Skip to main content

Full text of "Higher mathematics for students of chemistry and physics, with special reference to practical work"

See other formats


HIGHER  MATHEMATICS 


FOK 


STUDENTS  OF  CHEMISTRY  AND 
PHYSICS 


WITH  SPECIAL  REFERENCE  TO  PRACTICAL  WORE 


BY 


J.    W.     MELLOR,    D.Sc. 


NEW  IMPRESSION 


LONGMANS,    GREEN    AND    CO. 

89  PATERNOSTER   ROW,   LONDON,   E.C.  4 

NEW  YORK,  TORONTO 

BOMBAY,  CALCUTTA  AND  MADRAS 

1922 

[All  rights  reserved] 


2-87^ 


5 


37 


'*  The  first  thing  to  be  attended  to  in  reading  any  algebraio  treatise  is  the 
gaining  a  perfect  understanding  of  the  different  processes  there  exhibited, 
and  of  their  connection  with  one  another.  This  cannot  be  attained  by  a 
mere  reading  of  the  book,  however  great  the  attention  whioh  may  be  given. 
It  is  impossible  in  a  mathematical  work  to  fill  up  every  process  in  the 
manner  in  which  it  must  be  filled  up  in  the  mind  of  the  student  before 
he  can  be  said  to  have  completely  mastered  it.  Many  results  must  be  given 
of  which  the  details  are  suppressed,  such  are  the  additions,  multiplications, 
extractions  of  square  root,  etc.,  with  which  the  investigations  abound. 
These  must  not  be  taken  in  trust  by  the  student,  but  must  be  worked  by 
his  own  pen,  whioh  must  never  be  out  of  his  hand,  while  engaged  in  any 
algebraical  process." — De  Morgan,  On  the  Study  and  Difficulties  of  Mathe 
mattes,  188?. 


Made  in   Great  Britah* 


% 


^  fj  £*  r  \  n  €\ 


PREFACE  TO  THE  FOURTH  EDITION. 

The  fourth  edition  is  materially  the  same  as  the  third.  I 
have,  however,  corrected  the  misprints  which  have  been 
brought  to  my  notice  by  a  number  of  students  of  the  book, 
and  made  a  few  verbal  alterations  and  extensions  of  the  text. 
I  am  glad  to  say  that  a  German  edition  has  been  published  ; 
and  to  observe  that  a  large  number  of  examples,  etc.,  peculiar 
to  this  work  and  to  my  Chemical  Statics  and  Dynamics  have 
been  "  absorbed  "  into  current  literature. 

J.  W.  M. 

The  Villas,  Stoke-on-Tbent, 
13th  December,  1912. 


PREFACE  TO  THE  SECOND  EDITION. 

I  am  pleased  to  find  that  my  attempt  to  furnish  an  Intro-, 
duction  to  the  Mathematical  Treatment  of  the  Hypotheses 
and  Measurements  employed  in  scientific  work  has  been  so 
much  appreciated  by  students  of  Chemistry  and  Physics. 
In  this  edition,  the  subject-matter  has  been  rewritten,  and 
many  parts  have  been  extended  in  order  to  meet  the  growing 
tendency  on  the  part  of  physical  chemists  to  describe  their 
ideas  in  the  unequivocal  language  of  mathematics. 

J.  W.  M. 

Uh  July,  1905. 


vii 


PREFACE  TO  THE  FIRST  EDITION. 

It  is  almost  impossible  to  follow  the  later  developments  of 
physical  or  general  chemistry  without  a  working  knowledge 
of  higher  mathematics.  I  have  found  that  the  regular 
text-books  of  mathematics  rather  perplex  than  assist  the 
chemical  student  who  seeks  a  short  road  to  this  knowledge, 
for  it  is  not  easy  to  discover  the  relation  which  the  pure 
abstractions  of  formal  mathematics  bear  to  the  problems 
which  every  day  confront  the  student  of  Nature's  laws, 
and  realize  the  complementary  character  of  mathematical 
and  physical  processes. 

During  the  last  five  years  I  have  taken  note  of  the 
chief  difficulties  met  with  in  the  application  of  the  mathe- 
matician's x  and  y  to  physical  chemistry,  and,  as  these  notes 
have  grown,  I  have  sought  to  make  clear  how  experimental 
results  lend  themselves  to  mathematical  treatment.  I  have 
found  by  trial  that  it  is  possible  to  interest  chemical  students 
and  to  give  them  a  working  knowledge  of  mathematics 
by  manipulating  the  results  of  physical  or  chemical  ob- 
servations. 

I  should  have  hesitated  to  proceed  beyond  this  experi- 
mental stage  if  I  had  not  found  at  The  Owens  College  a 
set  of  students  eagerly  pursuing  work  in  different  branches 

of  physical  chemistry,  and  most  of  them  looking  for  help 

ix  ) 


x  .  PREFACE. 

in  the  discussion  of  their  results.  When  I  told  my  plan 
to  the  Professor  of  Chemistry  he  encouraged  me  to  write 
this  book.  It  has  been  my  aim  to  carry  out  his  suggestion, 
so  I  quote  his  letter  as  giving  the  spirit  of  the  book, 
which  I  only  wish  I  could  have  carried  out  to  the  letter. 

"The  Owens  College, 
"  Manchester. 
"My  Dear  Mellor, 

"  If  you  will  convert  your  ideas  into  words  and  write  a 
book  explaining  the  inwardness  of  mathematical  operations  as  applied 
to  chemical  results,  I  believe  you  will  confer  a  benefit  on  many  students 
of  chemistry.  We  chemists,  as  a  tribe,  fight  shy  of  any  symbols 
but  our  own.  I  know  very  well  you  have  the  power  of  winning  new 
results  in  chemistry  and  discussing  them  mathematically.  Can  you 
lead  us  up  the  high  hill  by  gentle  slopes?  Talk  to  us  chemically  to 
beguile  the  way  ?  Dose  us,  if  need  be,  '  with  learning  put  lightly,  like 
powder  in  jam '  ?  If  you  feel  you  have  it  in  you  to  lead  the  way  we 
will  try  to  follow,  and  perhaps  some  of  the  youngest  of  us  may  succeed 
Wouldn't  this  be  a  triumph  worth  working  for  ?     Try. 

"  Yours  very  truly, 

"H.  B.  Dixon." 
May,  1902. 


CONTENTS. 

(The  bracketed  numbers  refer  to  pages.) 

CHAPTER  I.— THE  DIFFERENTIAL  CALCULUS. 

§  1.  On  the  nature  of  mathematical  reasoning  (3)  ;  §  2.  The  differential  co- 
efficient (6)  ;  §  3.  Differentials  (10)  ;  §  4.  Orders  of  magnitude  (10)  ; 
§  5.  Zero  and  infinity  (12)  ;  §  6.  Limiting  values  (13) ;  §  7.  The  differ- 
ential coefficient  of  a  differential  coefficient  (17)  ;  §  8.  Notation  (19) ; 
§  9.  Functions  (19) ;  §  10.  Proportionality  and  the  variation  constant 
(22) ;  §  11.  The  laws  of  indices  and  logarithms  (24)  ;  §  12.  Differentia- 
tion, and  its  uses  (29) ;  §  13.  Is  differentiation  a  method  of  approxima- 
tion only  ?  (32) ;  §  14.  The  differentiation  of  algebraic  functions  (35) ; 
§  15.  The  gas  equations  of  Boyle  and  van  der  Waals  (46) ;  §  16.  The 
differentiation  of  trigonometrical  function's  (47) ;  §  17.  The  differentia- 
tion of  inverse  trigonometrical  functions.  The  differentiation  of  angles 
(49)  ;  §  18.  The  differentiation  of  logarithms  (51)  ;  §  19.  The  differ- 
ential coefficient  of  exponential  functions  (54)  ;  §  20.  The  "  compound 
interest  law  "  in  Nature  (56) ;  §  21.  Successive  differentiation  (64)  : 
§  22.  Partial  differentiation  (68) ;  §  23.  Euler's  theorem  on  homo 
geneous  functions  (75)  ;  §  24.  Successive  partial  differentiation  (76) ; 
§  25.  Complete  or  exact  differentials  (77) ;  §  26.  Integrating  factors 
(77) ;  §  27.  Illustrations  from  thermodynamics  (79). 

CHAPTER  II.— COORDINATE  OR  ANALYTICAL  GEOMETRY. 

§  28.  Cartesian  coordinates  (83) ;  §  29.  Graphical  representation  (85) ;  §  30. 
Practical  illustrations  of  graphical  representation  (86) ;  §  81.  Properties 
of  straight  lines  (89)  ;  §  32.  Curves  satisfying  conditions  (93) ;  §  33. 
Changing  the  coordinate  axes  (96)  ;  §  34.  The  circle  and  its  equation 
(97)  ;  §  35.  The  parabola  and  its  equation  (99) ;  §  36.  The  ellipse  and 
its  equation  (100) ;  §  37.  The  hyperbola  and  its  equation  (101) ;  §  38.  The 
tangent  to  a  curve  (102) ;  §  39.  A  study  of  curves  (106) ;  §  40.  The  rec- 
tangular or  equilateral  hyperbola  (109) ;  §  41.  Illustrations  of  hyper- 
bolic curves  (110) ;  §  42.  Polar  coordinates  (114) ;  §  43.  Spiral  curves 
(116) ;  §  44.  Trilinear  coordinates  and  ^riangular  diagrams  (118) ;  §  45. 
Orders  of  curves  (120) ;  §  46.  Coordinate  geometry  in  three  dimensions. 

xi 


xii  CONTENTS. 

Geometry  in  space  (121) ;  §  47.  Lines  in  three  dimensions  (127) ;  §  48. 
Surfaces  and  planes  (132) ;  §  49.  Pt- iodic  or  harmonic  motion  (135) ; 
§  50.  Generalized  forces  and  coordinates  (139). 

CHAPTER  III.— FUNCTIONS  WITH  SINGULAR  PROPERTIES. 

§  51.  Continuous  and  discontinuous  functions  (142) ;  §  52.  Discontinuity  ac- 
companied by  "  breaks  "  (143)  ;  §  53.  The  existence  of  hydrates  in 
solution  (145) ;  §  54.  The  smoothing  of  curves  (148) ;  §  55.  Discon- 
tinuity accompanied  by  change  of  direction  (149) ;  §  56.  The  triple 
point  (151) ;  §  57.  Maximum  and  minimum  values  of  a  function  (154) ; 
§  58.  How  to  find  maximum  and  minimum  values  of  a  function  (155) ; 
§  59.  Points  of  inflexion  (158) ;  §  60.  How  to  find  whether  a  curve  is 
concave  or  convex  (159) ;  §  61.  How  to  find  points  of  inflexion  (160) ; 
§  62.  Six  problems  in  maxima  and  minima  (161) ;  §  63.  Singular  points 
(168) ;  §  64.  pv-Gurves  (172) ;  §  65.  Imaginary  quantities  (176) ;  §  66. 
Curvature  (178) ;  §  67.  Envelopes  (182). 

CHAPTER  IV.— THE  INTEGRAL  CALCULUS. 

