 

NOr 

HSe- 

Cr(H20)5(OH)+2 

QH6COO- 

OOCCOO-2 

CH3COO- 

Al(H20)6(OH)+J 

HCO,- 

HS- 

HPO4-1 

SO,"2 

NH, 

CO3-2 

Te"2 

HOr 

po4-» 

s_2 

OH- 
O2 

NH2 


[H+][OH"]  = 


very  large 
very  large 
very  large 
very  large 
very  large 
large 

5.4  X  10-2 
1.7  X  10~2 
1.3  X  10- 
7.1  X  10- 

6.0  X  10- 

2.3  X  10- 
6.7  X  10- 

5.1  X  10"< 

1.7  X  10"« 

1.5  X  10- 

6.6  X  10" 

5.4  X  10- 

1.8  X  10"s 
1.4  X  10- 
4.4  X  10-' 

1.0  x  10-7 

6.3  X  10"8 

6.2  X  lO-8 

5.7  X  lO"10 
4.7  X  10-11 
1.0  X  10"11 

2.4  X  lO"12 
4.4  X  10-" 

1.3  X  10"13 

1.0  X  10-" 
<10"36 
very  small 


451 


APPENDIX 


3 


STANDARD  OXIDATION 
POTENTIALS  FOR 
HALF-REACTIONS 

IONIC  CONCENTRATIONS,  1  M  IN  WATER  AT  25 °C 

All  ions  are  aquated 


HALF-REACTION 


E°  (volts) 


Very  strong 

reducing 

agents 


Li  - 

Rb  - 
K- 
Cs- 
Ba- 

Sr  ■ 
Ca- 
Na- 
Mg- 
Al- 
Mn 
H,fo)  +  20H-  ■ 
Zn- 
Cr 
H2Te 
2Ag  +  S~2 
Fe 
H*{g)  ■ 
Cr-2 
H2Se 
Co 
Ni 
Sn 
Pb 

H2S(tf) 

Sn4* 

Cu+ 

S02(<?)  +  2H20 

Cu 

Cu 

21- 

H,02 


e~  +  Lr 
e-  +  Rb+ 
er  +K+ 
e-  +Cs+ 
2e-  +  Ba41 
2*-  +  Sr« 
2e-  -fCa4* 
e-  +  Na+ 
2e~  +  Mg42 
3e-  +  Al4* 
2e-  +  Mn41 
2e~  +  2H,0 
2e-  +  Zn-5 
3e-  +  Cr-3 
2e-  +  Te  +  2H+ 

•  2e-  +  Ag:S 

■  2*"  +  Fe+2 

2e~  +  2H+  (10-7  Af) 

•  e~  +  Cr+3 

-  2e~  +  Se  +  2H+ 

■  2e~  +  Co~- 

■  2e~  +  Ni^ 

-  2e-  +  Sn-5 

-  2<r  +  Pb*2 

•  2e-  +  2H+ 

-  2e~  +  S  +  2H+ 

-  2e-  +  Sn4* 

-  er  +  Cu+2 

-  2e~  +  SOr*  +  4H+ 

-  2e~  +  Cu'2 

-  er  +  Cu+ 

-  2e-  +  I, 

-  2e~  +  0,fa)  +  2H+ 


3.00 

Very  weak 

2.92 

oxidizing 

2.92 

agents 

2.92 

2.90 

2.89 

2.87 

1 

2.71 

2.37 

1.66 

1.18 

0.83 

0.76 
0.74 

O 

0.72 
0.69 

9. 

N* 

5" 

3 

C.44 
0.414 

0.41 
0.40 

(jo 

3- 

0.28 

0.25 

S3 

g 

(A 

0.14 

0.13 

0.00 

-0.14 

-0.15 

-0.15 

-0.17 

-0.34 

-0.52 

-0.53 

L 

-0.68 

r 

452 


APP.    3    I    STANDARD    OXIDATION    POTENTIALS    FOR    HALF-REACTIONS 


453 


APPENDIX  3— (Continued) 


HALF-REACTION 


E°  (VOLTS) 

-0.77 

-0.78 

-0.78 

O 

-0.79 

g. 

-0.80 

3 

-0.815 

05 

-0.96 

3 

-1.00 
-1.06 

3 
OS 

3" 

-1.23 

3 
O 

-1.28 

g 

-1.33 

l 

-1.36 
-1.50 

>f 

-1.52 

Very  strong 

-1.77 

oxidizing 

-2.87 

agents 

Very  weak 

reducing 

agents 


Fe-- 

N02((?)  +  H20 

Hg(0 

Hg(0 

Ag 

HsO 

NOfo)  +  2H20 

Au  +  4C1- 

2Br- 

H20 

Mn+2  +  2H20 

2Cr+>  +  7H20 

2C1- 

Au 

Mn+2  +  4H20 

2H20 

2F- 


e~  +  Fe*3 

e~  +  NCV  +  2H+ 

2e~  +  Hg+2 

e~  +  *Hg2« 

e~  +  Ag+ 

2e~  +  iO^g)  +  2H+  (10-7  M) 

3e-  +  N03"  +  4H+ 

3e"  +  AuCU- 

2e~  +  Br2(0 

2e~  +  hOtig)  +  2H+ 

2e~  +  MnO,  +  4H+ 

6e~  +  Cr2CV2  +  14H+ 

2e~  +  a,(g) 

3e~  +  Au+3 

5er  +  MnCV  +  8H+ 

2e~  +  H202  +  2H+ 

2e~  +  Ft(g) 


APPENDIX 


4 


NAMES,  FORMULAS, 
AND  CHARGES  OF  SOME 
COMMON  IONS 


POSITIVE   IONS   (CATIONS) 


aluminum 

ammonium 

barium 

calcium 

chromium  (II),  chromous 

chromium  (III),  chromic 

cobalt  (II),  cobaltous 

copper  (I),*  cuprous 

copper  (II),  cupric 

hydrogen,  hydronium 

iron  (II),*  ferrous 

iron  (III),  ferric 

lead 

lithium 

magnesium 

manganese  (II),  manganous 

mercury  (I),*  mercurous 

mercury  (II),  mercuric 

potassium 

silver 

sodium 

strontium 

tin  (II),*  stannous 

tin  (IV),  stannic 

zinc 


Al+» 

NH4+ 

Ba+2 

Ca+2 

Cr*2 

Cr+J 

Co"2 

Cu" 

Cu+2 

H+,  H3CT 

Fe+2 

Fe+3 

Pb+2 

Li+ 

Mg« 

Mn*2 

Hgo"2 

Hg"2 

K+ 

Ag+ 

Na+ 

Sr2 

Sn"2 

Sir" 

Zn"2 


NEGATIVE    IONS   (ANIONS) 


acetate  CH3COO- 

bromide  Br- 

carbonate  CO.T2 

hydrogen  carbonate  ion,  bicarbonate  HCO3- 

chlorate  CIO3- 

chloride  Cl- 

chlorite  CKV" 

chromate  Cr04-2 

dichromate  Cr^Or-2 

fluoride  F~ 

hydroxide  OH- 

hypochlorite  CIO- 

iodide  I~ 

nitrate  NO3- 

nitrite  NOr 

oxalate  CjOi-2 

hydrogen  oxalate  ion,  binoxalate  HGO,- 

perchlorate  ClOr 

permanganate  Mn04- 

phosphate  POj-3 

monohydrogen  phosphate  HPO<-2 

dihydrogen  phosphate  H2POr 

sulfate  SOi-2 

hydrogen  sulfate  ion,  bisulfate  HSO," 

sulfide  S-2 

hydrogen  sulfide  ion,  bisulfide  HS" 

sulfite  S03-a 

hydrogen  sulfite  ion,  bisulfite  HSO3- 


*  Aqueous  solutions  are  readily  oxidized  by  air. 

Note:  In  ionic  compounds  the  relative  number  of  positive  and  negative  ions  is  such  that  the  sum  of  their  electric 
charges  is  zero. 
454 


Index 


Absolute  temperature,  57 

Kelvin  scale,  58 
Absolute  zero,  58 

Abundance   of  elements   in   earth's 
crust,  see  Elements,  abundance 
in  earth's  crust 
Acetaldehyde  structure,  332 
Acetamide,  338 
Acetanilide,  344 
Acetic  acid 

in  biochemistry,  428 

structure,  333 
Acetone 

solubility  of  ethane  in,  313 

structure,  335 
Acetylene,  43 
Acid-base 

Br0nsted-Lowry  theory,  194 

contrast  definitions,  194 

indicators,  190 

reactions,  188 

titrations,  188 
Acids,  183 

aqueous,  179 

carboxylic,  334 

derivatives  of  organic,  337 

equilibrium  calculations,  192 

experimental  introduction,  183 

names  of  common,  183 

naming  of  organic,  339 

properties  of,  183 

relative  strengths,  192,  451 

strength  of,  190 

summary,  185 

weak,  190,  193 
Actinides,  414 
Actinium 

electron  configuration,  415 

oxidation  number,  414 
Activated  complex,  134 
Activation  energy,  132,  134,  369 
Activities  of  Science,  / 
Addition  polymerization,  346 
Additivity  of  reaction  heats, 

law  of,  111 
Adipic  acid,  347 
Adsorption,  138 
Age  of  earth,  442 
Air 

as  source  of  elements,  442 

composition,  438 

oxygen  content,  228 
Alanine,  347 
Alcohols,  330 

naming.  338 

oxidation  of,  336 
Aldehydes,  332 
Alkali  metals,  93 


atom  models,  98 

atomic  volume,  95 

chemistry,  95 

ionic  bonds,  95 

properties,  table,  94 

reactions  with  chlorine,  95 

reaction  with  water,  96 
Alkaline  earth  elements 

atomic  radii,  378 

chemical  properties,  381 

electron  configuration,  378 

ionization  energies,  379 

occurrence,  384 

preparation,  384 

properties  of,  377,  381 

solubilities,  382 
Alkaline  earth  hydroxides 

heat  of  reaction  to  form,  382 

K.p,  383 
Alkaline  earth  oxides,  heat  of  reac- 
tion with  water,  382 
Alkaline  earth  sulfates,  K,p,  382 
Alkanes,  341 

naming,  338 
Alkyl  group,  336 
Alloys,  309 

Alnico,  406 

copper,  71,  309 

covalent  bonds,  and,  305 

gold,  71 

hardness  and  strength,  311 

nickel,  407 
Alnico,  406 

Alpha  carbon  atoms,  348 
Alpha  decay,  417,  443 
Alpha  particle,  417 

scattering,  245 
Aluminum 

boiling  point,  365 

compounds,  102 

heat  of  vaporization,  365 

hydration  energy,  368 

hydroxide,  371 

ionization  energies,  269,  374 

metallic  solid,  365 

occurrence,  373 

properties,  101 

preparation,  238,  373 

reducing  agent,  367 
Alums,  403 
Americium 

electron  configuration,  415 

oxidation  numbers,  414 
Amides,  338 

naming,  339 
Amines,  336 

naming,  338 
Amino  acids,  347,  348,  432 


Ammonia 

a  base,  184 

boiling  point,  64 

complexes,  392,  395,  408 

complex  with  Ag+,  154 

Haber  process  for,  150 

and  hydrogen  chloride,  24 

model  of,  21 

molar  volume,  60,  64 

production,  150 

P-V  behavior  of,  19,  51,60 

solubility,  20 
Ampere,  241 
Amphoteric,  371 

complexes,  396 
Analogy 

billiard  ball,  6,  18 

car  collision,  126,  129 

dartboard,  261 

garbage  collector,  233 

golf-ball,  155 

lost  child,  3 

mountain  pass,  132 

notched  beam,  256 
.  station-wagon,  155 
Angstrom,  A.  J.,  258 
Aniline,  344 
Anions,  170,  207 
Anode,  207 
Anode  sludge,  408 
Anthracite,  321 
"Antifreeze,"  72 
Antimony,  31 
Aq  notation,  79 
Aqueous,  79 
Aqueous  solutions 