§  68.  Integration  (184) ;  §  69.  Table  of  standard  integrals  (192) ;  §  70.  The 
simpler  methods  of  integration  (192) ;  §  71.  How  to  find,  a  value  foi 
the  integration  constant  (198) ;  §  72.  Integration  by  the  substitution  of 
a  new  variable  (200) ;  §  73.  Integration  by  parts  (205) ;  §  74.  Integra- 
tion by  successive  reduction  (206);  §  75.  Reduction  formulss— for  re- 
ference (208)  ;  §  76.  Integration  by  resolution  into  partial  fractions 
(212)  ;  §  77.  The  velocity  of  chemical  reactions  (218) ;  §  78.  Chemical 
equilibria — incomplete  or  reversible  reactions  (225) ;  §  79.  Fractional 
precipitation  (229) ;  §  80.  Areas  enclosed  by  curves.  To  evaluate  de- 
finite integrals  (231) ;  §  81.  Mean  values  of  integrals  (234) ;  §  82.  Area? 
bounded  by  curves.  Work  diagrams  (237) ;  §  83.  Definite  integrals  and 
their  properties  (240) ;  §  84.  To  find  the  length  of  any  curve  (245) ; 
§  85.  To  find  the  area  of  a  surface  of  revolution  (247) ;  §  86.  To  find  the 
volume  of  a  solid  of  revolution  (248) ;  §  87.  Successive  integration. 
Multiple  integrals  (249) ;  §  88.  The  isothermal  expansion  of  gases  (254) ; 
§  89.  The  adiabatic  expansion  of  gases  (257) ;  §  90.  The  influence  of 
temperature  on  chemical  and  physical  changes  (262). 

CHAPTER  V.— INFINITE  SERIES  AND  THEIR  USES. 

§  91.  What  is  an  infinite  series  ?  (266) ;  §  92.  Washing  precipitates  (269) ; 
'  §  93.  Tests  for  convergent  series  (271) ;  §  94.  Approximate  calculations 
in  scientific  work  (273) ;  §  95.  Approximate  calculations  by  means  of 
infinite  series  (276) ;  §  96.  Maclaurin's  theorem  (280) ;  §  97.  Useful  de- 
ductions from  Maclaurin's  theorem  (282) ;  §  98.  Taylor's  theorem 
(286) ;  §  99.  The  contact  of  curves  (291)  ;  §  100.  Extension  of  Taylor's 
theorem  (292) ;  §  101.  The  determination  of  maximum  and  minimum 
values  of  a  function  by  means  of  Taylor's  series  (293) ;  §  102.  La- 
grange's theorem  (301) ;   §  103.  Indeterminate  functions  (304) ;  §  104. 


CONTENTS.  xiii 

The  calculus  of  finite  differences  (308);  §  105.  Interpolation  (310); 
§  106.  Differential  coefficients  from  numerical  observations  (318) ;  §  107. 
How  to  represent  a  set  of  observations  by  means  of  a  formula  (322) ; 
§  108.  To  evaluate  the  constants  in  empirical  or  theoretical  formulae 
(324) ;  §  109.  Substitutes  for  integration  (333) ;  §  110.  Approximate 
integration  (335) ;  §  111.  Integration  by  infinite  series  (341) ;  §  112 
The  hyperbolic  functions  (346). 

CHAPTER  VI.— HOW  TO  SOLVE  NUMERICAL  EQUATIONS. 

§  118.  Some  general  properties  of  the  roots  of  equations  (352) ;  §  114.  Graphio 
methods  for  the  approximate  solution  of  numerical  equations  (353); 
§  115.  Newton's  method  for  the  approximate  solution  of  numerical 
equations  (358) ;  §  116.  How  to  separate  equal  roots  from  an  equation 
(359) ;  §  117.  Sturm's  method  of  locating  the  real  and  unequal  roots 
of  a  numerical  equation  (360);  §  118.  Horner's  method  for  approxi- 
mating to  the  real  roots  of  numerical  equations  (363) ;  §  119.  Van  der 
Waals'  equation  (367). 

CHAPTER  VII.— HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS. 

§  120.  The  solution  of  a  differential  equation  by  the  separation  of  the  vari- 
ables (370)  ;  §  121.  What  is  a  differential  equation  ?  (374)  ;  §  122. 
Exact  differential  equations  of  the  first  order  (378) ;  §  123.  How  to  find 
integrating  factors  (381) ;  §  124.  Physical  meaning  of  exact  differentials 
(384) ;  §  125.  Linear  differential  equations  of  the  first  order  (387) ; 
§  126.  Differential  equations  of  the  first  order  and  of  the  first  or  higher 
degree — Solution  by  differentiation  (390);  §  127.  Clairaut's  equation 
(391) ;  §  128.  Singular  solutions  (392) ;  §  129.  Symbols  of  operation 
(396);  §180.  Equations  of  oscillatory  motion  (396);  §  131.  The  linear 
equation  of  the  second  order  (399) ;  §  132.  Damped  oscillations  (404) ; 
§  133.  Some  degenerates  (410) ;  §  134.  Forced  oscillations  (413) ;  §  135. 
How  to  find  particular  integrals  (418) ;  §  136.  The  gamma  function 
(423) ;  §  137.  Elliptic  integrals  (426) ;  §  138.  The  exact  linear  differ- 
ential equation  (431) ;  §  139.  The  velocity  of  consecutive  chemioal 
reactions  (433) ;  §  140.  Simultaneous  equations  with  constant  coeffici- 
ents (441) ;  §  141.  Simultaneous  equations  with  variable  coefficients 
(444) ;  §  142.  Partial  differential  equations  (448) ;  §  143.  What  is  the 
solution  of  a  partial  differential  equation  ?  (449) ;  §  144.  The  linear 
partial  differential  equation  of  the  first  order  (452) ;  §  145.  Some 
special  forms  (454)  ;  §  146.  The  linear  partial  equation  of  the  second 
order  (457) ;  §  147.  The  approximate  integration  of  differential  equa- 
tions (463). 

CHAPTER  VIII.— FOURIER'S  THEOREM. 

§  148.  Fourier's  series  (468) ;  §  149.  Evaluation  of  the  constants  in  Fourier's 
series  (470) ;  §  150.  The  development  of  a  function  in  a  trigono- 
metrical series  (473) ;  §  151.  Extension  of  Fourier's  series  (477) ;  §  152. 


xiv  CONTENTS. 

Fourier's  linear  diffusion  law  (481) ;  §  153.  Application  to  the  diffusion 
of  salts  in  solution  (483) ;  §  154.  Application  to  problems  on  the  con- 
duction of  heat  (493). 

CHAPTER  IX.— PROBABILITY  AND  THE  THEORY  OF  ERRORS. 

§  155.  Probability  (498) ;  §  156.  Application  to  the  kinetic  theory  of  gases 
(504) ;  §  157.  Errors  of  observation  (510) ;  §  158.  The  "  law  "  of  errors 
(511) ;  §  159.  The  probability  integral  (518) ;  §  160.  The  best  repre- 
sentative value  for  a  set  of  observations  (518) ;  §  161.  The  probable 
error  (521) ;  §  162.  Mean  and  average  errors  (524) ;  §  153.  Numerical 
values  of  the  probability  integrals  (531) ;  §  164.  Maxwell's  law  of  dis- 
tribution of  molecular  velocities  (534) ;  §  165.  Constant  errors  (537)  ; 
§  166.  Proportional  errors  (539) ;  §  167.  Observations  of  different  de- 
grees of  accuracy  (548) ;  §  168.  Observations  limited  by  conditions 
(555) ;  §  169.  Gauss'  method  of  solving  a  set  of  linear  observation  equa- 
tions (557) ;  §  170.  When  to  reject  suspected  observations  (563). 

CHAPTER  X.— THE  CALCULUS  OF  VARIATIONS. 

§  171.  Differentials  and  variations  (567) ;  §  172.  The  variation  of  a  func^on 
(568) ;  §  173.  The  variation  of  an  integral  with  fixed  limits  (569) ;  §  174. 
Maximum  or  minimum  values  of  a  definite  integral  (570) ;  §  175.  The 
variation  of  an  integral  with  variable  limits  (573)  ;  §  176.  Relative 
maxima  and  minima  (575) ;  §  177.  The  differentiation  of  definite  in- 
tegrals (577) ;   §  178.   Double  and  triple  integrals  (577). 

CHAPTER  XL—DETERMINANTS. 

§  179.  Simultaneous  equations  (580) ;  §  180.  The  expansion  of  determinants 
(583) ;  §  181.  The  solution  of  simultaneous  equations  (584) ;  §  182.  Test 
for  consistent  equations  (585) ;  §  183.  Fundamental  properties  of  de- 
terminants (587) ;  §  184.  The  multiplication  of  determinants  (589) ; 
§  185.  The  differentiation  of  determinants  (590)  ;  §  186.  Jacobians  and 
Hessians  (591) ;  §  187.  Illustrations  from  thermodynamics  (594) ;  §  188. 
Study  of  surfaces  (595). 

APPENDIX    I.— COLLECTION    OF   FORMULA   AND   TABLES   FOR 

REFERENCE. 

§  189.  Calculations  with  small  quantities  (601) ;  §  190.  Permutations  and 
combinations  (602) ;  §  191.  Mensuration  formulae  (603) ;  §  192.  Plane 
trigonometry  (606) ;  §  193.  Relations  among  the  hyperbolic  functions 
(612). 


CONTENTS.  xv 


APPENDIX  II.— REFERENCE  TABLES. 

E.  Singular  values  of  functions  (168) ;  II.  Standard  integrals  (193) ;  III. 
Standard  integrals  (Hyperbolic  functions)  (349  and  614) ;  IV.  Numerical 
values  of  the  hyperbolic  sines,  cosines,  «*,  and  e~x  (616) ;  V.  Common 
logarithms  of  the  gamma  function  (426) ;  VI.  Numerical  values  of  the 

factor  °'f74      (619) ;    VII.  Numerical  values  of  the  factor 


>/n-  1  \fn(n  -  )1 

(619) ;  VIII.  Numerical  values  of  the  factor    j       _  ^     (620) ;     IX. 

Numerical  values  of  the  factor  — - t=*  (620) ;    X.    Numerical  values 

of   the  probability  integral    -^  I    e  d  (hx),  (621) ;  XI.  Numerical 

values  of  the  probability  integral  —j=-\9  '<*(;:)  i  (622);  XII.  Nu- 
merical values  for  the  application  of  Ghauvenet's  criterion  (623) ;  XIII. 
Circular  or  radian  measure  of  angles  (624) ;  XIV.  Numerical  values  of 
some  trigonometrical  ratios  (609) ;  XV.  Signs  of  the  trigonometrical 
ratios  (610) ;  XVI.  Comparison  of  hyperbolic  and  trigonometrical 
functions  (614);  XVII.  Numerical  values  of  e*2  and  of  e~x%  (626); 
XVIII.  Natural  logarithms  of  numbers  (627). 


INTRODUCTION. 

"  Bient6t  le  calcul  math^matique  sera  tout  aussi  utile  au  chimiste 
que  la  balance."  * — P.  Schutzenberger. 

When  Isaac  Newton  communicated  the  manuscript  of  his 
"  Methodus  fluxionum  "  to  his  friends  in  1669  he  furnished 
science  with  its  most  powerful  and  subtle  instrument  of 
research.  The  states  and  conditions  of  matter,  as  they 
occur  in  Nature,  are  in  a  state  of  perpetual  flux,  and  these 
qualities  may  be  effectively  studied  by  the  Newtonian  method 
whenever  they  can  be  referred  to  number  or  subjected  to 
measurement  (real  or  imaginary).  By  the  aid  of  Newton's 
calculus  the  mode  of  action  of  natural  changes  from  moment 
to  moment  can  be  portrayed  as  faithfully  as  these  words 
represent  the  thoughts  at  present  in  my  mind.  From  this, 
the  law  which  controls  the  whole  process  can  be  determined 
with  unmistakable  certainty  by  pure  calculation — the  so- 
called  Higher  Mathematics. 

This  work  starts  from  the  thesis2  that  so  far  as  the 
investigator  is  concerned, 

Higher  Mathematics  is  the  art  of  reasoning  about  the 
numerical  relations  between  natural  phenomena ;  and  the 
several  sections  of  Higher  Mathematics  are  different  modes 
of  viewing  these  relations. 

1  Translated :  "Ere  long  mathematics  will  be  as  useful  to  the  chemist  as  the 
balance ".     (1880.) 

2  In  the  Annalen  der  NaturphUosophie,  1,  50,  1902,  W.  Ostwald  maintains  that 
mathematics  is  only  a  language  in  which  the  results  of  experiments  may  be  conveni  - 
ently  expressed  ;  and  from  tli is  standpoint  criticises  I.  Kant's  Metaphysical  Founda- 
tions of  Natural  Science. 

xvii  6  * 


xviii  INTRODUCTION. 

For  instance,  I  have  assumed  that  the  purpose  of  the 
Differential  Calculus  is  to  inquire  how  natural  phenomena 
change  from  moment  to  moment.  This  change  may  be 
uniform  and  simple  (Chapter  I.);  or  it  may  be  associated 
with  certain  so-called  "  singularities  "  (Chapter  III.).  The 
Integral  Calculus  (Chapters  IV.  and  VII.)  attempts  to  deduce 
the  fundamental  principle  governing  the  whole  course  of 
any  natural  process  from  the  law  regulating  the  momentary 
states.  Coordinate  Geometry  (Chapter  II.)  is  concerned 
with  the  study  of  natural  processes  by  means  of  "  pictures  " 
or  geometrical  figures.  Infinite  Series  (Chapters  V.  and 
VIII.)  furnish  approximate  ideas  about  natural  processes 
when  other  attempts  fail.  From  this,  then,  we  proceed  to 
study  the  various  methods  —  tools  —  to  be  employed  in 
Higher  Mathematics. 

This  limitation  of  the  scope  of  Higher  Mathematics 
enables  us  to  dispense  with  many  of  the  formal  proofs  of 
rules  and  principles.  Much  of  Sidgwick's l  trenchant  indict- 
ment of  the  educational  value  of  formal  logic  might  be  urged 
against  the  subtle  formalities  which  prevail  in  "  school  " 
mathematics.  While  none  but  logical  reasoning  could  be 
for  a  moment  tolerated,  yet  too  often  "  its  most  frequent 
work  is  to  build  a  perns  asinorum  over  chasms  that  shrewd 
people  can  bestride  without  such  a  structure  ".2 

So  far  as  the  tyro  is  concerned  theoretical  demonstrations 
are  by  no  means  so  convincing  as  is  sometimes  supposed. 
It  is  as  necessary  to  learn  to  "  think  in  letters  "  and  to 
handle  numbers  and  quantities  by  their  symbols  as  it  is  to 
learn  to  swim  or  to  ride  a  bicycle.  The  inutility  of  "  general 
proofs  "is  an  everyday  experience  to  the  teacher.  The  be- 
ginner only  acquires  confidence  by  reasoning  about  something 
which  allows  him  to  test  whether  his  results  are  true  or 
false  ;  he  is  really  convinced  only  after  the  principle  has 
been  verified  by  actual  measurement  or  by  arithmetical  il- 
lustration.    "  The  best  of  all  proofs,"  said  Oliver  Heaviside 

>  A.  Sidgwick,  The  Use  of  Words  in  Reasoning.     (A.  &  C.  Black,  London.) 
2  0.  W.  Holmes,  The  Autocrat  of  the  Breakfast  Table.    (W.  Scott,  London.) 


INTRODUCTION.  xix 

in  a  recent  number  of  the  Electrician,  "  is  to  set  out  the  fact 
descriptively  so  that  it  can  be  seen  to  be  a  fact  ".  Re- 
membering also  that  the  majority  of  students  are  only 
interested  in  mathematics  so  far  as  it  is  brought  to  bear 
directly  on  problems  connected  with  their  own  work,  I  have, 
especially  in  the  earlier  parts,  explained  any  troublesome  prin- 
ciple or  rule  in  terms  of  some  well-known  natural  process.  For 
example,  the  meaning  of  the  differential  coefficient  and  of  a 
limiting  ratio  is  first  explained  in  terms  of  the  velocity  of  a 
chemical  reaction  ;  the  differentiation  of  exponential  functions 
leads  us  to  compound  interest  and  hence  to  the  "  Compound 
Interest  Law "  in  Nature ;  the  general  equations  of  the 
straight  line  are  deduced  frorn  solubility  curves ;  discon- 
tinuous functions  lead  us  to  discuss  Mendeleeff's  work  on  the 
existence  of  hydrates  in  solutions ;  Wilhelmy's  law  of  mass 
action  prepares  us  for  a  detailed  study  of  processes  of  inte- 
gration ;  Harcourt  and  Esson's  work  introduces  the  study  of 
simultaneous  differential  equations  ;  the  equations  of  motion 
serve  as  a  basis  for  the  treatment  of  differential  equations  of 
the  second  order  ;  Fourier's  series  is  applied  to  diffusion 
phenomena,  etc.,  etc.  Unfortunately,  this  plan  has  caused 
the  work  to  assume  more  formidable  dimensions  than  if  the 
precise  and  rigorous  language  of  the  mathematicians  had 
been  retained  throughout. 

I  have  sometimes  found  it  convenient  to  evade  a  tedious 
demonstration  by  reference  to  the  "  regular  text-books".  In 
such  cases,  if  the  student  wants  to  "  dig  deeper,"  one  of  the 
following  works,  according  to  subject,  will  be  found  sufficient : 
B.  Williamson's  Differential  Calculus,  also  the  same  author's 
Integral  Calculus,  London,  1899  ;  A.  E.  Forsyth's  Differential 
Equations,  London,  1902 ;  W.  W.  Johnson's  Differential 
Equations,   New  York,    1899. 

Of  course,  it  is  not  always  advisable  to  evade  proofs  in 
this  summary  way.  The  fundamental  assumptions — the  so- 
called  premises — employed  in  deducing  some  formulae  must 
be  carefully  checked  and  clearly  understood.  However 
correct  the  reasoning  may  have  been,  any  limitations  intro- 
duced as  premises  must,  of  necessity,  reappear  in  the  con- 


xx  INTRODUCTION. 

elusions.  The  resulting  formulae  can,  in  consequence,  only 
be  applied  to  data  which  satisfy  the  limiting  conditions. 
The  results  deduced  in  Chapter  IX.  exemplify,  in  a  forcible 
manner,  the  perils  which  attend  the  indiscriminate  applica- 
tion of  mathematical  formulae  to  experimental  data.  Some 
formulae  are  particularly  liable  to  mislead.  The  "  probable 
error  "  is  one  of  the  greatest  sinners  in  this  respect. 

The  teaching  of  mathematics  by  means  of  abstract 
problems  is  a  good  old  practice  easily  abused.  The  abuse 
has  given  rise  to  a  widespread  conviction  that  "  mathematics 
is  the  art  of  problem  solving,"  or,  perhaps,  the  prejudice 
dates  from  certain  painful  reminiscences  associated  with 
the  arithmetic  of  our  school-days. 

Under  the  heading  "  Examples "  I  have  collected 
laboratory  measurements,  well-known  formulae,  practical 
problems  and  exercises  to  illustrate  the  text  immediately 
preceding.  A  few  of  the  problems  are  abstract  exercises  in 
pure  mathematics,  old  friends,  which  have  run  through 
dozens  of  text-books.  The  greater  number,  however,  are 
based  upon  measurements,  etc.,  recorded  in  papers  in  the 
current  science  journals  (Continental,  American  or  British) 
and  are  used  in  this  connection  for  the  first  time. 

It  can  serve  no  useful  purpose  to  disguise  the  fact  that  a 
certain  amount  of  drilling,  nay,  even  of  drudgery,  is  neces- 
sary in  some  stages,  if  mathematics  is  to  be  of  real  use  as 
a  working  tool,  and  not  employed  simply  for  quoting  the 
results  of  others.  The  proper  thing,  obviously,  is  to  make 
the  beginner  feel  that  he  is  gaining  strength  and  power 
during  the  drilling.  In  order  to  guide  the  student  along 
the  right  path,  hints  and  explanations  have  been  appended 
to  those  exercises  which  have  been  found  to  present  any 
difficulty.  The  subject-matter  contains  no  difficulty  which 
has  not  been  mastered  by  beginners  of  average  ability  with- 
out the  help  of  a  teacher. 

The  student  of  this  work  is  supposed  to  possess  a  work- 
ing knowledge  of  elementary  algebra  so  far  as  to  be  able  to 
solve  a  set  of  simple  simultaneous  equations,  and  to  know 
the   meaning  of  a   few   trigonometrical  formulae.      If   any 


INTRODUCTION.  xxi 

difficulty  should  arise  on  this  head,  it  is  very  possible  that 
the  appendix  will  contain  what  is  required  on  the  subject. 
I  have,  indeed,  every  reason  to  suppose  that  beginners  in  the 
study  of  Higher  Mathematics  most  frequently  find  their 
ideas  on  the  questions  discussed  in  §§  10,  11,  and  the  appen- 
dix, have  grown  so  rusty  with  neglect  as  to  require  refur- 
bishing. 

I  have  also  assumed  that  the  reader  is  acquainted  with 
the  elementary  principles  of  chemistry  and  physics.  Should 
any  illustration  involve  some  phenomenon  with  which  he 
is  not  acquainted,  there  are  two  remedies — to  skip  it,  or  to 
look  up  some  text-book.  There  is  no  special  reason  why  the 
student  should  waste  time  with  illustrations  in  which  he  has 
no  interest. 

It  will  be  found  necessary  to  procure  a  set  of  mathe- 
matical tables  containing  the  common  logarithms  of  numbers 
and  numerical  values  of  the  natural  trigonometrical  ratios. 
Such  sets  can  be  purchased  from  a  penny  upwards.  The 
other  numerical  tables  required  for  reference  in  Higher 
Mathematics  are  reproduced  in  Appendix  II. 


HIGHEK  MATHEMATICS 


FOR 


STUDENTS  OF  CHEMISTRY  AND  PHYSICS 


CHAPTER  I. 

THE  DIFFERENTIAL  CALCULUS. 