of  acids,  179 

of  bases,  179 

electrical  conductivity  of,  78 

of  electrolytes,  78,  313 

precipitation  reactions  in,  80 
Argon,  91 

boiling  point,  374 

heat  of  vaporization,  105,  374 

ionization  energy,  268 

melting  point,  307 

occurrence,  373 

preparation,  374 

properties,  101 

use   105 
Aromatic  compounds,  344 
Arrhenius,  Svante,  198 
Arsenic,  ionization  energy,  410 
Asbestos,  310 
Aspirin,  346 
Astatine 

oxidation  number,  414 

properties,  97 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


455 


456 


INDEX 


Astronaut,  231 
Atmosphere,  437 

composition,  437 

one,  54 

properties,  437 
Atmospheric  pressure,  53 
Atom 

chemical   evidence   for   existence, 
234 

defined,  21 

mass  and  charge  of  parts,  87 

nuclear  model,  86 

number  per  molecule,  26 

sizes,  88 

structure,  86 
Atomic  hydrogen  spectrum,  253 
Atomic  number.  88 

and  periodic  table,  89 

table,  inside  back  cover 
Atomic  orbitals,  262,  263 
Atomic  pile,  120 
Atomic  theory,  17,  22,  28,  234 

as  a  model,  17 

chemical  evidence  for,  234 

of  John  Dalton,  236 

review.  34 
Atomic  velocity  distribution,  130, 131 
Atomic  volume,  94,  98 

alkali  metals,  94 

halogens.  97 

inert  gases.  91 

third-row  elements,  101 
Atomic  weight.  33 

table,  inside  back  cover 
Atoms,  21 

conservation  of,  40 

electrical  nature  of.  236 

measuring  dimensions  of,  245 

AVOGADRO,    AMADEO 

hypothesis,  25,  52 
hypothesis  and  kinetic  theory,  58 
law.  25 
number,  33 
Azo  dyes,  344 


Bacteria,  434 
Balancing  reactions,  42 

by  half-reactions,  218 

by  oxidation  number,  219 

oxidation-reduction  reactions,  217, 
219 
Balmer,  J.  J.,  258 
Balmer  series,  258 
Barite,  385 
Barium 

atomic  size,  379 

chemistry,  382 

electron  configuration,  378 

heat  of  vaporization,  305 

hydroxide,  Ksp,  383 

ionization  energies,  379 

occurrence,  385 

properties,  381 
Barometer,  53 

standard,  54 
Barrier,  potential  energy,  134 
Bases,  185 

aqueous.  179 

common,  185 

experimental  introduction,  183 

properties  of,  183,  184 
Basic  oxides,  382 
Battery,  storage 

Edison.  406 

lead,  406 
Bauxite.  373 


Benzene 

derivatives,  343 

modification  of  functional  group, 
344 

representation  of.  343 

substitution  reactions,  343 
Benzoic  acid,  192 
Berkelium,  oxidation   numbers  for, 

414 
Beryl,  385 
Beryllium 

atomic  size,  379 

boiling  point,  374 

bonding  capacity,  285 

chemistry  of,  382 

electron  configuration,  378 

heat  of  vaporization,  374 

ionization  energies,  379 

occurrence,  384 

preparation,  385 

properties,  381 

structure,  381 
Beryllium  difluoride,  dipole  in,  293 
Berzelius,  Jons,  30 
Bessemer  converter,  404 
Beta  decay,  417 
Beta  particle,  417 
Bicarbonate  ion,  184 
Bidentate,  395 
Billiard  ball  analogy,  6,  18 

and  kinetic  energy,  114 
Billiard   ball  collision,  conservation 

of  energy  in,  1 14 
Binding  energy,  121,  418 
Biochemistry,  421 
Bismuth,  oxidation  numbers,  414 
Blast  furnace,  404 
Bohr,  Niels,  259 
Boiling  point,  67 

elevation,  325 

normal,  68 

pure  substances,  table,  67 
Bond 

breaking  and  rate,  125 

chemical,  31,  35,  274 

covalent,  see  Covalent  bond 

covalent  and  ionic  contrasted,  287 

covalent  and  network  solids,  302, 
309 

double,  295 

energies  and  electric  dipoles,  290 

energy,  119 

energy  of  halogens,  355 

hydrogen,  see  Hydrogen  bond 

infrared  identification,  250 

ionic,  see  Ionic  bond 

length,  276,  295 

origin  of  stability,  215 

overlap,  277 

reason  for  forming,  274 

stability,  275 

stability  and   acid   strength,    187, 
188 

type,  286 

types  in  fluorine  compounds,  289 
Bonding 

capacity  and  molecular  shape,  293 

dsp\  395 

</V\  395 

electron  dot  representations,  278 

in  complex  ions,  395 

in  fluorine,  278 

in  gaseous  lithium  fluoride,  287 

in  oxygen  molecule,  295 

in  solids  and  liquids,  300 

metallic,  303 

orbital  representation,  278 


p>,  291 

P3,  291 

representations,  278 

sp.  292 

sp\  292 

sp3,  292 

tetrahedral,  292 
Bonding  capacity  of 

beryllium,  285 

boron,  285 

carbon,  284 

lithium,  286 

nitrogen,  283 

second-row  elements,  281 
Bordeaux  mixture,  408 
Boron 

boiling  point,  374 

bonding  capacity,  285 

heat  of  vaporization,  374 

ionization  energies,  273 
Boron  trifluoride,  dipole  in,  294 
Boyle's  Law,  17 

Brackets,  concentration  notation,  151 
Brass,  311 
Bromate  ion,  360 
Bromine 

atomic  volume,  410 

boiling  point,  307 

color,  352 

covalent  radius,  354 

electron  configuration,  353 

electron  dot  representation,  353 

ionic  radius,  355 

ionization  energy,  410 

melting  point,  307 

orbital  representation,  353 

preparation.  356,  361 

properties,  97,  355 

van  der  Waals  radius,  354 
Bronsted-Lowry  theory,  194 
Burning 

candle,  449 

methane,  41 

paraffin,  44 
Butane,  338 

properties,  341 
/.ro-butane  properties,  341 
Butanoic  acid,  properties,  308 
1-Butanol,  338 
Butyl  alcohol,  338 
1-Butylamine,  338 
Butyramide,  339 
Butyric  acid,  339 


Calcium 

atomic  size,  379 
atomic  volume,  410 
chemistry,  382 

electron  configuration,  271,  378 
heat  of  vaporization,  305 
ionization  energies,  379 
occurrence,  385 
properties.  381 

Calcium  carbonate,  384 

Calcium  hydroxide,  62,  383 

Calculations 

based  on  chemical  equations,  44 
of  H+  concentration,  192 
gas  volume-gas  volume,  227 
liquid  volume-volume,  230 
weight-gas  volume.  226 
weight-liquid  volume,  228 
weight-weight,  226 

Californium 
electron  configuration,  415 
oxidation  number,  414 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


INDEX 


457 


Calorie,  10 

Calorimeter.  /// 

Calorimetry.  Ill 

Candle,  description  of  burning,  449 

Caprslamide.  339 

Capr\lic  acid,  339 

Carbohydrates,  422 

Carbon 

alpha  (a),  348 

boiling  point,  374 

bonding  in.  284 

compounds,  sources  of,  321 

compounds,  composition  and 
structure.  323 

co\alent  radius,  365 

electron  configuration.  265 

heat  of  vaporization,  374 

ionization  energy,  268 
Carbon  dioxide,  28 

molar  volume.  51.  60 
Carbon  disulfide,  infrared  spectrum, 

249 
Carbon  monoxide 

absorption  by.  251 

boiling  point.  64 

molar  volume.  51.  60,  64 

poisoning,  398 
Carbon  tetrabromide 

boiling  point.  307 

bond  length.  354 

melting  point,  307 
Carbon  tetrachloride 

boiling  point,  307 

bond  length,  354 

infrared  spectrum  249 

melting  point,  307 

poison.  82 

solubility  of  ethane  in,  313 

use.  82 
Carbon  tetrafluoride 

boiling  point,  307 

bond  length.  354 

dipole  in.  294 

melting  point,  307 
Carbon  tetraiodide 

boiling  point,  307 

bond  length.  354 

melting  point,  307 
Carbonyl  group.  335 
Carboxyl  group.  334,  337 
Carboxylic  acids,  333,  334,  425 
Car  collision  analogy,  126,  129 
Cast  iron,  404 
Catalysis,  135 
Catalysts 

action  of,  135 

enzymes,  138 

and  equilibrium,  148 

examples  of,  137 

in  formic  acid  decomposition,  137 

in  manufacture  of  H2S04,  227 

and  rusting,  405 
Cathode,  207 
Cathodic  protection,  405 
Cations,  170,  207 
Cell 

dry,  403 

electrochemical,  199,  206 
Cellulose,  425 

structure,  431 
Centigrade  temperature,  57,  58 

relation  to  °K,  58 
Cerite,  413 
Cerium 

properties,  412 

source,  413 
Certainty,  absence  of,  // 


Cesium,  23 

chemistry.  95 

heat  of  vaporization,  305 

properties.  94 
Chain  reaction,  419 
Chalcocite,  46 
Chalcopyrite,  408 
Charge 

conservation  of,  80 

electric,  75 

electron,  241 

positive,  242,  243 

separation,  312 
Charge  mass  ratio  of  electrons, 

240 
Charles,  Jacques,  57 
Charles'  law,  58 
Chemical  bonding,  see  Bonding 
Chemical  bonds,  see  Bond 
Chemical  change,  38 
Chemical  energy,  1 19 
Chemical  equations,  see  Equations 
Chemical   equilibrium,   law  of.    152 
Chemical  formulas,  see  Formula 
Chemical  kinetics.  124 
Chemical  reactions,  see  Reactions 
Chemical  stability.  30 
Chemical  symbols.  30 

not  from  common  names,  31 

see  inside  back  cover 
Chemotherapy.  434 
Chlorate  ion,  360 
Chloric  acid,  359 
Chlorides 

chemistry  of,  99 

of  alkali  metals.  93,  103 

of  third-row  elements,  103 
Chlorine 

boiling  point,  374 

color,  352 

compounds,  102 

covalent  radius,  354 

electron  configuration,  353 

heat  of  fusion,  69 

heat  of  reaction  to  form  atoms,  290 

heat  of  vaporization,  347 

ionic  radius,  355 

ionization  energy,  268 

melting  point,  307 

molar  volume,  60,  64 

occurrence,  373 

oxidation  numbers,  359 

oxidizing  agent,  369 

oxyacids,  359,  373 

preparation,  356,  374 

properties,  97,  101,  355 

structure,  366 

use,  106 

van  der  Waals  radius,  354 
Chloroacetic  acid,  349 
Chlorobenzene,  345 
Chloromethane 

solubility  in  CC14,  313 

solubility  in  acetone,  313 
Chlorophyll,  431 

structure,  397 
Chlorous  acid,  359 
Chromate  ion,  structure,  402 
Chromatography,  413 
Chromium 

electron  configuration,  389 

oxidation  number,  391 

properties,  400,  401 

radius,  399 
Chromium(III)  oxide,  402 
Chromium  trihydroxide,  396,  402 
Cis-trans  isomerism,  296,  394 