M  The  philosopher  may  be  delighted  with  the  extent  of  his  views,  the 
artificer  with  the  readiness  of  his  hands,  but  let  the  one  remember 
that  without  mechanical  performance,  refined  speculation  is  an 
empty  dream,  and  the  other  that  without  theoretical  reasoning, 
dexterity  is  little  more  than  brute  instinct." — S.  Johnson. 

§  1.  On  the  Nature  of  Mathematical  Reasoning. 

Herbert  Spencer  has  defined  a  law  of  Nature  as  a  proposition 
stating  that  a  certain  uniformity  has  been  observed  in  the  relations 
between  certain  phenomena.  In  this  sense  a  law  of  Nature  ex- 
presses a  mathematical  relation  between  the  phenomena  under 
consideration.  Every  physical  law,  therefore,  can  be  represented 
in  the  form  of  a  mathematical  equation.  One  of  the  chief  objects 
of  scientific  investigation  is  to  find  out  how  one  thing  depends  on 
another,  and  to  express  this  relationship  in  the  form  of  a  mathe- 
matical equation — symbolic  or  otherwise — is  the  experimenter's 
ideal  goal.1 

There  is  in  some  minds  an  erroneous  notion  that  the  methods 
of  higher  mathematics  are  prohibitively  difficult.  Any  difficulty 
that  might  arise  is  rather  due  to   the   complicated  nature  of  the 


lfThus  M.  Berthelot,  in  the  preface  to  his  celebrated  Essai  de  MScanique  Ghimique 
fond&e  sur  la  thermochemie  of  1879,  described  his  work  as  an  attempt  to  base  chemistry 
wholly  on  those  mechanical  principles  which  prevail  in  various  branches  of  physical 
science.  E.  Kant,  in  the  preface  to  his  Metaphysischen  Anfangsgrilnden  der  Natur- 
wissenscliaft,  has  said  that  in  every  department  of  physical  science  there  is  only  so 
much  science,  properly  so  called,  as  there  is  mathematics.  As  a  consequence,  he 
denied  to  chemistry  the  name  "science  ".  But  there  was  no  "Journal  of  Physical 
Chemistry"  in  his  time  (1786). 


4  HIGHER  MATHEMATICS.  §  1. 

phenomena  alone.  A.  Comte  has  said  in  his  Philosophie  Positive, 
**  our  feeble  minds  can  no  longer  trace  the  logical  consequences  of 
the  laws  of  natural  phenomena  whenever  we  attempt  to  simul- 
taneously include  more  than  two  or  three  essential  factors  V  In 
consequence  it  is  generally  found  expedient  to  introduce  "  simplifying 
assumptions  "  into  the  mathematical  analysis.  For  example,  in 
the  theory  of  solutions  we  pretend  that  the  dissolved  substance 
behaves  as  if  it  were  an  indifferent  gas.  The  kinetic  theory  of 
gases,  thermodynamics,  and  other  branches  of  applied  mathematics 
are  full  of  such  assumptions. 

By  no  process  of  sound  reasoning  can  a  conclusion  drawn  from 
limited  data  have  more  than  a  limited  application.  Even  when 
the  comparison  between  the  observed  and  calculated  results  is 
considered  satisfactory,  the  errors  of  observation  may  quite  obscure 
the  imperfections  of  formulae  based  on  incomplete  or  simplified 
premises.  Given  a  sufficient  number  of  "  if's,"  there  is  no  end  to 
the  weaving  of  "  cobwebs  of  learning  admirable  for  the  fineness 
of  thread  and  work,  but  of  no  substance  or  profit  ".2  The  only 
safeguard  is  to  compare  the  deductions  of  mathematics  with  ob- 
servation and  experiment  "  for  the  very  simple  reason  that  they 
are  only  deductions,  and  the  premises  from  which  they  are  made 
may  be  inaccurate  or  incomplete.  We  must  remember  that  we 
cannot  get  more  out  of  the  mathematical  mill  than  we  put  into  it> 
though  we  may  get  it  in  a  form  infinitely  more  useful  for  our 
purpose."  3 

The  first  clause  of  this  last  sentence  is  often  quoted  in  a 
parrot-like  way  as  an  objection  to  mathematics.  Nothing  but 
real  ignorance  as  to  the  nature  of  mathematical  reasoning  could 
give  rise  to  such  a  thought.  No  process  of  sound  reasoning 
can  establish  a  result  not  contained  in  the  premises.  It  is 
admitted  on  all  sides  that  any  demonstration  is  vicious  if  it 
contains   in   the  conclusion    anything    more    than    was  assumed 


*I  believe  that  this  is  the  key  to  the  interpretation  of  Comte's  strange  remarks  : 
"  Every  attempt  to  employ  mathematical  methods  in  the  study  of  chemical  questions 
must  be  considered  profoundly  irrational  and  contrary  to  the  spirit  of  chemistry. . . . 
If  mathematical  analysis  should  ever  hold  a  prominent  place  in  chemistry — an  aber- 
ration which  is  happily  almost  impossible — it  would  occasion  a  rapid  and  widespread 
degeneration  of  that  science." — Philosophie  Positive,  1830. 

2  F.  Bacon's  The  Advancement  of  Learning,  Oxford  edit.,  32,  1869. 

?  J.  Hopkinson's  James  Forrest  Lecture,  1§94. 


§  1.  THE  DIFFERENTIAL  CALCULUS.  5 

in  the  premises.1  Why  then  is  mathematics  singled  out  and 
condemned  for  possessing  the  essential  attribute  of  all  sound 
reasoning  ? 

Logic  and  mathematics  are  both  mere  tools  by  which  "  the 
decisions  of  the  mind  are  worked  out  with  accuracy,"  but  both 
must  be  directed  by  the  mind.  I  do  not  know  if  it  is  any  easier  to 
see  a  fallacy  in  the  assertion  that  "  when  the  sun  shines  it  is 
day;  the  sun  always  shines,  therefore  it  is  always  day,"  than  in 
the  statement  that  since  (-§-  3)2  =  (f  -  2)2,  we  get,  on  extracting 
roots,  f  -  3  =  J  -  2  ;  or  3  =  2.  We  must  possess  a  clear  conception 
of  any  physical  process  before  we  can  attempt  to  apply  mathe- 
matical methods  ;  mathematics  has  no  symbols  for  confused  ideas. 

It  has  been  said  that  no  science  is  established  on  a  firm  basis 
unless  its  generalizations  can  be  expressed  in  terms  of  number,  and 
it  is  the  special  province  of  mathematics  to  assist  the  investigator 
in  finding  numerical  relations  between  phenomena.  After  experi- 
ment, then  mathematics.  While  a  science  is  in  the  experimental 
or  observational  stage,  there  is  little  scope  for  discerning  numerical 
relations.  It  is  only  after  the  different  workers  have  "  collected 
data "  that  the  mathematician  is  able  to  deduce  the  required 
generalization.  Thus  a  Maxwell  followed  Faraday,  and  a  Newton 
completed  Kepler. 

It  must  not  be  supposed,  however,  that  these  remarks  are 
intended  to  imply  that  a  law  of  Nature  has  ever  been  represented 
by  a  mathematical  expression  with  perfect  exactness.  In  the  best 
of  generalizations,  hypothetical  conditions  invariably  replace  the 
complex  state  of  things  which  actually  obtains  in  Nature. 

Most,  if  not  all,  the  formulae  of  physics  and  chemistry  are  in 
the  earlier  stages  of  a  process  of  evolution.  For  example,  some 
exact  experiments  by  Forbes,  and  by  Tait,  indicate  that  Fourier's 
formula  for  the  conduction  of  heat  gives  somewhat  discordant 
results  on  account  of  the  inexact  simplifying  assumption:  "the 
quantity  of  heat  passing  along  a  given  line  is  proportional  to  the 
rate  of   change  of   temperature";   Weber  has   pointed  out  that 


1  Inductive  reasoning  is,  of  course,  good  guessing,  not  sound  reasoning,  but  the 
finest  results  in  science  have  been  obtained  in  this  way.  Calling  the  guess  a  "  working 
hypothesis,"  its  consequences  are  tested  Dy  experiment  in  every  conceivable  way.  For 
example,  the  brilliant  work  of  Fresnel  was  the  sequel  of  Young's  undulatory  theory 
of  light,  and  Hertz's  finest  work  was  suggested  by  Maxwell's  electro-magnetic  theories. 


6  HIGHER  MATHEMATICS.  §  2. 

Fick's  equation  for  the  diffusion  of  salts  in  solution  must  be 
modified  to  allow  for  the  decreasing  diffusivity  of  the  salt  with 
increasing  concentration  ;  and  finally,  van  der  Waals,  Clausius, 
Bankine,  Sarrau,  etc.,  have  attempted  to  correct  the  simple  gas 
equation:  pv  =  BT,  by  making  certain  assumptions  as  to  the 
internal  structure  of  the  gas. 

There  is  a  prevailing  impression  that  once  a  mathematical 
formula  has  been  theoretically  deduced,  the  law,  embodied  in 
the  formula,  has  been  sufficiently  demonstrated,  provided  the 
differences  between  the  "  calculated  "  and  the  "  observed  "  results 
fall  within  the  limits  of  experimental  error.  The  important  point, 
already  emphasized,  is  quite  overlooked,  namely,  that  any  discrep- 
ancy between  theory  and  fact  is  masked  by  errors  of  observation. 
With  improved  instruments,  and  better  methods  of  measurement, 
more  accurate  data  are  from  time  to  time  available.  The  errors  of 
observation  being  thus  reduced,  the  approximate  nature  of  the 
formulae  becomes  more  and  more  apparent.  Ultimately,  the  dis- 
crepancy between  theory  and  fact  becomes  too  great  to  be  ignored- 
It  is  then  necessary  to  "go  over  the  fundamentals  ".  New  formulae 
must  be  obtained  embodying  less  of  hypothesis,  more  of  fact.  Thus, 
from  the  first  bold  guess  of  an  original  mind,  succeeding  genera- 
tions progress  step  by  step  towards  a  comprehensive  and  a  complete 
formulation  of  the  several  laws  of  Nature. 

§  2.  The  Differential  Coefficient. 

Heracleitos  has  said  that  "  everything  is  in  motion,"  and  daily 
experience  teaches  us  that  changes  are  continually  taking  place  in 
the  properties  of  bodies  around  us.  Change  of  position,  change 
of  motion,  of  temperature,  volume,  and  chemical  composition 
are  but  a  few  of  the  myriad  changes  associated  with  bodies  in 
general. 

Higher  mathematics,  in  general,  deals  with  magnitudes  which 
change  in  a  continuous  manner.  In  order  to  render  such  a  process 
susceptible  to  mathematical  treatment,  the  magnitude  is  supposed 
to  change  during  a  series  of  very  short  intervals  of  time.  The 
shorter  the  interval  the  more  uniform  the  process.  This  conception 
is  of  fundamental  importance.  To  illustrate,  let  us  consider  the 
chemical  reaction  denoted  by  the  equation : 

Cane  sugar  — >  Invert  sugar. 


§  2.         THE  DIFFERENTIAL  CALCULUS.  7 

The  velocity  of  the  reaction,  or  the  amount 1  of  cane  sugar  trans- 
formed in  unit  time,  will  be 

Velocity  of  chemical  action  =  Amounmt  of  s"bsufcance  Produced  (1) 

Time  of  observation  v  ' 

This  expression  only  determines  the  average  velocity,  V,  of  the  re- 
action during  the  time  of  observation.  If  we  let  xx  denote  the 
amount  of  substance  present  at  the  time,  tv  when  the  observation 
commences,  and  x2  the  amount  present  at  the  time  t2,  the  average 
velocity  of  the  reaction  will  be 

y  =  xi  ~x%  •    ,-.  y=  _    .  (2) 

t2  —   i\  ot 

where  Sx  and  U  respectively  denote  differences  xx  -  x2i  and  t2  -  tr 
As  a  matter  of  fact  the  reaction  progresses  more  and  more  slowly 
as  time  goes  on.  Of  course,  if  sixty  grams  of  invert  sugar  were 
produced  at  the  end  of  one  minute,  and  the  velocity  of  the  reaction 
was  quite  uniform  during  the  time  of  observation ,  it  follows  that 
one  gram  of  invert  sugar  would  be  produced  every  second.  /We  ^) 
understand  the  mean  or  average  velocity  of  a  reaction  in  any  / 
given  interval  of  time,  to  be  the  amount  of  substance  which  would 
be  formed  in  unit  time  if  the  velocity  remained  uniform  and  con- 
stant throughout  the  interval  in  question.  But  the  velocity  is  not 
uniform — it  seldom  is  in  natural  changes.  In  consequence,  the 
average  velocity,  sixty  grams  per  minute,  does  not  represent  the 
rate  of  formation  of  invert  sugar  during  any  particular  second,  but 
simply  the  fact  observed,  namely,  the  mean  rate  of  formation  of 
invert  sugar  during  the  time  of  observation/^ 

Again,  if  we  measured  the  velocity  of  {he  reaction  during  one 
second,  and  found  that  half  a  gram  of  invert  sugar  was  formed  in 
that  interval  of  time,  we  could  only  say  that  invert  sugar  was  pro- 
duced at  the  rate  of  half  a  gram  per  second  during  the  time  of  ob- 
servation. But  in  that  case,  the  average  velocity  would  more 
accurately  represent  the  actual  velocity  during  the  time  of  obser- 
vation, because  there  is  less  time  for  the  velocity  of  the  reaction  to 
vary  during  one  than  during  sixty  seconds. 


1  By  "  amount  of  substance  "  we  understand  "  number  of  gram-molecules  "  per 
litre  of  solution.  "One  gram-molecule  "  is  the  molecular  weight  of  the  substance 
expressed  in  grams.  E.g.,  18  grms.  of  water  is  1  gram-molecule;  27  grms.  is  1*5 
gram-molecules;  36  grms.  is  2  gram-molecules,  etc.  We  use  the  terms  "amount," 
"quantity,"   "concentration,"  and  "active  mass"  synonymously. 


< 


8  HIGHER  MATHEMATICS.  §  2. 

/  By  shortening  the  time  of  observation   the  average  velocity 

approaches  more  and  more  rfearly  to  the  actual  velocity  of  the  re- 
action during  the  whole  time  of  observation.  In  order  to  measure 
the  velocity  of  the  reaction  at  any  instant  of  time,  it  would  be 
necessary  to  measure  the  amount  of  substance  formed  during  an 
infinitely  short  instant  of  time.  But  any  measurement  we  can 
possibly  make  must  occupy  some  time,  and  consequently  the 
velocity  of  the  particle'has  time  to  alter  while  the  measurement  is 
in  progress.  It  is  thus  a  physical  impossibility  to  measure  the 
velocity,  at  any  instant ;  but,  in  spite  of  this  fact,  it  is  frequently 
necessary  to  reason  about  this  ideal  condition. 

We  therefore  understand  by  velocity  at  any  instant,  the  mean 
or  average  velocity  during  a  very  small  interval  of  time,  with  the 
proviso  that  we  can  get  as  near  as  we  please  to  the  actual  velocity 
at  any  instant  by  taking  the  time  of  observation  sufficiently  small. 
An  instantaneous  velocity  is  represented  by  the  symbol 

Jf  =  F' ^ 

where  dx  is  the  symbol  used  by  mathematicians  to  represent  an 
infinitely  small  amount  of  something  (in  the  above  illustration,  invert 
sugar),  and  dt  a  correspondingly  short  interval  of  time.  Hence 
it  follows  that  neither  of  these  symbols  per  se  is  of  any  practical 
value,  but  their  quotient  stands  for  a  perfectly  definite  conception, 
namely,  the  rate  of  chemical  transformation  measured  during  an 
interval  of  time  so  small  that  all  possibility  of  error  due  to  vari- 
ation of  speed  is  eliminated. 

Numerical  Illustration. — The  rate  of  conversion  of  acetochloranilide 
into  p-chloracetanilide,  just  exactly  four  minutes  after  the  reaction  had  started, 
was  found  to  be  4-42  gram-molecules  per  minute.  The  "  time  of  observation  " 
was  infinitely  small.  When  the  measurement  occupied  the  whole  four 
minutes,  the  average  velocity  was  found  to  be  8*87  gram-molecules  per 
minute ;  when  the  measurement  occupied  two  minutes,  the  average  velocity 
was  5-90  units  per  minute  ;  and  finally,  when  the  time  of  observation  occupied 
one  minute,  the  reaction  apparently  progressed  at  the  rate  of  4*70  units  per 
minute.  Obviously  then  we  approximate  more  closely  to  the  actual  velocity, 
4-42  gram-molecules  per  minute,  the  smaller  the  time  of  observation. 

The  idea  of  an  instantaneous  velocity,  measured  during  an 
interval  of  time  so  small  that  no  perceptible  error  can  affect  the 
result,  is  constantly  recurring  in  physical  problems,  and  we  shall 
soon   see   that  the   so-called    "  methods  of  differentiation  "   will 


§  2.  THE  DIFFERENTIAL  CALCULUS.  9? 

actually  enable  us  to  find  the  velocity  or  rate  of  change  under  these 
conditions.  The  quotient  dx[dtjs_kp.o\yn  as  the  differential  co- 
efficient of  x  with  respect  to  t.  The  value  of  x  obviously  depends 
upon  what  value  is  assigned  to  t,  the  time  of  observation ;  for  this 
reason,  x  is  called  the  dependent  variable,  t  the  independent 
variable.  The  differential  coefficient  is  the  only  true  measure  of 
a  velocity  at  any  instant  of  time.  Our  "  independent  variable  " 
is  sometimes  called  the  principal  variable;  our  "dependent 
variable  "  the  subsidiary  variable. 

Just  as  the  idea  of  the  velocity  of  a  chemical  reaction  represents 
the  amount  of  substance  formed  in  a  given  time,  so  the  velocity  of 
any  motion  can  be  expressed  in  terms  of  the  differential  coefficient 
of  a  distance  with  respect  to  time,  be  the  motion  that  of  a  train, 
tramcar,  bullet,  sound-wave,  water  in  a  pipe,  or  of  an  electric 
current. 

The  term  "  velocity  "  not  only  includes  the  rate  of  motion,  but 
also  the  direction  of  the  motion.  If  we  agree  to  represent  the 
velocity  of  a  train  travelling  southwards  to  London,  positive,  a 
train  going  northwards  to  Aberdeen  would  be  travelling  with  a 
negative  velocity.  Again,  we  may  conventionally  agree  to  consider 
the  rate  of  formation  of  invert  sugar  from  cane  sugar  as  a  positive 
velocity,  the  rate  of  decomposition  of  cane  sugar  into  invert  sugar 
as  a  negative  velocity. 

It  is  not  necessary,  for  our  present  purpose,  to  enter  into  re- 
fined distinctions  between  rate,  speed,  and  velocity.  Velocity  is 
of  course  directed  speed.  I  shall  use  the  three  terms  synonym- 
ously. 

The  concept  velocity  need  not  be  associated  with  bodies. 
Every  one  is  familiar  with  such  terms  as  "  the  velocity  of  light," 
"the  velocity  of  sound,"  and  "  the  velocity  of  an  explosion- wave  ". 
The  chemical  student  will  soon  adapt  the  idea  to  such  phrases  as, 
"the  velocity  of  chemical  action,"  "  the  speed  of  catalysis,"  "the 
rate  of  dissociation,"  "the  velocity  of  diffusion,"  "the  rate  of 
evaporation,"  etc.  It  requires  no  great  mental  effort  to  extend 
the  notion  still  further.  If  a  quantity  of  heat  is  added  to  a  sub- 
stance at  a  uniform  rate,  the  quantity  of  heat,  Q,  added  per  degree 
rise  of  temperature,  0,  corresponds  exactly  with  the  idea  of  a 
distance  traversed  per  second  of  time.  Specific  heat,  therefore, 
may  be  represented  by  the  differential  coefficient  dQ/dO.  Simi- 
larly, the  increase  in  volume  per  degree  rise  of  temperature  is 


,10  HIGHER  MATHEMATICS.  §  4. 

represented  by  the  differential  coefficient  dv/dO  ;  the  decrease  in 
volume  per  unit  of  pressure,  p,  is  represented  by  the  ratio  -  dvjdp, 
where  the  negative  sign  signifies  that  the  volume  decreases  with 
increase  of  pressure.  In  these  examples,  it  has  been  assumed  that 
unit  mass  or  unit  volume  of  substance  is  operated  upon,  and  there- 
fore the  differential  coefficients  respectively  represent  specific  heat, 
coefficient  of  expansion,  and  coefficient  of  compressibility. 

From  these  and  similar  illustrations  which  will  occur  to  the 
reader,  it  will  be  evident  that  the  conception  called  by  mathe- 
maticians "the  differential  coefficient"  is  not  new.  Every  one 
consciously  or  unconsciously  uses  it  whenever  a  "  rate,"  "  speed," 
or  a  "  velocity  "  is  in  question. 

§  3.  Differentials. 

It  is  sometimes  convenient  to  regard  dx  and  dt,  or  more  generally 
dx  and  dy,  as  very  small  quantities  which  determine  the  course  of 
any  particular  process  under  investigation.  These  small  magni- 
tudes are  called  differentials  or  infinitesimals.  Some  one  has 
defined  differentials  as  small  quantities  "verging  on  nothing". 
Differentials  may  be  treated  like  ordinary  algebraic  magnitudes. 
The  quantity  of  invert  sugar  formed  in  the  time  dt  is  represented 
by  the  differential  dx.  Hence  from  (3),  if  dx/dt  =  V,  we  may 
write  in  the  language  of  differentials 

dx  =  V.  dt. 

I  suppose  that  the  beginner  has  only  built  up  a  vague  idea  of 
the  magnitude  of  differentials  or  infinitesimals.  They  seem  at 
once  to  exist  and  not  to  exist.  I  will  now  try  to  make  the  concept 
more  clearly  defined. 

§  i.  Orders  of  Magnitude. 