Citric  acid.  428 
Cleaning  solution,  403 
Coal.  321 
Coal  gas,  322 
Coal  tar,  322 
Cobalt 

atomic  radius,  399 

electron  configuration,  389 

ore,  410 

oxidation  numbers,  391 

properties,  400,  406 
Coke,  322,  404 
Collision  theory,  126 

concentration  effect,  126 

and  organic  mechanism,  331 

temperature  effect,  129 
Color,  399 

Colorimetric  analysis,  151 
Communicating    scientific    informa- 
tion, 12 
Combining  volumes,  law  of,  26,  236 
Combustion,  heat  of  hydrogen,  40 
Complex  ions,  392 

amphoteric,  396 

bonding  in,  395 

formation.  413 

geometry  of,  393 

in  nature,  396 

isomers,  394 

linear,  395 

octahedral,  393 

significance  of,  395 

square  planar,  395 

tetrahedral,  394 

weak  acids,  396 
Compound,  28 

bonding  in,  306 
Concentration 

and  equilibrium,  148 

and  E  zero's,  213 

and  Le  Chatelier's  Principle,  149 

effect  on  reaction  rate,  126,  128 

molar,  72 

of  solids  in  equilibria  expressions, 
153 

of  water  in  equilibria  expressions, 
154 

pH.  190 
Conceptual  definition,  195 
Condensed  phases,  27,  68,  78 

electrical  properties,  78 
Conductivity,  electrical 

in  metals,  81 

in  water  solutions,  78 

of  solids,  80 
Condensation  polymerization,  346 
Conservation 

of  atoms,  40 

of  charge,  80,  218 

of  mass,  40 
Conservation  of  energy 

in  a  billiard  ball  collision,  114 

in  a  chemical  reaction,  115 

in  a  stretched  rubber  band,  114 

law  of,  113,  117,  207 
Constant  heat  summation  law,  1 1 1 
Contact  process,  H5SO4,  227 
Coordination  number,  393 
Copper 

alloys,  conductivity  of,  309,  311 

atomic  radius,  399 

electron  configuration,  389 

heat  of  fusion,  69 

heat  of  melting,  69 

native,  408 

occurrence,  408 

oxidation  numbers,  391 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


458 


I  N  D  L;  X 


Copper  (continued) 

properties,  400.  408 
Core,  of  earth,  440 
Corrosion,  405 
Coulomb,  241 
Coulombic  forces,  416 
Coulson,  C.  A.,  252 
Covalent  bonds,  274,  277,  288 

elements  that  form  solids  using, 
302 

in  halogens,  97 

vs.  ionic,  287 

and  network  solids,  302,  309 
Covalent  radius,  354 

of  halogens,  354 
Crawford,  Bryce,  Jr.,  274 
Cryolite,  393 
Crystal 

ionic,  81,  311 

metal,  304 

molecular,  81,  302 
Crystallization,  70,  71 

and  equilibrium,  144 

fractional,  413 
Curium,  oxidation  number,  414 
Cyclohexane,  properties,  341 
Cyclopentane,  340 
Cyclopentene,  342 

d  orbitals,  262 

dsp1  bonding,  395 

d2sp*  bonding,  395 

"Dacron,"  347 

Dalton,  John,  12,  38,  236 

atomic  theory,  236 
Dartboard  analogy,  261 
Da  Vinci,  Leonardo,  1 
Debye,  Peter  J.,  320 
Decomposition 

of  formic  acid,  137 

of  water,  40,  115 
Definite  composition  of  compounds, 

Law  of,  234 
Definitions,  see  under  word  of  in- 
terest 

conceptual,  195 

experiment,  2 

operational,  195 
Delta  H  (AH),  110 
Delta  H  cross  (AHt),  135 
Density,  28 
Derived  quantity,  10 
Deuterium,  90,  123,  419 
Deuterium  bromide,  infrared  spec- 
trum, 248 
Diamond,  structure,  302,  365 
Diatomic,  31 
Diborane,  285 
Dichloroethylene,  isomers  of,  297 

properties,  308 
Dichromate  ion,  402 
Diffraction 

of  X-rays,  248 

patterns,  248 
Dipoles,  288 

BF3,  294 

BeF2,  293 

and  bond  energy,  290 

CF4,  294 

F20,  294 

geometrical  sum,  293 

LiF,  293 

molecular,  293 

and  molecular  shape,  293 

of  ionic  bond,  288 

and  solvent  properties,  313 


Disaccharides,  424 
Discharge  tube,  239,  255 
Disproportionation,  361 
Dissociation,  179 

constant  of  HF,  361 
Dissolving  rate,  164 
Distillation,  70 
Disulfur  dichloride,  103 
Divalent,  282 
Divanadium  pentoxide,  401 

DbBEREINER,   J.   W.,    104 

Dolomite,  385 

Double  arrows,  use  of,  146 

Double  bonds,  295 

Dry  cells,  199,  403 

Dyes,  azo,  344 

Dynamic  nature  of  equilibrium,  144, 

165 
Dysprosium,  properties,  412 

E  =  mc\  121,  418 
E°,  209 

and  equilibrium,  215 

table,  452 
Earth 

age  of,  442 

chemistry  of,  437 

core  of,  440 

crust  of,  441 

data,  444 

parts  of,  438 
Edison  storage  battery,  406 
Einstein,  Albert,  121 
Einsteinium,  oxidation  number,  414 
Elastic  collision,  6 
Electrical  nature  of  atoms,  236 
Electrical  phenomena,  74 
Electrical  properties  of  condensed 

phases,  78 
Electric  arc  furnace,  404 
Electric  charge,  75 

detection,  74 

effect  of  distance,  76 

in  matter,  77 

interactions,  75 

negative,  77 

positive,  77 

production,  76 

types,  76 
Electric   conductivity,   see  Conduc- 
tivity 
Electric  current,  78 
Electric  dipoles,  see  Dipoles 
Electric  discharge,  239 
Electric  force,  76,  77 
Electricity,  fundamental  unit,  241 
Electrochemical  cell 

chemistry  of,  199 

and  Le  Chatelier's  Principle.  214 

operation,  206 

standard  half  cell,  21C 
Electrodes,  207 
Electrolysis,  220,  221 

apparatus,  40 

cells,  238 

of  water,  40,  115 
Electrolytes,  169,  179 

strong,  180 

weak,  180 
Electrolytic  conduction,  220,  see  also 

Conductivity,  electrical 
Electrometer,  75 
Electron,  77 

affinity,  280 

affinity  of  fluorine  atom,  280 

behavior  in  metals,  304 


capture,  417 

charge,  241 

charge/mass  ratio,  240 

competition  for,  205 

deflection,  239 

evidence  for  existence,  236 

losing  tendency,  207 

mass  of,  87,  241 

"sea"  of,  304 

"seeing,"  239 

transfer,  201,  202 

valence,  269 

van  der  Waals  forces  and,  306 
Electron  configuration,  265 

of  alkaline  earths,  378 

of  halogens,  352 

of  lanthanides,  415 

of  transition  elements,  389 
Electron  dot  representation  of  chemi- 
cal bonding,  278 
Electron  populations  of  inert  gases, 

92,  263 
Elements,  29 

abundance  in  earth's  crust,  441 

availability,  441 

covalent  solids  formed  by,  302 

discovery,  30 

in  ocean,  440 

known  to  ancients,  30,  441 

metals,  303 

molecular  crystals  formed  by,  301 

naming,  30,  31 

on  Sun,  447 

symbols,  see  inside  back  cover 
e/m,  240,  243 
Emerald,  385 

Empirical  formula,  81,  323,  324 
Endothermic  reaction,  40,  135 
Energy 

activation,  132,  134,  369 

binding,  121,418 

changes  on  warming,  119 

chemical,  119 

effect  on  equilibrium,  167 

kinetic,  see  Kinetic  energy 

law  of  conservation  of,  113,  117, 
207 

molecular,  118 

nuclear,  119,  418 

of  hydrogen  bond,  315 

of  light,  253 

of  motion,  see  Kinetic  energy 

of  position,  see  Potential  energy 

potential,  see  Potential  energy 

stored  in  living  organisms,  428 

stored  in  a  molecule,  117 

stored  in  the  nucleus,  110 
Energy  levels,  259 

notched  beam  analogy,  256 

of  H  atom,  258,  259,  264 

of  many-electron  atoms,  265,  266 
Enzymes,  138,  427,452 
Equations 

balancing,  42 

calculations  based  upon,  44,  226 

chemical,  41 

symbols  used  in,  41,  146 

writing,  42 
Equilibrium,  67,  144,  145,  158 

altering,  147 

and  chemical  reactions,  145 

and  E°,  215 

and  Le  Chatelier's  Principle,  149 

and  rate,  155 

and  solubility,  144,  163 

attainment  of,  148 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


INDEX 


459 


Equilibrium  (continued) 

calculations,  192 

constant,  151,  table,  154 

crystallization  and,  144 

dynamic  nature  of,  144,  165 

effect  of  catalyst,  148 

effect  of  concentration,  148;  of  en- 
ergy, 167;  of  randomness,  166; 
of  temperature,  67.  148,  167 

factors  determining,  155,  158 

law  of  chemical,  152,  173 

liquid-gas,  66 

qualitative  aspects  of,  142 

quantitative  aspects  of,  151 

recognizing,  143 

state  of,  142,  147 

sugars,  423 

thermal,  56 

use  of  double  arrows,  146 
Equilibrium  constant,  151 

for  water,  181 

table,  154 

table  (acids),  191,  451 
Equilibrium  Law  relation,  table,  154 
Erbium,  properties,  412 
Esters,  337 
Ethane,  323 

properties,  341 

solubility  in  CC14;  in  acetone,  313 
Ethanol,  323 

boiling  point,  329 

determining    structural    formula, 
326 
Ethylamine,  338 
N-Ethyl  acetamide,  338 
Ethyl  bromide,  328 

reactions  of,  330 
Ethyl  iodide,  336 
Ethylene,  346 

chemical  reactivity,  296 

double  bond  in,  296 
Ethylene  glycol,  325 
Ethyl  group,  329 
Europium,  properties,  412 
Exothermic  reaction,  40,  135 
Experiment,  2 