If  a  small  number  n  be  divided  into  a  million  parts,  each  part,1 
n  x  10  ~ 6  is  so  very  small  that  it  may  for  all  practical  purposes  be 
neglected  in  comparison  with  n.  If  we  agree  to  call  n  a  magnitude 
of  the  first  order,  the  quantity  w  x  10"6  is  a  magnitude  of  the 
second  order.     If  one  of  these  parts  be  again  subdivided  into  a 

1  Note  104  =  unity  followed  by  four  cyphers,  or  10,000.  10  -  4  is  a  decimal  point 
followed  by  three  cyphers  and  unity,  or  10  ~  4  =  101000  =  0*0001.  This  notation  is  in 
general  use. 


§  4.  THE  DIFFERENTIAL  CALCULUS.  11 

million  parts,  each  part,  n  x  10  ~  12,  is  extremely  small  when 
compared  with  n,  and  the  quantity  n  x  10  ~  12  is  a  magnitude  of 
the  third  order.  We  thus  obtain  a  series  of  magnitudes  of  the 
first,  second,  and  higher  orders, 


n, 


l.OOOOOO'    1000000.000000' 

each  one  of  which  is  negligibly  small  in  comparison  with  those 
which  precede  it,  and  very  large  relative  to  those  which  follow. 

This  idea  is  of  great  practical  use  in  the  reduction  of  intricate 
expressions  to  a  simpler  form  more  easily  manipulated.  It  is 
usual  to  reject  magnitudes  of  a  higher  order  than  those  under 
investigation  when  the  resulting  error  is  so  small  that  it  is  out- 
side the  limits  of  the  "errors  of  observation"  peculiar  to  that 
method  of  investigation. 

Having  selected  our  unit  of  smallness,  we  decide  what  part  of 
this  is  going  to  be  regarded  as  a  small  quantity  of  the  first  order. 
Small  quantities  of  the  second. order  then  bear  the  same  ratio  to 
magnitudes  of  the  first  order,  as  the  latter  bear  to  the  unit  of 
measurement.  In  the  "theory  of  the  moon,"  for  example,  we 
are  told  that  y1^  is  reckoned  small  in  comparison  with  unity  ;  (x1^)2 
is  a  small  magnitude  of  the  second  order ;  (yV)3  of  the  third  order, 
etc.  Calculations  have  been  made  up  to  the  sixth  or  seventh 
orders  of  small  quantities. 

In  order  to  prevent  any  misconception  it  might  be  pointed  out 
that  "great"  and  "small"  in  mathematics,  like  "hot"  and 
"cold"  in  physics,  are  purely  relative  terms.  The  astronomer 
in  calculating  interstellar  distances  comprising  millions  of  miles 
takes  no  notice  of  a  few  thousand  miles  ;  while  the  physicist  dare 
not  neglect  distances  of  the  order  of  the  ten  thousandth  of  an  inch 
in  his  measurements  of  the  wave  length  of  light. 

A  term,  therefore,  is  not  to  be  rejected  simply  because  it  seems 
small  in  an  absolute  sense,  but  only  when  it  appears  small  in 
comparison  with  a  much  larger  magnitude,  and  when  an  exact 
determination  of  this  small  quantity  has  no  appreciable  effect  on 
the  magnitude  of  the  larger.  In  making  up  a  litre  of  normal 
oxalic  acid  solution,  the  weighing  of  the  63  grams  of  acid  required 
need  not  be  more  accurate  than  to  the  tenth  of  a  gram.  In  many 
forms  of  analytical  work,  however,  the  thousandth  of  a  gram  is  of 
fundamental  importance  ;  an  error  of  a  tenth  of  a  gram  would 
stultify  the  result. 


12  HIGHER  MATHEMATICS.  §  5. 

§  5.  Zero  and  Infinity. 

The  words  " infinitely  small"  were  used  in  the  second  para- 
graph. It  is,  of  course,  impossible  to  conceive  of  an  infinitely  small 
or  of  an  infinitely  great  magnitude,  for  if  it  were  possible  to  retain 
either  of  these  quantities  before  the  mind  for  a  moment,  it  would 
be  just  as  easy  to  think  of  a  smaller  or  a  greater  as  the  case  might 
be.  In  mathematical  thought  the  word  "  infinity  "  (written  :  go) 
signifies  the  properties  possessed  by  a  magnitude  greater  than  any 
finite  magnitude  that  can  be  named.  For  instance,  the  greater 
we  make  the  radius  of  a  circle,  the  more  approximately  does  the 
circumference  approach  a  straight  line,  until,  when  the  radius  is 
made  infinitely  great,  the  circumference  may,  without  committing 
any  sensible  error,  be  taken  to  represent  a  straight  line.  The  con- 
sequences of  the  above  definition  of  infinity  have  led  to  some  of 
the  most  important  results  of  higher  mathematics.  To  sum- 
marize, infinity  represents  neither  the  magnitude  nor  the  value 
of  any  particular  quantity.  The  term  "infinity"  is  simply  an 
abbreviation  for  the  property  of  growing  large  without  limit.  E.g., 
"tan  90°  =  oo "  means  that  as  an  angle  approaches  90°,  its  tan- 
gent grows  indefinitely  large.  Now  for  the  opposite  of  greatness — 
smallness. 

In  mathematics  two  meanings  are  given  to  the  word  "  zero  ". 
The  ordinary  meaning  of  the  word  implies  the  total  absence  of 
magnitude  ;  we  shall  call  this  absolute  zero.  Nothing  remains 
when  the  thing  spoken  of  or  thought  about  is  taken  away.  If  four 
units  be  taken  from  four  units  absolutely  nothing  remains.  There 
is,  however,  another  meaning  to  be  attached  to  the  word  different 
from  the  destruction  of  the  thing  itself.  If  a  small  number  be 
divided  by  a  billion  we  get  a  small  fraction.  If  this  fraction  be 
raised  to  the  billionth  power  we  get  a  number  still  more  nearly 
equal  to  absolute  zero.  By  continuing  this  process  as  long  as  we 
please  we  continually  approach,  but  never  actually  reach,  the 
absolute  zero.  In  this  relative  sense,  zero — relative  zero — is 
defined  as  "an  infinitely  small "  or  "a  vanishingly  small  number," 
or  "a  number  smaller  than  any  assignable  fraction  of  unity". 
For  example,  we  might  consider  a  point  as  an  infinitely  small 
circle  or  an  infinitely  short  line.  To  put  these  ideas  tersely,  ab- 
solute zero  implies  that  the  thing  and  all  the  properties  are  absent 
relative  zero  implies   that  however  small  the   thing  may  be  its 


§  6.  THE  DIFFERENTIAL  CALCULUS.  13 

'property  of  growing  small  without  limit  is  alone  retained  in  the 
mind. 

Examples. — (1)  After  the  reader  has  verified  the  following  results  he  will 
understand  the  special  meaning  to  be  attached  to  the  zero  and  infinity  of 
mathematical  reasoning,  n  denotes  any  finite  number ;  and  "  ?  "  an  in- 
determinate magnitude,  that  is,  one  whose  exact  value  cannot  be  determined. 

00+00=00;  00-  00  =  ?;  wxO  =  0;  0x0  =  0;  n  x  00=00; 
0/0  =  ? ;  n{0  =  00 ;  0/n  =  0 ;    oo/O  =  00  ;  0/  00  =  0 ;  n\  00  =  0 ;    00/n  =  00. 

(2)  Let  y  =  1/(1  -  x)  and  put  x  =  1,  then  y  =  00 ;  if  x  <  1,  y  is  positive  ; 
and  when  x  >  1,  y  is  negative.1  Thus  a  variable  magnitude  may  change  its 
sign  when  it  becomes  infinite. 

If  the  reader  has  access  to  the  Transactions  of  the  Cambridge 
Philosophical  Society  (11.  145,  1871),  A.  de  Morgan's  paper  "  On 
Infinity,"  is  worth  reading  in  connection  with  this  subject. 

§  6.  Limiting  Values. 

I.  The  sum  of  an  infinite  number  of  terms  may  have  a  finite 
value.     Converting  J  into  a  decimal  fraction  we  obtain 
£  =  0*11111  .  . .  continued  to  infinity, 
i  =  tV  +  rhv  +  toV*  +  ...  to  infinity, 
that  is  to  say,  the  sum  of  an  infinite  number  of  terms  is  equal  to  1 
— a  finite  term  !     If  we  were  to  attempt  to  perform  this  summa- 
tion we  should  find  that  as  long  as  the  number  of  terms  is  finite 
we  could  never  actually  obtain  the  result  £.     We  should  be  ever 
getting  nearer  but  never  get  actually  there. 

If  we  omit  all  terms  after  the  first,  the  result  is  ^  less  than  i  ; 
if  we  omit  all  terms  after  the  third,  the  result  is  — 1  too  little ; 
and  if  we  omit  all  terms  after  the  sixth,  the  result  is  9, 000.000 
less  than  i,  that  is  to  say,  the  sum  of  these  terms  continually 
approaches  but  is  never  actually  equal  to  J  as  long  as  the  number 
of  terms  is  finite.  ^  is  then  said  to  be  the  limiting  value  of  the 
sum  of  this  series  of  terms. 

Again,  the  perimeter  of  a  polygon  inscribed  in  a  circle  is  less 
than  the  sum  of  the  arcs  of  the  circle,  i.e.,  less  than  the  circum- 


1  The  signs  of  inequality  are  as  follows:  "  ^  "  denotes  "is  not  equal  to"  ; 
">,"  "  is  greater  than  "  ;  "^>,"  "  is  not  greater  than "  ;  "<,"  " is  less  than  "  ; 
and  "<£,"  "  is  not  less  than  ".  For  "  =E  "  read  "  is  equivalent  to  "  or  "  is  identical 
with  ".  The  symbol  >—  has  been  used  in  place  of  the  phrase  "is  greater  than  or' 
equal  to,"  and  — <,  in  place  of  "is  equal  to  or  less  than". 


14 


HIGHER  MATHEMATICS. 


§6- 


ference  of  the  circle.  In  Fig.  1,  let  the  arcs  AaB,  BbC...be 
bisected  at  a,  b  . . .  Join  A  a,  aB,  Bb,  ... 
Although  the  perimeter  of  the  second  poly- 
gon is  greater  than  the  first,  it  is  still  less 
than  the  circumference  of  the  circle.  In 
a  similar  way,  if  the  arcs  of  this  second 
polygon  are  bisected,  we  get  a  third  poly- 
gon whose  perimeter  approaches  yet  nearer 
to  the  circumference  of  the  circle.  By 
continuing  this  process,  a  polygon  may  be 
obtained  as  nearly  equal  to  the  circum- 
ference of  a  circle  as  we  please.  The  circumference  of  the  circle  is 
thus  the  limiting  value  of  the  perimeter  of  an  inscribed  polygon, 
when  the  number  of  its  sides  is  increased  indefinitely. 

In  general,  when  a  variable  magnitude  x  continually  approaches 
nearer  and  nearer  to  a  constant  value  n  so  that  x  can  be  made  to 
differ  from  n  by  a  quantity  less  than  any  assignable  magnitude,  n 
is  said  to  be  the  limiting  value  of  x. 

From  page  8,  it  follows  that  dx/dt  is  the  limiting  value1  of 
Sx/St,  when  t  is  made  less  than  any  finite  quantity,  however  small. 
This  is  written,  for  brevity, 

dx 
dt 


Sx 


I 


in  words  "dx/dt  is  the  limiting  value  of  Sx/U  when  t  becomes 
zero  "  or  rather  relative  zero,  i.e.,  small  without  limit.  This  no- 
tation is  frequently  employed. 

The  sign  "  =  "  when  used  in  connection  with  differential  co- 
efficients does  not  mean  "is  equal  to,"  but  rather  "can  be  made 
as  nearly  equal  to  as  we  please  ".  We  could  replace  the  usual 
"  =  "  by  some  other  symbol,  say  "  =^,"  if  it  were  worth  while.2 

II.  The  value  of  a  limiting  ratio  depends  on  the  relation  be- 
tween the  two  variables.     Strictly  speaking,  the  limiting  value  of 


1  Although  differential  quotients  are,  in  this  work,  written  in  the  form  "dx/dt," 

cPxjdt2  .  . .  ,  the  student  in  working  through  the  examples  and  demonstrations,  should 

..     dx    d2x 
write  -=-.,    -ttjs. 

dt'    dP 

2  The  symbol  "  x  =  0  "  is  sometimes  used  for  the  phrase  "  as  x  approaches  zero  ". 

"lim  "      "lim  " 

=  0  '        M=0  '  or  "•£"  are  also  used  instead  of  our  "Lt^o,     meaning  "the 

limit  of .  . .  as  jc  approaches  zero", 


II 


The  former  method  is  used  to  economize  space. 


§6. 


THE  DIFFERENTIAL  CALCULUS. 


15 


the  ratio  Sx/St  has  the  form  g,  and  as  such  is  indeterminate — in- 
determinate, because  g-  may  have  any  numerical  value  we  please. 
It  is  not  difficult  to  see  this,  for  example,  g  =  0,  because  0x0  =  0; 
£  =  1,  because  0x1  =  0;  J  =  2,  because  0  x  2  =  0 ;  §  =  15, 
because  0  x  15  =  0 ;  §  =  999,999,  because  0  x  999,999  =  0,  etc. 

Example.— There  is  a  "hoary-headed"  puzzle  which  goes  like  this: 
Given  x  =  a ;  .\  x2  =  xa ;  .*.  x2  -  a2  =  xa  -  a? ;  .•.  (a;  -  a)  (x  +  a)  =  a(aj  -  a) ; 
.\  x  +  a  =  a;  .-.  2a  =  a  ;  .'.2  =  1.  Where  is  the  fallacy?  Ansr.  The  step 
(x  -  a)  (x  +  a)  =  a(x  -  a)  means  (x  +  a)  x  0  =  a  x  0,  i.e.,  no  times  a;  +  a  = 
no  times  a,  and  it  does  not  necessarily  follow  that  x  +  a  is  therefore  equal 
to  a. 

For  all  practical  purposes  the  differential  coefficient  dx/dt  is  to  be 
regarded  as  a  fraction  or  quotient.  The  quotient  dx/dt  may  also 
be  called  a  "  rate-measurer,"  because  it  determines  the  velocity  or 
rate  with  which  one  quantity  varies  when  an  extremely  small 
variation  is  given  to  the  other.  The  actual  value  of  the  ratio  dx/dt 
depends  on  the  relation  subsisting  between  x  and  t. 

Consider  the  following  three  illustrations  :  If  the  point  P  move 
on  the  circumference  of  the  circle  towards  a 
fixed  point  Q  (Fig.  2),  the  arc  x  will  diminish 
at  the  same  time  as  the  chord  y.  By 
bringing  the  point  P  sufficiently  near  to  Q, 
we  obtain  an  arc  and  its  chord  each  less 
than  any  given  line,  that  is,  the  arc  and 
the  chord  continually  approach  a  ratio  of 
equality.  Or,  the  limiting  value  of  the  ratio 
Sx/Sy  is  unity. 


Fig.  2. 


T  Sx 


~dy  ~ 


It  is  easy  to  show  this  numerically.  Let  us  start  with  an  angle 
of  60°  and  compare  the  length,  dx,  of  the  arc  with  the  length, 
dy,  of  the  corresponding  chord. 


Angle  at 
Centre. 

dx. 

dy. 

dx 
dy 

60° 

1-0472 

1-0000 

1-0472 

30° 

0-5236 

05176 

1-0115 

10° 

0-1745 

0-1743 

1-0013 

5° 

0-0873 

0-0872 

1-0003 

1° 

0-0175 

00175 

1-0000 

16 


HIGHER  MATHEMATICS. 


§6. 


A  chord  of  1°  does  not  differ  from  the  corresponding  arc  in  the  first 
four  decimal  places  ;  if  the  angle  is  1',  the  agreement  extends 
through  the  first  seven  decimal  places  ;  and  if  the  angle  be  1",  the 
agreement  extends  through  the  first  fifteen  decimals.  The  arc  and 
its  chord  thus  approach  a  ratio  of  equality. 

If  ABC  (Fig.  3)  be  any  right-angled  triangle  such  that  AB  =  BG  ; 
by  Pythagoras'  theorem  or  Euclid,  i.,  47,  and  vi.,  4, 

AB  :  AG  =  x  :  y  =  1  :  J2. 
If  a  line  DE,  moving  towards  A,  remain  parallel  to  BC,  this  pro- 
portion will  remain  constant  even  though  each  side  of  the  triangle 
ADE  is  made  less  than  any  assignable  magnitude,  however  small. 
That  is 

T  Sx  _  dx  _   1 

^  =  GBy~d^-J2' 

Let  ABC  be  a  triangle  inscribed  in  a  circle  (Fig.  4).     Draw  AD 
perpendicular  to  BC.     Then  by  Euclid,  vi.,  8 

BC  :  AC  =  AC  :  DC  =  x  :  y. 

If  A  approaches  C  until  the  chord  AC  becomes  indefinitely 
small,  DC  will  also  become  indefinitely  small.     The  above  propor- 


Fig.  3. 


Fig.  4. 


tion,  however,  remains.  When  the  ratio  BC  :  AC  becomes  in- 
finitely great,  the  ratio  of  AC  to  DC  will  also  become  infinitely 
great,  or 

T ,        Sx      dx 

It  therefore  follows  at  once  that  although  two  quantities  may 
become  infinitely  small  their  limiting  ratio  may  have  any  finite  or 
infinite  value  whatever. 

Example. — Point  out  the  error  in  the  following  deduction  :  "  If  A B  (Fig.  3) 
is  a  perpendicular  erected  upon  the  straight  line  BC,  and  C  is  any  point; 


§  7.  THE  DIFFERENCIAL  CALCULUS.  17 

upon  BG,  then  AG  is  greater  than  AB,  however  near  C  may  be  to  B,  and, 
therefore,  the  same  is  true  at  the  limit,  when  G  coincides  with  B".  Hint. — 
The  proper  way  to  put  it  is  to  say  that  AC  becomes  more  and  more  nearly 
equal  to  AB,  as  C  approaches  B,  etc. 

Two  quantities  are  generally  said  to  be  equal  when  their 
difference  is  zero.  This  does  not  hold  when  dealing  with  differ- 
entials. The  difference  between  two  infinitely  small  quantities 
may  be  zero  and  yet  the  quantities  are  not  equal.  Infinitesimals 
can  only  be  regarded  as  equal  when  their  ratio  is  unity. 

§  7.  The  Differential  Coefficient  of  a  Differential  Coefficient. 

Velocity  itself  is  generally  changing.  The  velocity  of  a  falling 
stone  gradually  increases  during  its  descent,  while,  if  a  stone  is 
projected  upwards,  its  velocity  gradually  decreases  during  its  ascent. 
Instead  of  using  the  awkward  term  "  the  velocity  of  a  velocity," 
the  word  acceleration  is  employed.  If  the  velocity  is  increasing 
at  a  uniform  rate,  the  acceleration,  F,  or  rate  of  change  of  velocity, 
or  rate  of  change  of  motion,  is  evidently 

Increase  of  velocity      -&       V, 2  —   V1        SV 
Acceleration  =  -     -^^ 1  F  -    ^  _  ^    -  y,  (1) 

where  V1  and  V2  respectively  denote  the  velocities  at  the  beginning, 
tv  and  end,  t2,  of  the  interval  of  time  under  consideration  ;  and  SV 
denotes  the  small  change  of  velocity  during  the  interval  of  time  St. 
In  order  to  fix  these  ideas  we  shall  consider  a  familiar  ex- 
periment, namely,  that  of  a  stone  falling  from  a  vertical  height. 
Observation  shows  that  if  the  stone  falls  from  a  position  of  rest, 
its  velocity,  at  the  end  of  1,  2,  3,  4,  and  5  seconds  is 

32,  64,  96,  128,  160, 
feet  per  second  respectively.  In  other  words,  the  velocity  of  the 
descending  stone  is  increasing  from  moment  to  moment.  The 
above  reasoning  still  holds  good.  Let  ds  denote  the  distance  tra- 
versed during  the  infinitely  short  interval  of  time  dt.  The  velocity 
of  descent,  at  any  instant,  is  evidently 

ds 

Tt  =  v- m: 

Next  consider  the  rate  at  which  the  velocity  changes  from  one 
moment  to  another.  This  change  is  obviously  the  limit  of  the 
ratio  SV/St,  when  St  is  zero.     In  other  words 

-ft  -  JR         •        ,        .        .        (3) 

B 


18  HIGHER  MATHEMATICS.  §  7. 


Substituting  for  7,  we  obtain  the  second  differential  coefficient 

■  ,  f© 

which  is  more  conveniently  written 

d2s       „  /  ds2\ 

it>-F>{«°*p-w}  ■    -  •  ■■•  m 

This  expression  represents  the  rate  at  which  the  velocity  is  increas- 
ing at  any  instant  of  time.  In  this  particular  example  the  acceler- 
ation is  due  to  the  earth's  gravitational  force,  and  g  is  usually 
written  instead  of  F. 

The  ratio  d2x/dt2  is  called  the  second  differential  coefficient 
of  x  with  respect  to  t.  Just  as  the  first  differential  coefficient  of  x 
with  respect  to  t  signifies  a  velocity,  so  does  the  second  differential 
coefficient  of  x  with  respect  to  t  denote  an  acceleration. 

The  velocity  of  most  chemical  reactions  gradually  diminishes  as 
time  goes  on.  Thus,  the  rate  of  transformation  of  cane  sugar  into 
invert  sugar,  after  the  elapse  of  0,  30,  60,  90,  and  130  minutes 
was  found  to  be  represented  by  the  numbers l 

15-4,  13-7,  12-4,  114,  9-7, 
respectively. 

If  the  velocity  of  a  body  increases,  the  velocity  gained  per 
second  is  called  its  acceleration  ;  while  if  its  velocity  decreases,  its 
acceleration  is  really  a  retardation.  Mathematicians  often  prefix 
a  negative  sign  to  show  that  the  velocity  is  diminishing.  Thus, 
the  rate  at  which  the  velocity  of  the  chemical  reaction  changes  is, 
with  the  above  notation, 

d2x 

F  =  ~dt2 (5) 

In  our  two  illustrations,  the  stone  had  an  acceleration  of  32 
units  (feet  per  second)  per  second ;  the  chemical  reaction  had  an 
acceleration  of  -  0*00073  units  per  second,  or  of  -  0*044  units  per 
minute.     See  also  page  155. 