Experimental  errors,  see  Uncertainty 
Eyring,  Henry,  124,  141 
E  zero,  209,  table,  452 

/orbitals,  262 

Fable,  Lost  Child  in  Woods,  3 

Faraday,  Michael,  237 

Fats,  425 

Fermentation,  426 

Fermium,  oxidation  number,  414 

Ferromanganese,  403 

Fission,  nuclear,  120,  419 

Flea,  88 

Fluorescence,  239,  409 

Fluorides 

bonding  capacity,  293 

bond  types  in  second  row,  286 

of  second-row  elements,  melting 
and  boiling  points,  286 

orbitals,  293 

shape,  293 
Fluorine 

boiling  point,  374 

bonding,  278 

color,  352 

complex,  393 

covalent  radius,  354 

electron  affinity,  280 

electron  configuration,  265 

electron  dot  representation,  353 


heat  of  reaction  to  form  atoms,  290 

heat  of  vaporization,  374 

ionic  character  of  bonds,  288 

ionic  radius,  355 

ionization  energy,  268 

melting  point,  307 

orbital  representation,  353 

preparation,  356 

properties,  355 

special  remarks,  361 

van  der  Waals  radius,  354 
Fluorine  compounds,  bond  type,  289 
Fluorine  oxide 

dipole  in,  294 

molecular  shape,  291 
Fluorocarbons,  362 
Fool's  gold,  404 
Force,  electric,  76,  77 
Formaldehyde,  332 
Formamide,  339 
Formation,  heat  of,  113 
Formic  acid,  197,  333 

catalytic  decomposition,  137 
Formula 

chemical,  30 

empirical,  81,  323,  324 

molecular,  31,  323,  325 

structural,  31,  323,  326 
Fourth  row  of  periodic  table,  271, 

387 
Fractional  crystallization,  413 
Freezing  point  lowering,  325,  393 
"Freon,"  362 
Frequency  of  light,  246 

relation  to  wave  length,  251 
Fructose,  423 
Fumaric  acid,  428 

properties,  308 

structure,  316 
Functional  groups,  330,  335 
Fundamental 

property,  78 

unit  of  electricity,  241 
Furnace,  electric  arc,  404 
Fusion,  heat  of,  68 

pure  substances,  table,  69 
Fusion,  nuclear,  121,  419 


Gadolinite,  413 
Gadolinium,  properties,  412 
Gallium 

atomic  radius,  399 

ionization  energy,  410 
Galena,  373 

Galvanized  iron,  405,  409 
Garbage  collector  analogy,  233 
Gas 

elements  found  as,  65 

ideal,  60 

inert,  90 

and  kinetic  theory,  49,  53 

liquid-gas  equilibrium,  66 

liquid-gas  phase  change,  66 

measuring  pressure,  53 

model  of,  18,  23 

molar  volume,  49,  50,  51,  60 

molecules  in,  274 

molecular  weight,  34,  51 

natural.  322 

perfect,  59 

pressure,  53 

pressure,  cause  of,  54 

properties  of,  20 

review,  61 

solubility,  20,  167 


volume,  calculation  of,  227 

volume,  change  with  temperature, 
57 

volume  measurement,  24 
Gasoline,  46,  63,  126,  341 
Gastric  juice,  1 38 
Gaviota  Pass,  136 
Gay-Lussac,  Joseph  L.,  49 
Geiger  counter,  240 
Generalization,  4,  153 

melting  of  solids,  4 

reliability  of,  59 
Geometry 

molecular,  290-297 

of  complex  ions,  393 
Germanium,  ionization  energy,  410 
Glucose,  422 
Glutamic  acid,  347 
Glyceraldehyde,  427 
Glyceric  acid,  427 
Glycerol,  425 
Glycine,  347 
Gold 

alloy,  71 

oxidation  numbers,  414 
Goldschmidt  reaction,  401 
Golf  ball  analogy,  155 
Graphite 

burning,  46 

structure,  302 
Gypsum,  385 


h  (Planck's  constant),  254 

Hand  AH,  110 

[H+],  calculation  of,   192,  see  also 

Hydrogen  ion 
Haber,  Fritz,  151 
Haber  process,  140,  150 
Hafnium,  oxidation  number,  414 
Haldane,  J.  B.  S.,  436 
Half-cell  potentials 

effect  of  concentration,  213 

measuring,  210 

standard,  210 

table  of,  211,452 
Half-cell  reactions,  201 
Half-life,  416 
Half-reaction,  201 

balancing,  218 

potentials,  452 
Halides 

chemistry  of,  99 

ions,  98 
Hall,  C.  M.,  96.  373 
Halogens 

atom  models,  98 

bond  energies,  355 

chemistry,  98 

color,  352 

covalent  bonds,  97 

covalent  radius,  355 

electron  configuration,  352 

ionization  energies,  353 

oxyacids,  358 

positive  oxidation  states,  358 

preparation.  356 

properties,  96;  table,  97.  352.  355 

reactions,  energy  required  for.  357 

reactions  of  compounds,  356 

reduction  of,  357 

sizes  of  atoms  and  ions,  354 

toxicity.  352 
Hardness  of  metals  and  alloys,  31 1 
Hard  water,  384 
Heat  and  chemical  reactions,  108 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


460 


INDEX 


Heat  content,  110,  116 

change  during  a  reaction,  110 

of  a  substance,  109 
Heat  of  combustion  of 

diamond,  122 

graphite,  122 

hydrazine,  47 

hydrogen,  40 

methane,  123 
Heat  of  formation.  113 
Heat  of  reaction,  135 

between  elements,  table,  112 

oxidation  of  HC1,  160 

oxidation  of  sulfur  dioxide,  161 

predicting,  112 
Heat  of  reaction  to  form 

ammonia,  1 12 

Br  atoms,  290 

carbon  dioxide,  112 

carbon  monoxide,  112 

CI  atoms,  290 

CO  +  H2,  110 

ethane,  112 

F  atoms,  290 

H  atoms,  274 

hydrogen  chloride,  160 

hydrogen  iodide,  112 

iron(lll)  oxide,  162 

Li  atoms,  290 

Li  +  Br,  290 

Li  +  F,  290 

Na  +  CI,  290 

NH3  products,  114 

Na  atoms,  290 

NO,  112 

N02,  112 

N*3  products,  114 

P4Oio,  140 

propane,  112 

sulfur  dioxide,  112 

sulfuric  acid,  112 

water,  109,  112 
Heat  of  solution,  166 

Cl2  in  H20.  167 

h  in  benzene,  167 

I2  in  ethyl  alcohol,  166 

I2  in  CCU,  166 

02  in  HoO,  167 

N20  in  H20,  167 
Heat  of  vaporization,  66;  see  also 

Vaporization 
Helium,  91 

boiling  point,  63 

heat  of  vaporization,  105 

interaction  between  atoms,  277 

ionization  energy,  268 

molar  volume,  60 

on  Sun,  447 

source,  91 
Hematite,  404 
Hemin,  structure  of,  397 
Hess's  Law,  111 
Heterogeneous,  70 

systems  and  reaction  rate,  126 
n-Hexane  properties,  341 
Hibernation,  2 

HlLDEBRAND,    JOEL    H.,    163 

Holmium,  properties,  412 
Homogeneous,  70 

systems  and  reaction  rate,  126 
Hydration,  313 
Hydrazine.  46,  47,  231 
Hydrides  of  third-row  elements,  102 

boiling  point  of,  315 
Hydrocarbons,  340 

unsaturated,  342 


Hydrochloric  acid,  42 
Hydrofluoric  acid,  storage  of,  361 
Hydrogen,  99 

apparent  ionization  energy,  289 

boiling  point,  63 

bonding  in,  274 

burning,  25,  41 

chemistry,  100 

energy  level  scheme,  258,  264 

freezing  point,  63 

heat  of  combustion,  40 

helium  nuclear  reaction.  419 

ion,  and  weak  acids,  193 

ion  (aqueous),  187 

ion  as  catalyst,  137 

ion,  competition  for  in  weak  acid, 
194 

ion  concentration,  calculation  of, 
192 

ion,  hydrated  form,  187 

ion,  nature  of,  185 

ionic  character  of  bonds,  289 

ionization  energy,  268 

isotopes,  90,  123,  419 

molar  heat  of  combustion,  40 

molar  volume,  60 

production,  62 

properties,  table,  100 

solubility,  20 

spectrum  of,  253,  255 
Hydrogenation,  407 
Hydrogen  atom,  253 

energy  level  scheme,  258,  264 

energy  levels  of,  259 

heat  of  reaction  to  form,  274 

light  emitted  by,  254 

orbitals,  263 

periodic  table  and,  263 

quantum  mechanics  and,  259 

quantum  numbers  and,  260 

spectrum  of,  253,  255 
Hydrogen  bonds,  314,  329,  361 

energy  of,  315 

inter-  and  intramolecular,  316 

nature  of,  316 

and  proteins,  432 

representation  of,  315 

significance  of,  316 

sugars,  424 
Hydrogen  bromide 

boiling  point,  315 

infrared  spectrum,  248 
Hydrogen  chloride 

ammonia  mixture,  24 

boiling  point,  64 

heat  of  oxidation,  160 

heat  of  reaction  to  form,  160 

model  of,  21 

molar  volume,  60,  64 

P-V  relation,  19 
Hydrogen  fluoride 

boiling  point,  315 

bonding,  280 

melting  point,  99 
Hydrogen  halides,  315 
Hydrogen  iodide,  boiling  point,  315 
Hydrogen  peroxide  bonding,  283,  295 
Hydrogen  selenide,  boiling  point,  315 
Hydrogen  sulfide,  boiling  point,  315 
Hydrogen    telluride,    boiling    point, 

315 
Hydrolysis 

of  fats,  426 

of  starch,  426 

of  sugars,  426 
Hydrometer,  407 


Hydronium  ion,  187 

concentration  calculation,  192 

concentration  and  /?H,  190 

model,  186 
Hydroquinone,  345 
Hydrosphere,  437 

composition,  439 
Hydroxide  ion,  106,  180 
Hydroxides  of  third  row,  371 
Hydroxylamine,  251 
Hydroxyl  group,  329 
Hypobromite  ion,  422 
Hypochlorite  ion,  361 
Hypochlorous  acid,  structure,  359 
Hypophosphorous  acid,  372 
Hypothesis,  Avogadro's,  25,  52 


Ice 

bonds  in,  315 

melting  of.  69 

structure,  69,  315 
Ideal  gas,  60 

see  also  Perfect  gas 
Ilmenite,  401 
Indicators 

acid-base,  190 

litmus,  21,  189 
Inductive  reasoning,  3 
Inert  gases.  90 

atom  models,  98 

compounds,  91 

electron  population,  92,  263 

properties,  table,  91 

stability  of  structure,  93 
Information 

accumulating,  75 

organizing,  75 
Infrared,  247 
Infrared  spectra,  248 

of  HBr,  DBr,  248 

of  CCU,  CS2,  249 
Infrared  spectroscopy,  249 
Invisible  ink,  406 
Iodate  ion,  360 
Iodimetry,  358,  358 
Iodine 

boiling  point,  307 

color,  352 

covalent  radius,  354 

electron  configuration,  353 

ionic  radius,  355 

melting  point,  307 

preparation,  231,  356 

properties  of,  97,  355 

solubility  in  CC14,  166 

solubility  in  ethanol,  163 

van  der  Waals  radius,  354 
Ion,  78 

exchange  resin,  413 

formation,  86 

hydrogen  (aqueous),  185 

names,  formulas,  charges  of  com- 
mon, 454 
Ionic  bond,  287,  288 

dipole  of,  288 

in  alkali  metal  halides,  95 

vs.  covalent,  287 
Ionic  character,  287 
Ionic  crystal,  81.  J77 
Ionic  radius,  355 
Ionic  solids,  79,  57,  i77 

electrical  conductivity,  80 

properties  of,  572 

solubility  in  water,  79 

stability  of,  J77 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


INDEX 


461 


Ionization 

lithium,  267 

magnesium,  270 

sodium,  270 
Ionization  energy,  267 

alkaline  earths.  379 

and  atomic  number,  268 

and  the  periodic  table,  267 

and  valence  electrons,  269 

halogens,  353 

measurement  of,  268 

successive.  269 

table  of,  268 

trends,  268 
Iridium,  oxidation  numbers,  414 
Iron 

atomic  radius,  399 

cast,  404 

complex  with  oxalate,  395 

electron  configuration,  389 

galvanized,  405,  409 

manufacture  of,  404 

occurrence,  404 

oxidation  numbers,  391 

properties.  400.  403 

rusting  of.  45,  85.  .405 
Iron(III)  oxide,  heat  of  reaction  to 

form.  162 
Iron  pyrites,  46,  404 
Isomers 

cis-  and  trans-,  296,  394 

of  complex  ions,  394 

of  C;H60,  327 

structural,  327 
/.yo-prefix,  341 
Isotopes,  90 

and  mass  number,  89 

and  mass  spectrograph,  243 

vital  statistics,  table,  89 

Jupiter,  data  on,  444 

K 

magnitude,  154 

table,  154 
Ka 

determination,  192 

values,  191 
K.p,  174 

table,  174 
Ku,  181 

change  with  temperature,  181 

values  at  various  temperatures,  181 
Kcal,  40 

Kelvin  temperature  scale,  58 
Ketones,  334 
Kerosene,  231,  341 
Kilo,  40 
Kilocalorie,  40 
Kinetic  energy,  53,  114 

billiard  ball  analogy,  6,  114 

distribution,  130,  131 

formula  for,  59 

of  a  moving  particle,  59 

relation   to  temperature,  56,   131 
Kinetic  theory,  49,  52,  53 

and  Avogadro's  Hypothesis,  58 

review.  61 
Kinetics,  chemical,  124 
Knudsen  cell,  63 
Kroll  process,  368 
Krypton,  91 

atomic  volume,  410 

boiling  point,  307 

heat  of  vaporization,  105 

Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


ionization  energy,  410 
melting  point,  307 

Laboratory,  2 

Lanthanide  elements,  411,  389 

contraction,  413 

electron  configurations,  415 

occurrence  and  preparation,  413 

oxidation  numbers,  414 

properties,  412 
Lanthanum 

heat  of  vaporization,  305 

properties,  412 

source,  413 
Latimer,  Wendell  M.,  199 
Lattice,  81 
Laws,  14,  117 

additivity  of  reaction  heats,  111 

Avogadro's,  25,  52 

Boyle's,  17 

Charles',  58 

chemical  equilibrium,  152,  173 

chemical  equilibrium  derived,  155 

combining  volumes,  236 

conservation  of  charge,  80 

conservation  of  energy,  113,  117 

conservation  of  mass,  40 

constant  heat  summation,  111 

definite  composition,  234 

Hess's,  111 

octaves,  104 

of  nature,  117 

simple  multiple  proportions,  235 
Lead,  oxidation  numbers,  414 
Lead  chamber  process,  H2S04,  227 
Lead  storage  battery,  406 
Le  Chatelier,  Henry  L.,  149 
Le  Chatelier's   Principle,   149,    181, 
188,  337,  360 

and  concentration,  149 

and  electrochemical  cells,  214 

and  equilibrium,  149 

and  pressure,  149 

and  temperature,  150 
Levels,  energy,  259 
Lewis,  G.  N.,  48,  142 
Light,  246 

a  form  of  energy,  253 

emitted  by  hot  tungsten,  255 

emitted  by  hydrogen  atoms,  254 

spectrum,  247 
Lime,  manufacture,  143 
Limestone,  385,  404 
Linear  complex,  395 
Line  spectra,  255 
Liquid,  27,  65 

gas  equilibrium,  66 

gas  phase  change,  66 

solid  phase  change,  5,  68 

volume  calculations,  230 
Lister,  346 
Lithium 

bo  ling  point,  374 

bonding  capacity,  286 

bonding  in  gaseous,  287 

heat  of  reaction  to  form  atoms,  290 

heat  of  vaporization,  374 

ionization  energy,  268 

ionization,  267 

properties,  94 
Lithium  fluoride 

bonding  in  gaseous,  287 

dipole  in,  293 

heat  of  reaction  to  form  atoms,  290 
Lithosphere,  437 

composition,  440 


properties,  440 
Litmus,  21,  189 

color  in  acids,  183 

color  in  bases,  184 
Los  Angeles  to  San  Francisco  anal- 
ogy, 132 
Lost  Child  fable,  3 
"Lucite,"  347 
Lutetium,  properties,  412 
Lyman,  T.,  258 


M,  (symbol  for  molar  concentration), 

72 
Macroscopic 

changes  and  equilibrium,  143,  147 

properties,  118 
Magnesium 

atomic  size,  379 

boiling  point,  365 

chemistry,  382 

complex  formation,  396 

compounds,  102 

electron  configuration,  378 

heat  of  vaporization,  365 

hydration  energy,  368 

ionization  energies,  269,  374 

ionization,  270 

metallic  solid,  365 

occurrence,  373,  385 

oxide,  heat  of  reaction  to  form,  375 

preparation,  373 

properties,  101,  381 

reducing  agent,  367 
Magnesium  hydroxide,  370 

K,p,  383 
Magnesium  sulfate,  Kap,  383 
Magnetic  field,  240 
Magnetite,  404 
Maleic  a  id 

properties,  308 

structure,  316 
Manganese 

atomic  radius,  399 

electron  configuration,  389 

oxidation  numbers,  391 

properties,  400,  403 
Manganese  dioxide,  403 
Manometer,  53 
Mantle,  440 
Manufacture  of 

aluminum,  373 

ammonia,  140,  150 

helium,  91 

hydrogen,  62 

iodine,  231 

lime,  143 

magnesium,  373 

methanol,  160 

nitric  acid,  45,  232 

nitrogen  compounds,  35,  150 

phosphorus,  376 

sodium,  238 

sodium  carbonate,  230 

sulfuric  acid,  225,  369 

titanium,  368 

uranium,  35 
Many-electron  atoms 

energy  level  diagram,  266 

energy  levels  of,  265 
Mars,  data  on,  444 
Mass 

conservation  of,  40 

of  atom  and  its  parts,  87 

of  electron,  87,  241 
Mass  defects,  121 


462 


INDEX 


Mass-energy  relationship,  121 
Mass  number,  90,  120 
Mass  spectrograph,  242,  443 
Mass  spectrum  of  neon,  242 
Matter 

electrical  nature,  74, 11 

fundamental  property,  78 
McMillan,  Edwin  M.,  420 
Measurement  of  gas  pressure.  53 
Mechanism  of  reaction,  127,  128 

CH3Br  +  OH-,  331 

decomposition  of  formic  acid,  138, 
139 

oxidation  of  HBr,  128 
Melting 

of  ice,  69 

of  solids,  generalization,  4 
Melting,  heat  of,  see  Fusion,  heat  of 
Melting  point 

alkali  metals,  94 

alkaline  earths,  381 

halogens,  355  • 

inert  gases,  91 

paradichlorobenzene,  9 

sodium  chloride,  69 

third-row  elements,  101 

transition  elements,  400 

water,  69 
Mendeleev,  Dimitri,  104,  107 
Mendelevium,  oxidation  number,  414 
Mercuric  perchlorate,  237 
Mercurous  perchlorate,  237 
Mercury,  oxidation  numbers,  414 
Mercury  (planet),  data  on,  444 
Metabolism,  oxidative,  429 
Metallic 

alloys,  309 

bond, 303 

elements,  303 

radius,  380 

substances,  81 
Metals 

alkali,  94 

characteristic  properties,  81,  303 

conductivity,  81 

electron  behavior  in,  304 

hardness  and  strength,  311 

heats  of  vaporization,  305 

location  in  periodic  table,  304 

properties  explained,  305 
Meteorites,  404,  444 
Methane 

burning,  41 

molar  volume,  60 

properties,  341 

solubility  in  CCl*;  in  acetone,  313 

structural  formula,  332 
Methanol,  structural  formula,  332 
Methyl 

acetate,  338 

alcohol,  338 

amine,  338 

ammonium  ion,  348 

bromide,  330 

butyrate,  339 

caprylate,  339 

ether,  boiling  point,  313 

formate,  339 

group,  330 

octanoate.  339 

orange,  344 

propionate.  339 

salicylate,  346 
yV-Methyl  acetamide,  348 
Meyer,  Lothar,  104 
Mica,  310 


Microscopic  processes,  and  equilib- 
rium, 147 
Microwave  spectroscopy,  249 
Millikan,  Robert,  241 

oil  drop  experiment,  241 
Minerals,  373,  385 
Miscible,  176 
Model 

atomic  theory,  34 

electron-proton,  76 

gases,  23 

kinetic  theory  of  gases,  53 

nuclear  atom,  86 

particle,  18 

pressure  of  gases,  18 

scientific,  18 

system,  7 
Models 

alkali  atoms,  98 

halogen  atoms,  98 

inert  gas  atoms,  98 

molecular,  various  representations, 
32 

water,  31 
Molar  concentration,  72 
Molar  heat 

of  combustion,  see  Heat  of  com- 
bustion 

of  fusion,  see  Fusion,  heat  of 

of  melting,  see  Fusion,  heat  of 

of  vaporization,  see  Vaporization, 
heat  of 
Molar  volume,  50 

alkali  metals,  94 

gases,  comparison,  50,  51,  60 

halogens,  97 

inert  gases,  91 

table  of,  60 

third-row  elements,  101 
Mole,  32,  32,  33 

in  calculations,  44,  225 

volume  of,  see  Molar  volume 
Molecular  architecture,  290 
Molecular  crystal,  81 
Molecular  formula,  31 

determination,  325 
Molecular  rotation,  249 
Molecular  shape 

BF3,  292 

BeF2,  292 

CF4,  292 

CH<,  292 

F20,  291 

H20,  291 

NF3,  291 

NH3,  291 

and  melting  point,  308 

and  van  der  Waals  forces,  307 
Molecular  size,  and  van  der  Waals 

forces,  307 
Molecular  solids,  102,  301,  306 
Molecular  substances  and  van  der 

Waals  forces,  306 
Molecular    structure,    experimental 

determination  of,  324 
Molecular  velocities,  distribution  of, 

131 
Molecular  vibrations,  249 
Molecular  weight,  33,  33 

boiling  point  correlation,  307 

calculation,  33 

determination,  325 
Molecules,  21,  274 

energy  of,  118 

measuring  dimensions,  245 

models  of,  21 


number  of  atoms  in,  26 

polar,  288 

relative  weights,  25 

representations  of,  32 

weights  of,  22 
Momentum,  59 
Monazite,  413 
Monel  metal,  407 
Monomer,  346 
Monosaccharides,  424 
Morgan,  G.  T.,  224 
Motion,  types  of,  118 
Mountain  pass  analogy,  132 
Multiple  proportions,  Law  of,  235 
Myristic  acid,  425 


n  (principal  quantum  number),  259, 

261 
Names 

of  common  ions,  454 

organic,  339 
Natural  gas,  46,  322 
Negative  charge,  77,  see  also  Electron 
Negative  ion,  87,  207 
Negligible  solubility,  73 
Neodymium 

properties,  412 

source,  413 
Neon,  91 

boiling  point,  374 

electron  configuration,  265 

heat  of  fusion,  69 

heat  of  vaporization,  105,  374 

ionization  energy,  268 

mass  spectrum  of,  242 

melting  point,  307 

sign,  239 
Neopentane,  341 

properties,  308 
Neptune,  data  on,  444 
Neptunium 

electron  configuration,  415 

oxidation  numbers,  414 
Net  reaction,  201 
Network  solid,  102 

and  covalent  bonds,  302,  309 

diamond  and  graphite,  302 

one-dimensional,  309 

three-dimensional,  309 

two-dimensional,  309 
Neutral,  77,  189 
Neutron 

properties,  87 

ratio  to  proton,  417 

relation  to  mass  number,  90 
Newlands,  J.  A.  R.,  85,  104 
Newton,  Isaac,  17 
Nichrome,  402 
Nickel 

atomic  radius,  399 

alloys  of,  407 

electron  configuration,  389 

oxidation  number,  391 

properties,  400,  406 
Nickel  carbonyl,  410 
Nitration,  344 

Nitric  acid,  manufacture,  45,  232 
Nitric  oxide 

as  catalyst,  227 

reaction  with  oxygen,  26 

solubility,  20 
Nitrobenzene,  344 
Nitrogen 

boiling  point,  63,  374 

bonding  capacity,  283 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


INDBX 


463 


Nitrogen  (continued) 

compounds,  manufacture  of,   35, 
150 

density,  50 

electron  configuration,  265 

freezing  point,  63 

heat  of  vaporization,  374 

ionization  energy,  268 

molar  volume,  49,  51,  60 

reaction  with  hydrogen,  150 
Nitrogen  dioxide,  as  catalyst,  227 
Nitrous  acid,  KA,  191 
Noble  gases,  see  Inert  gases 
Nomenclature 

common  ions,  454 

organic  compounds,  339 
Normal  boiling  point,  68 
Notched  beam  analogy  to  H  spec- 
trum, 256 
Novocaine,  345 
M-pentane,  308,  340 
nu,  246 
Nuclear 

atom  model,  86 

changes,  121 

charge,  86 

energy,  119,4/8 

fission,  120,  419 

forces,  87 

fuels,  35,  413 

fusion,  121,  419 

power  plant,  82 

reactions,  120,  443 

stability,  416 
Nucleon,  120,  416 
Nucleus 

energy  stored  in,  120 

radius,  88,  251 

"seeing,"  244 
"Nylon,"  347 


Observation,  2,  15 
Ocean,  composition  of,  440 
Ocean  water,  composition,  439 
w-Octadecane,  properties,  341 
Octahedral  complex,  393 
Octane,  46,  338 

properties,  341 
Octanamide,  339 
Octanoic  acid,  339 
1-Octanol,  338 
Octaves,  Law  of,  104 
Octyl  alcohol,  338 
1-Octylamine,  338 

OH  molecules,  reaction  between,  282 
Oil-drop  experiment,  241 
Oil  of  wintergreen,  346 
Oleomargarine,  407 
Open  hearth  furnace,  404 
Operational  definition,  195 
Orbital   representation   of  chemical 

bonding,  278 
Orbitals 

atomic,  262,  263 

d  and  /,  262 

molecular  shape  and,  293 

overlap,  277 

p,  262 

principal  quantum  number  and, 
261 

s,261 
Ores,  442 

Organic  compounds,  322 
Osmium,  oxidation  numbers,  414 
Ostwald,  W.,  65 

Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  section*. 


Overall  reaction,  201 
Overlap  and  bonding,  277 
Oxidation,  202 

of  alcohols,  336 

of  ammonia,  113 

of  food,  426 

of  hydrogen  chloride,  160 

of  organic  compounds,  332 

of  sulfur,  225,  369 

of  sulfur  dioxide,  225,  369 
Oxidation  numbers,  215,  216 

balancing  redox   reactions  using, 
219 

rules  for  assigning,  219 
Oxidation  potentials,  211,  452 
Oxidation-reduction,  202,  216 

self,  361 
Oxidation-reduction  reactions,  202 

balancing  with  half-reactions,  217 

balancing  with  oxidation  numbers, 
219 
Oxides 

basic,  382 

third-row  elements,  102 
Oxidizing  agent,  215 
Oxyacids 

of  chlorine,  359,  373 

of  phosphorus,  371,  372 

of  sulfur,  371 
Oxygen 

abundance  in  earth's  crust,  441 

boiling  point,  63,  374 

bonding  capacity,  281 

bonding  in  molecules,  295 

bond  length,  295 

double  bond  in,  295 

electron  configuration,  265 

fluorine  compounds,  283 

freezing  point,  63 

heat  of  vaporization,  374 

ionization  energy,  268 

isotopes,  90 

mass  of  atom,  33 

molar  volume,  51,  60 

molecule,  27 

preparation  from  KC103,  46 

pressure-volume  behavior,  18 

reaction  with  nitric  oxide,  26 

solubility,  20 
Ozone,  439 

p  electron,  262 

p  orbitals,  262 

p2  bonding,  291 

p3  bonding,  291 

Packing  in  crystals,  81 

Palmitic  acid,  425 

/raraaminobenzoic  acid,  434 

/raradichlorobenzene,  8 

Paraffin,  43 

paraphthallic  acid,  347 

Partial  pressure,  55,  67,  438 

Particles,  fundamental,  table,  78,  87 

Pauli  Principle,  267 

Pauling,  Linus  C,  299,  300 

Peat,  321 

«-Pentane,  340 

properties,  308 
i.vo- Pen  tane,  340 
Pepsin,  138 
Perchloric  acid,  359 
Perfect  gas,  59 

molar  volume,  60 
Pcrfluoroethane,  362 
Perfluoroethylene,  362 
Periodic  table,  85,  267,  see  inside 


front  cover 

and  atomic  number,  89 

elements  in  each  row  of,  272 

fourth  row,  271,  387 

historical  development,  103 

and  hydrogen  atom,  263 

and  ionization  energy,  267 

second  column,  377 

seventh  column,  352 

seventh  row,  413 

sixth  row,  41 1 
Permanganate  ion 

oxidation  of  Fe+2,  125 

oxidation  of  H2S,  218 

oxidation  of  oxalate  ion,  125 
Petroleum,  322 
pH,  190 
Phase,  70 
Phase  change,  5 

liquid-gas,  66 

solid-liquid,  5,  68 
Phases,  5 

condensed,  27,  68,  78 
Phenacetin,  345 
Phenol,  345 
Phosphides,  368 
Phosphoric  acid,  372 

KA,  191 
Phosphorous  acid,  372 
Phosphorus 

black,  365 

boiling  point,  374 

chemistry  of,  368 

compounds,  102 

heat  of  vaporization,  374 

ionization  energy,  268 

melting  point,  374 

occurrence,  373 

oxyacids,  371,  372 

preparation,  374,  376 

properties,  101 

structure,  366 

white,  120,  365,  366,  369 
Photon,  254 

Photosynthesis,  254,  430 
Pig  iron,  404 
Pitchblende,  385,  442 
Planck's  constant,  254 
Planets 

atmosphere  of,  445 

chemistry  of,  444 

lithosphere  of,  446 
Plastics,  3<6 
Platinum 

catalyst,  action  of,  227 

dioxide,  230 

oxidation  numbers,  414 
"Plexiglas,"  347 
Pluto,  data  on,  444 
Plutonium 

electron  configuration,  415 

oxidation  numbers,  414 
Polar  molecule,  288 
Polonium,  oxidation  numbers,  414 
Polyethylene,  347 
Polymerization,  346 

types  of,  346 
Polymers,  346,  431,  432 
Polysaccharides,  425 
Polystyrene,  345 
Positive  charge,  77,  see  also  Proton 

formation  of,  242 
Positive  ions.  207 

charge  to  mass  ratio,  243 

"seeing,"  242 
Positron,  218 


464 


INDEX 


Potassium,  93 

atomic  radius,  399 

atomic  volume,  410 

chemistry,  95 

electron  configuration,  271 

heat  of  vaporization,  305 

ionization  energy,  268 

properties,  94 
Potassium  chrome  alum,  393 
Potassium  permanganate,  218 
Potential  energy,  115 

barrier,  134 

diagrams,  134,  138,  139,  331 
Potentials,  half-cell 

effect  of  concentration,  213 

measuring,  210 

table,  211,452 
Praseodymium 

properties,  412 

source,  413 
Precipitation 

and  equilibrium,  144 

for  separation,  176 

from  aqueous  solution,  80 

rate,  164 
Predictions 

based  on  E°'s,  212 

of  equilibrium  conditions,  149 

of  formation  of  precipitate,  175 

of  heat  of  reaction,  112 
Predominant  reacting  species,  80 
Pressure,  55 

and  Le  Chatelier's  Principle,  149 

and  reaction  rate,  126 

atmospheric,  53 

cause,  54 

measurement,  53 

of  gases,  53 

partial,  55,  67,  145 

standard,  54 

vapor,  66,  67,  145 
Pressure-volume  relationships,  13, 18 

of  ammonia,  table,  19,  51,  60 

of  hydrogen  chloride,  table,  19 

of  other  gases,  19 

of  oxygen,  table,  14,  18 
Principle 

Le  Chatelier's,  149 

Pauli,  267 
Principal  quantum  number,  260 
Procaine,  345 
Product,  39 

Promethium,  properties,  412 
Promoting  electrons,  284 
Propane,  338 

properties,  341 
1-Propanol,  335 
2-Propanol,  335 
Protactinium,  415 

electron  configuration,  415 

oxidation  numbers,  414 
Propionamide,  339 
Propionic  acid,  338 
Propyl  alcohol,  338 
1-Propylamine,  338 
Propylene,  342 
Propyl  group,  330 
Protein,  348,  432 

helical  structure,  433 

hydrogen  bonds  in,  432 
Proton,  77 

and  atomic  number,  89 

competition  for,  193 

donors,  396 

mass  of,  87,  121 

relation  to  atomic  number,  88 


transfer,  194 

transfer  theory  of  acids,  194 
Ptyalin,  138 

Pullman,  Alberte,  421 
Pullman,  Bernard,  421 
Purification,  70 
Pyruvic  acid,  426 

Qualitative  aspects  of  equilibrium, 

142 
Qualitative  presentation  of  data,  14 
Quantitative  aspects  of  equilibrium, 

151 
Quantitative  analysis,  infrared,  250 
Quantitative  presentation  of  data,  14 
Quantum  mechanics,  259,  260 

and  the  hydrogen  atom,  259 
Quantum  number,  260 

and  hydrogen  atom,  260 

and  orbitals,  261 

principal,  260 

Radioactive  nuclei,  413 
Radioactivity,  416,  442 

types  of,  417 
Radium 

electron  configuration,  378 

occurrence,  385 
Radius 

covalent,  354 

double  bond,  alkaline  earths,  380 

double  bond,  oxygen,  295 

ionic,  355,  380 

metallic,  380 

van  der  Waals,  354,  380 
Radon,  91 

boiling  point,  307 

heat  of  vaporization,  105 

melting  point,  307 
Randall,  M.,  142 
Randomness,  effect  on  equilibrium, 
166,  447 

in  dissolving,  313 
Rare  earths,  272,  see  also  Lantha- 

nides 
Rate 

determining  step,  128 

effect  of  concentration  on,  126, 128 

of  dissolving,  164 

of  opposing  reactions,  148,  155 

of  reactions,  124 
Reactant,  40 

Reacting  species,  predominant,  80 
Reaction  coordinate,  133 
Reaction  heat,  135 

additivity  of,  111 

measurement  of,  111 
Reaction  rates,  124 

factors  affecting,  125 
Reactions,  38,  129 

acid- base,  188 

balancing,  42,  217,  219 

calcium  carbonate  decomposition, 
143 

chemical,  principles  of,  38 

conservation  of  energy  in,  115 

effect  of  heat  on,  108 

endothermic,  40,  135 

equations  for  chemical,  41 

equilibrium  in,  145 

exothermic,  40,  135 

half-cell,  201 

mechanism  of,  127,  128 

nuclear,  120,  121,  419 

oxidation-reduction,  202,  216 


predicting  from  E°,  212 

redox,  203 

substitution,  of  benzene,  344 
Reactive,  284 
Reasoning,  inductive,  3 
Redox,  203 
Reducing  agent,  215 
Reduction,  202 
Regularities,  search  for,  3 
Reichenbach,  Hans,  233 
Relativity,  theory  of,  121 
Repulsive  forces  in  H2,  275 
Resin,  ion  exchange,  413 
Rhenium,  oxidation  number,  414 
Roasting,  46 

Robinson,  Sir  Robert,  321,  351 
Rotational  motion,  118 

and  microwave  spectroscopy,  249 
Rubber  band,  conservation  of  energy 

in,  114 
Rubidium 

heat  of  vaporization,  305 

ionization  energy,  410 

properties  of,  94 
Rule,  14 

Rusting,  45,  85,  405 
Rutherford,  Ernest,  244 
Rutherford 

nuclear  atom,  244 

scattering  experiment,  244 
Rutile,  401 


s  electron,  261 
s  orbitals,  261 
sp  bonding,  292 
sp2  bonding,  292 
sp3  bonding,  292 
Salicylic  acid,  345 
Salt  bridge,  201 
Samarium,  properties,  412 

source,  413 
San  Francisco  to  Los  Angeles  anal- 
ogy, 132 
Saponification,  426 
"Saran,"  347 
Saturated 

hydrocarbons,  326,  340 

solutions,  72 

vapor,  145 
Saturn,  data  on,  444 
Scandium 

atomic  radius,  399 

atomic  volume,  410 

electron  configuration,  389 

heat  of  vaporization,  305 

oxidation  number,  391 

properties,  400 
Scattering  of  alpha  particles,  245 
Science,  2,  15 

activities  of,  1 

communicating  information,  12 

review,  15 
"Sea"  of  electrons,  304 
Seaborg,  Glenn  T.,  420 
Second  column  of  periodic  table,  371 
Second-row  elements,   bonding  ca- 
pacity. 281 
Selenium,  ionization  energy,  410 
Self  oxidation-reduction,  361 
Separation  of  charge,  312 
Separations 

by  crystallization,  413 

by  distillation,  70 

by  precipitation,  176 
Seventh  column  of  periodic  table,  352 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


INDEX 


465 


Sc\cnih  row  of  periodic  tabic.  411 
occurrence,  413 

Significant  figures.  9.  12 

see  also  Uncertainty 
Silica.  309 

dissolves  in  HF,  361 
Silicates.  310 
Silicic  acid,  375 
Silicon 

boiling  point,  374 

chemistry  of.  368 

compounds  of.  102 

crystal  structure,  365 

heat  of  vaporization.  374 

ionization  energy,  268 

melting  point.  374 

occurrence.  373 

ox\  acids.  371 

preparation,  373 

properties,  101 
Silver 

complex  with  ammonia.  395 

nitrate,  79 
Simple  multiple  proportions,  Law  of, 

235 
Sixth  row  of  the  periodic  table,  41 1 
Slaking.  382 
Smog,  125 
Slightly  soluble,  73 
Sodium 

boiling  point.  365 

compounds.  102 

coolant  in  nuclear  pile.  82 

electron  configuration,  265 

heat  of  fusion.  69 

heat  of  vaporization.  365 

hydration  energy.  368 

ionization  energies,  269,  374 

ionization  of.  270 

metallic  solid,  365 

occurrence,  373 

preparation,  238,  356,  373 

properties,  94,  101 

reducing  agent,  367 
Sodium  benzoate,  192 
Sodium  carbonate 

a  base,  184 

manufacture  of,  230 
Sodium  chloride 

crystal.  312 

electrolysis  of,  356 

heat  of  fusion,  69 

lattice,  81 

melting  point,  69 

resemblance  to  inert  gases,  93 
Sodium  hydroxide,  328,  370 

catalyst,  338 
Sodium  hypochlorite,  360 
Sodium  myristate,  426 
Sodium  stearate,  426 
Sodium  uranyl  acetate,  17J 
Solar  atmosphere,  447 
Solids,  27,  65 

atoms  in,  27 

concentration  of  in  equilibrium  ex- 
pressions, 154 

electrical  conductivity  of,  80 

ionic,  79,  81,  311 

and  liquids,  bonding  in,  300 

melting  of,  4,  69 

metallic,  81 

molecular,  102,  301,  306 

network,  102 

solid-liquid  phase  change,  5,  68 

solutions,  71 
Solubility,  72,  72,  144,  164 


alkali  compounds,  170 
alkaline  earth  compounds,  382 

ammonium  compounds,  170 

bromides.  172 

carbonates.  173 

chlorides,  172 

common  compounds,  table  of,  171 

dynamic  nature,  164 

electrolytes  in  water.  313 

equilibrium.  144.  163 

factors  fixing.  165 

gases,  20,  167 

hydrogen  compounds,  170 

hydroxides.  171 

iodides.  172 

iodine  in  ethyl  alcohol.  163 

ionic  solids  in  water,  79 

phosphates.  173 

product,  174 

products,  table  of,  174 

qualitative.  170 

quantitative,  173 

range  of,  73 

sulfates,  172 

sulfides,  172 

sulfites.  173 

table  of.  171 
Soluble.  73 
Solute,  72 
Solutions,  69,  70 

aqueous,  79,  168 

expressing  composition  of,  72 

gaseous,  71 

heat  of.  166 

liquid,  71 

and  pure  substances,  70 

saturated,  72 

solid,  71 

variations  of  properties,  73 
Solvent,  72 

ionizing,  169 
Solvent  properties  and  dipoles,  313 
Sorbitol,  423 
Sprensen  /?H  scale,  190 
Space,  interstellar,  448 
Spectrograph 

mass,  242 

simple,  247 
Spectroscopy,  187 

infrared,  249 

microwave,  249 

X-ray,  248 
Spectrum 

atomic  hydrogen,  253,  255 

light,  247 

tungsten,  hot,  255 
Spontaneous 

chemical  reactions,  212 

endothermic  reactions,  157 

exothermic  reactions,  156 
Square  planar  complex,  395 
Stable,  284 
Stability 

origin  of  bond,  215 
Stalactites,  formation  of,  384 
Standard  half  cell,  210 
Standard  potentials,  209 

and  equilibrium,  275 

table,  452 
Standard  pressure,  54 
Standard  state,  210 
Standard  temperature,  54 
Starch,  425 

iodine  test,  358 

structure,  431 
Stars,  447 


State 

standard,  210 

steady,  144 
Stationary  states,  260 
Station-wagon  analogy,  155 
Stearic  acid,  425 
St  el,  403 

slag,  404 

varieties  of,  402.  404 

wool,  heaung,  7 
Stellar  atmospheres,  448 
Stock,  Alfred  E.,  386 
Stoichiometry,  224 
Stoichiometric  calculations, 

pattern  for,  225 
Storage  batteries 

Edison,  406 

lead,  406 
Stored  energy,  109 
STP,  53 

Strontianite,  385 
Strontium 

atomic  size,  379 

chemistry,  382 

electron  configuration,  378 

heat  of  vaporization,  305 

hydroxide,  K,p,  383 

ionization  energies,  379 

occurrence,  385 

properties,  381 
Structural  formula,  31 
Structural  isomers,  327 
Styrene,  345 
Sublimation,  176 
Substance,  28 

pure,  29,  65,  70 
Substitution    reactions   of  benzene, 

344 
Substrate,  433 
Sucrose,  424 
Sugars,  422 

hydrogen  bonds  in,  424 

hydrolysis  of,  425 

properties,  424 
Sulfanilamide,  434 
Sulfur 

boiling  point,  374 

chemistry,  368 

chloride,  103 

compounds,  102 

heat  of  vaporization,  374 

ionization  energy,  268 

occurrence,  373 

oxidation  of,  225,  369 

oxyacids,  371 

preparation,  374 

properties,  101 

structure,  366 
Sulfur  dioxide 

boiling  point,  64 

formation  of,  225,  369 

molar  volume,  60 

oxidation  of,  225,  369 
Sulfuric  acid 

as  oxidizing  agent,  229 

concentration  of,  228 

density  of,  229 

manufacture,  225 

contact  process,  229 

lead  chamber  process,  228 
Sulfurous  acid,  Ka,  191 
Sulfur  oxides,  225,  369 
Sulfur  oxyacids,  371 
Sulfur  trioxide 

formation  of,  225,  369 

hydration  of,  225 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


466 


INDEX 


Sun,  data  on,  119,  444,  447 
Superconductivity,  58 
Superfluidity,  58 
Symbols,  chemical,  30 

not  from  common  names,  table,  31 
System,  70 

homogeneous  and  heterogeneous, 
126 

open  and  closed,  144 


fi  .,  443 

M,  10 

Tantalum,  oxidation  number,  414 
"Teflon,"  347 
Tehachapi  mountains,  132 
Tejon  pass,  132 
Television  picture  lube,  409 
Temperature 
absolute,  57 
absolute  zero,  58 
earth's  center,  440 
effect  on  equilibrium,  67,  148,  167; 
on  gas  volume,  57;  on  rate,  129; 
on  K,r,  181 
Le  Chatelier's  Principle  and,  150 
Kelvin,  58 

kinetic  energy  and,  56,  131 
Tensile  strength  of  metals,  31 1 
Terbium,  properties,  412 
Termites,  426 
Tetrahedral,  186 

arrangement  of  H20  around  Li+ 

and  H+,  186 
complex,  394 
Tetrathionate  ion,  362 
Tetravalent,  284 

Thallium,  oxidation  numbers,  414 
Theory,  4,  14,  17 
atomic,  17,  22,  28,  234 
Br0nsted-Lowry,  194 
collision,  126 
kinetic,  52 
relativity,  121 
use  of,  28 
Thermal  equilibrium,  56 
Thermite  reaction,  122 
Thermometers,  56 
Thiosulfate  ion,  362 
Third-row  elements,  101 
compounds,  102 
physical  properties,  102 
properties,  table,  101 
Third  row  of  the  periodic  table,  364 
Thomson,  J.  J.,  244 
Thomson  model  of  atom,  244 
Thorium 

electron  configuration.  415 
oxidation  number,  414 
Thorium  oxide,  source,  413 
Threshold  energy,  130 

and  reaction  rate,  132 
Thulium,  properties,  412 
Tin 
catalyst  for  rusting,  405 
use,  406 
Titanium 

atomic  radius,  399 
electron  configuration,  389 
oxidation  numbers,  391 
production,  368 
properties.  400 
Titrations 


acid-base,  188 

definition,  189 
Trajectory,  261 
7>a/!5-isomers,  296,  394 
Transition  elements,  272,  387 

atomic  radii,  399 

complex  ions  of,  392 

electron  configuration,  389 

oxidation  numbers,  391 

properties,  390,  398 

reactivities,  390 

table  of,  400 
Transition  metals,  271 
Translational  motion,  118 
Tritium,  123,  419 
Trivalent,  283 
Tungsten 

carbides.  250 

oxidation  numbers,  414 

spectrum  of  hot,  255 

Ultraviolet  light,  248 

absorption  in  atmosphere,  438 
Uncertainly  in  science,  8 

how  to  indicate,  12 

experimental  errors,  Lab  Manual, 
App.  4 

in  derived  quantities,  10 

in  measurement,  8 
Univalent,  279 

Unsaturated,  hydrocarbons,  342 
Uranium 

compounds,  223 

electron  configuration,  415 

oxidation  number,  414 

preparation,  35 
Uranium  hexafluoride,  35 
Uranus,  data  on,  444 
Urea,  434 

Valence,  286 
Valence  electrons,  269 

and  ionization  energies,  269 
Vanadium 

atomic  radius,  399 

electron  configuration,  389 

oxidation  numbers,  391 

pentoxide  catalyst,  227 

properties,  400,  401 
van  der  Waals  forces,  301 

elements  that  form  molecular  crys- 
tals using,  301 

and  molecular  shape,  307 

and  molecular  size,  307 

and  molecular  substances,  306 

and  number  of  electrons,  306 
van  der  Waals  radius,  354 

halogens,  354 
Vanillin,  345 
Vaporization,  molar  heat  of,  66 

alkali  metals,  94 

alkaline  earths,  381 

copper,  67 

chlorine,  67 

inert  gases,  105 

metals,  305 

neon,  67 

pure  substances,  table,  67 

sodium,  67 

sodium  chloride,  67 

water,  66 
Vapor  pressure,  66,  67,  145 

see  also  Pressure 


Velocity  of  atoms  and  molecules 

distribution,  130 

measuring,  131 
Venus,  data  on,  444 
Vibrational  motion,  118 

and  infrared,  250 
Voltage,  207 
Volume,  50 

relation  to  pressure,  13,  18 

relation  to  temperature,  57 

see  also  Atomic  volume 

see  also  Molar  volume 

Wall,  F.  T.,  108 

Water 

as  a  base,  194 

as  an  acid,  194 

as  a  weak  electrolyte,  180 

concentration   in  equilibrium  ex- 
pressions, 154 

Dalton  and  formula,  250 

decomposition,  40,  115 

density,  154 

electrolysis,  40,  115 

formation  of,  39,  116 

H+  and  OH~  in,  181 

heat  of  fusion,  69 

liquid-gas  phase  change,  66 

model  of  molecule,  31 

molar  heat  of  vaporization,  66 

reaction  with  alkali  metals,  95 

solubility  of  ionic  solids  in,  171 

solution,  electrical  conductivity,  78 

vapor  pressure,  67 
Water  gas,  108 

.heat  effects  in  manufacture  of,  109 
Wavelength  of  light,  246 

relation  to  frequency,  251 
Weight-gas  volume  calculations,  226 
Weight-liquid   volume  calculations, 

228 
Weight-weight  calculations,  226 
Werner,  Alfred,  393 
Wintergreen,  oil  of,  340 
Wollaston,  W.  H.,  258 
Wondering  Why,  5,  8,  16,  155 
Woodward,  Robert  Burns,  435 
Work,  114 
Writing  equations,  42 

Xenon,  91 

boiling  point,  307 

heat  of  vaporization,  105 

melting  point,  307 
X-Rays,  248 

diffraction  patterns,  248 

Yankee  stadium,  88 

Yeast,  426 

Ytterbium,  properties,  412 

Yttrium,  heat  of  vaporization,  305 

Z,  (atomic  number),  389 
Zinc 

atomic  radius,  399 

complex  with  ammonia,  395 

electron  configuration,  389 

galvanizing,  405 

oxidation  of,  203 

oxidation  numbers,  391 

properties,  400,  409 
Zinc  blende,  409 


Boldface  numbers  refer  to  definitions.  Italic  numbers  refer  to  sections. 


t£      r\j      «o 

*    3:    *     8 


■   *  *S  o\  *  *$ 


On       q 

n 

*0         <N 

<o 


Co 

ir 

**» 

V 

I. 
<» 

& 


<o 

> 
a 

« 

K 


INTERNATIONAL  ATOMIC  WEIGHTS 


SYMBOL 


ATOMIC 
NUMBER 


ATOMIC 
WEIGHT 


NAME 


SYMBOL 


ATOMIC 
NUMBER 


ATOMIC 
WEIGHT 


Actinium 

Ac 

89 

(227) 

Mercury 

Hg 

80 

200.6 

Aluminum 

Al 

13 

27.0 

Molybdenum 

Mo 

42 

95.9 

Americium 

Am 

95 

(243) 

Neodymium 

Nd 

60 

144.2 

Antimony 

Sb 

51 

121.8 

Neon 

Ne 

10 

20.2 

Argon 

Ar 

18 

39.9 

Neptunium 

Np 

93 

(237) 

Arsenic 

As 

33 

74.9 

Nickel 

Ni 

28 

58.7 

Astatine 

At 

85 

(210) 

Niobium 

Nb 

41 

92.9 

Barium 

Ba 

56 

137.3 

Nitrogen 

N 

7 

14.01 

Berkelium 

Bk 

97 

245 

Osmium 

Os 

76 

190.2 

Beryllium 

Be 

4 

9.01 

Oxygen 

O 

8 

16.00 

Bismuth 

Bi 

83 

209.0 

Palladium 

Pd 

46 

106.4 

Boron 

B 

5 

10.8 

Phosphorus 

P 

15 

31.0 

Bromine 

Br 

35 

79.9 

Platinum 

Pt 

78 

195.1 

Cadmium 

Cd 

48 

112.4 

Plutonium 

Pu 

94 

(242) 

Calcium 

Ca 

20 

40.1 

Polonium 

Po 

84 

210 

Californium 

Cf 

98 

(251) 

Potassium 

K 

19 

39.1 

Carbon 

C 

6 

12.01 

Praseodymium 

Pr 

59 

140.9 

Cerium 

Ce 

58 

140.1 

Promethium 

Pm 

61 

(147) 

Cesium 

Cs 

55 

132.9 

Protactinium 

Pa 

91 

(231) 

Chlorine 

CI 

17 

35.5 

Radium 

Ra 

88 

(226) 

Chromium 

Cr 

24 

52.0 

Radon 

Rn 

86 

(222) 

Cobalt 

Co 

27 

58.9 

Rhenium 

Re 

75 

186.2 

Copper 

Cu 

29 

63.5 

Rhodium 

Rh 

45 

102.9 

Curium 

Cm 

96 

(247) 

Rubidium 

Rb 

37 

85.5 

Dysprosium 

Dy 

66 

162.5 

Ruthenium 

Ru 

44 

101.1 

Einsteinium 

Es 

99 

(254) 

Samarium 

Sm 

62 

150.4 

Erbium 

Er 

68 

167.3 

Scandium 

Sc 

21 

45.0 

Europium 

Eu 

63 

152.0 

Selenium 

Se 

34 

79.0 

Fermium 

Fm 

100 

(253) 

Silicon 

Si 

14 

28.1 

Fluorine 

F 

9 

19.0 

Silver 

Ag 

47 

107.9 

Francium 

Fr 

87 

(223) 

Sodium 

Na 

11 

23.0 

Gadolinium 

Gd 

64 

157.3 

Strontium 

Sr 

38 

87.6 

Gallium 

Ga 

31 

69.7 

Sulfur 

S 

16 

32.1 

Germanium 

Ge 

32 

72.6 

Tantalum 

Ta 

73 

180.9 

Gold 

Au 

79 

197.0 

Technetium 

Tc 

43 

(99) 

Hafnium 

Hf 

72 

178.5 

Tellurium 

Te 

52 

127.6 

Helium 

He 

2 

4.00 

Terbium 

Tb 

65 

158.9 

Holmium 

Ho 

67 

164.9 

Thallium 

Tl 

81 

204.4 

Hydrogen 

H 

1 

1.008 

Thorium 

Th 

90 

232.0 

Indium 

In 

49 

114.8 

Thulium 

Tm 

69 

168.9 

Iodine 

I 

53 

126.9 

Tin 

Sn 

50 

118.7 

Iridium 

Ir 

77 

192.2 

Titanium 

Ti 

22 

47.9 

Iron 

Fe 

26 

55.8 

Tungsten 

W 

74 

183.9 

Krypton 

Kr 

36 

83.8 

Uranium 

U 

92 

238.0 

Lanthanum 

La 

57 

138.9 

Vanadium 

V 

23 

50.9 

Lead 

Pb 

82 

207.2 

Xenon 

Xe 

54 

131.3 

Lithium 

Li 

3 

6.94 

Ytterbium 

Yb 

70 

173.0 

Lutetium 

Lu 

71 

175.0 

Yttrium 

Y 

39 

88.9 

Magnesium 

Mg 

12 

24.3 

Zinc 

Zn 

30 

65.4 

Manganese 

Mn 

25 

54.9 

Zirconium 

Zr 

40 

91.2 

Mendelevium 

Md 

101 

(256) 

Parenthetical  names  refer  to  radioactive  elements;  the  mass  number  (not  the  atomic  weight)  of  the  isotope  with  largest 
half-life  is  usually  given. 

*  Latest  values  recommended  by  the  International  Union  of  Pure  and  Applied  Chemistry,  1961. 


0-7167-0001-8