In  a  similar  way  it  can  be  shown  that  the  third  differential 
coefficient  represents  the  rate  of  change  of  acceleration  from  moment 
to  moment ;  and  so  on  for  the  higher  differential  coefficients 
dnx/dtn,  which  are  seldom,  if  ever,  used  in  practical  work.  A  few 
words  on  notation. 


Multiplied  by  103. 


§  9.  THE  DIFFERENTIAL  CALCULUS.  19 

§  8.  Notation. 

It  is  perhaps  needless  to  remark  that  the  letters  S,  A,  d,  d2,  . .  .* 
do  not  represent  algebraic  magnitudes.  They  cannot  be  dis- 
sociated from  the  appended  x  and  t.  These  letters  mean  nothing 
more  than  that  x  and  t  have  been  taken  small  enough  to  satisfy 
the  preceding  definitions. 

Some  mathematicians  reserve  the  symbols  Sx,  St,  Ax,  A  t, . .  . 
for  small  finite  quantities  ;  dx,  dt  . . .  have  no  meaning  per  se.  As 
a  matter  of  fact  the  symbols  dx,  dt . . .  are  constantly  used  in  place 

of  Sx,  St,  . . .,  or  Ax,  At.  ...  In  the  ratio  ~r-,  -77 is  the  symbol  of 

an  operation  performed  on  x,  as  much  as  the  symbols  "  +  "  or 

"/"  denote  the  operation  of  division.      In  the  present  case  the 

Sx 
operation  has  been  to  find  the  limiting  value  of  the  ratio  -^  when 

St  is  made  smaller  and  smaller  without  limit ;  but  we  constantly 
find  that  dx/dt  is  used  when  Sx/St  is  intended.  For  convenience, 
D  is  sometimes  used  as  a  symbol  for  the  operation  in  place  of  d/dx. 
The  notation  we  are  using  is  due  to  Leibnitz;2  Newton,  the  dis- 
coverer of  this  calculus,  superscribed  a  small  dot  over  the  de- 
pendent variable  for  the  first  differential  coefficient,  two  dots  for 
the  second,  thus  x,  x  .  . . 

dy  d 

In  special  cases,  besides  ~j-  and  y,  we  may  have  j-(y),  dyx, 

d2y         /  d  \2 
xy,  xv  x'  . . .  for  the  first  differential  coefficient ;  -j-2,  y,  i-rj  y, 

x2,  x"  . . .  for  the  second  differential  coefficient ;  and  so  on  for  the 
higher  coefficients,  or  derivatives  as  they  are  sometimes  called. 

The  operation  of  finding  the  value  of  the  differential  coefficients 
of  any  expression  is  called  differentiation.  The  differential 
calculus  is  that  branch  of  mathematics  which  deals  with  these 
operations. 

§  9.  Functions. 

If  the  pressure  to  which  a  gas  is  subject  be  altered,  it  is  known 
that  the  volume  of  the  gas  changes  in  a  proportional  way.     The 


1  For  ". . . "  read  "etc."  or  "and  so  on  ". 

2  The  history  of  the  subject  is  somewhat  sensational.    See  B.  Williamson's  article  in 
the  Encyclopaedia  Britannica,  Art.  "Infinitesimal  Calculus". 

B* 


20  HIGHER  MATHEMATICS.  §  9. 

two  magnitudes,  pressure  p  and  volume  v,  are  interdependent. 

Any  variation  of  the  one  is  followed  by  a  corresponding  variation 

of  the  other.     In  mathematical  language  this  idea  is  included  in 

the  word  "function  " ;  v  is  said  to  be  a  function  of  p.     The  two 

related  magnitudes  are  called  variables.     Any  magnitude  which 

remains  invariable  during  a  given  operation  is  called  a  constant. 

In  expressing  Boyle's  law  for  perfect  gases  we  write  this  idea 

thus: 

Dependent  variable  =  /  (independent  variable), 

or  v  =  f(p), 

meaning  that  "v  is  some  function  of  p".  There  is,  however,  no 
particular  reason  why  p  was  chosen  as  the  independent  variable. 
The  choice  of  the  dependent  variable  depends  on  the  conditions  of 
the  experiment  alone.     We  could  here  have  written 

p  =  f(v) 
just  as  correctly  as  v  =  f(p).      In  actions  involving   time  it   is 
customary,  though  not  essential,  to  regard  the  latter  as  the  in- 
dependent variable,  since  time  changes  in  a  most  uniform  and 
independent  way.  ■  Time  is  the  natural  independent  variable. 

In  the  same  way  the  area  of  a  circle  is  a  function  of  the  radius, 
so  is  the  volume  of  a  sphere ;  the  pressure  of  a  gas  is  a  function 
of  the  density ;  the  volume  of  a  gas  is  a  function  of  the  tempera- 
ture ;  the  amount  of  substance  formed  in  a  chemical  reaction  is  a 
function  of  the  time ;  the  velocity  of  an  explosion  wave  is  a  func- 
tion of  the  density  of  the  medium  ;  the  boiling  point  of  a  liquid  is  a 
function  of  the  atmospheric  pressure  ;  the  resistance  of  a  wire  to  the 
passage  of  an  electric  current  is  a  function  of  the  thickness  of  the 
wire ;  the  solubility  of  a  salt  is  a  function  of  the  temperature,  etc. 

The  independent  variable  may  be  denoted  by  x,  when  writing 
in  general  terms,  and  the  dependent  variable  by  y.  The  relation 
between  these  variables  is  variously  denoted  by  the  symbols : 

y  =/(*) ;  y  =  <K#) ;  y  =  F(x) ;  y  =  ${x) ;  y  =  Mx) . . . 

Any  one  of  these  expressions  means  nothing  more  than  that  "y  is 
some  function  ofx".  If  xv  yY ;  x2,  y2 ;  xz,  y3, . . .  are  corresponding 
values  of  x  and  y,  we  may  have 

y=Ax)\  2/i=/(*i);  yt.f?/Wv' 

"Let  y=f(%)"  means  "take  any  equation  which  will  enable 
you  to  calculate  y  when  the  value  of  x  is  known  ". 

The  word  "function"  in  mathematical  language  thus  implies 


I 


§  9.  THE  DIFFERENTIAL  CALCULUS.  21 

that  for  every  value  of  x  there  is  a  determinate  value  of  y.  If  v0 
and  p0  are  the  corresponding  values  of  the  pressure  and  volume  of 
a  gas  in  any  given  state,  v  and  p  their  respective  values  in  some 
other  state,  Boyle's  law  states  that  pv  =  p0v0.     Hence, 

n      Povo.  nr    _     Povo 

p  =  ;  or,  v  = • 

r  V  p 

The  value  of  p  or  of  v  can  therefore  be  determined  for  any 
assigned  value  of  v  or  p  as  the  case  might  be. 

A  similar  rule  applies  for  all  physical  changes  in  which  two 
magnitudes  simultaneously  change  their  values  according  to  some 
fixed  law.  It  is  quite  immaterial,  from  our  present  point  of  view, 
whether  or  not  any  mathematical  expression  for  the  function  f(x) 
is  known.  For  instance,  although  the  pressure  of  the  aqueous 
vapour  in  any  vessel  containing  water  and  steam  is  a  function  of 
the  temperature,  the  actual  form  of  the  expression  or  function 
showing  this  relation  is  not  known ;  but  the  laws  connecting  the 
volume  of  a  gas  with  its  temperature  and  pressure  are  known 
functions  —  Boyle's,  and  Charles'  laws.  The  concept  thus 
remains  even  though  it  is  impossible  to  assign  any  rule  for  cal- 
culating the  value  of  a  function.  In  such  cases  the  corresponding 
value  of  each  variable  can  only  be  determined  by  actual  obser- 
vation and  measurement.  In  other  words,  f(x)  is  a  convenient 
symbol  to  denote  any  mathematical  expression  containing  x. 

From  pages  8  and  17,  since 

V  -  /(»), 
the  differential  coefficient  dy/dx  is  another  function  of  x,  say  f(x), 

Similarly  the  second  derivative,  d2y/dx2,  is  another  function  of  x, 

and  so  on  for  the  higher  differential  functions. 

The  above  investigation  may  be  extended  to  functions  of  three 
or  more  variables.  Thus,  the  volume  of  a  gas  is  a  function  of  the 
pressure  and  temperature  ;  the  amount  of  light  absorbed  by  a 
solution  is  a  function  of  the  thickness  and  concentration  of  the 
solution ;  and  the  growth  of  a  tree  depends  upon  the  fertility  of  the 
soil,  the  rain,  solar  heat,  etc.     We  have  tacitly  assumed,  in  our 


22  HIGHER  MATHEMATICS.  §  10. 

preceding  illustration,  that  the  temperature  was  constant.     If  the 
pressure  and  temperature  vary  simultaneously,  we  write 

I  must  now  make  sure  that  the  reader  has  clear  ideas  upon  the 
subjects  discussed  in  the  two  following  articles ;  we  will  then  pass 
directly  to  the  "  calculus  "  itself. 

§  10.  Proportionality  and  the  Variation  Constant. 

When  two  quantities  are  so  related  that  any  variation  (increase 
or  decrease)  in  the  value  of  the  one  produces  a  proportional  varia- 
tion (increase  or  decrease)  in  the  value  of  the  other,  the  one 
quantity  is  said  to  be  directly  proportional  to  the  other,  or  to  vary 
as,  or  to  vary  directly  as,  the  other.  For  example,  the  pressure  of 
a  gas  is  proportional  to  its  density;  the  velocity  of  a  chemical 
reaction  is  proportional  to  the  amount  of  substance  taking  part  in 
the  reaction ;  and  the  area  of  a  circle  is  proportional  to  the  square 
of  the  radius. 

On  the  other  hand,  when  two  quantities  are  so  related  that  any 
increase  in  the  value  of  the  one  leads  to  a  proportional  decrease  in 
the  value  of  the  other,  the  one  quantity  is  said  to  be  inversely 
proportional  to,  or  to  vary  inversely  as  the  other.  Thus,  the  pres- 
sure of  a  gas  is  inversely  proportional  to  its  volume,  or  the  volume 
inversely  proportional  to  the  pressure ;  and  the  number  of  vibra- 
tions emitted  per  second  by  a  sounding  string  varies  inversely  as 
the  length  of  the  string. 

The  symbol  "  oc  "  denotes  variation.  For  x  oc  y,  read  "  x  varies 
as  y";  and  for  scoc  y'1,  read  "x  varies  inversely  as  y  ".  The 
variation  notation  is  nothing  but  abbreviated  proportion.  Let 
xv  y1;  x2,  y2 ; . . .  be  corresponding  values  of  x  and  y.  Then,  if  x 
varies  as  y, 

xi  '•  V\  ~  x2  '•  Vi  ~  xs  '  Vz  ~  •  •  • » 
or,  what  is  the  same  thing, 

Ui    y%    y%    '"'  K ) 

Since  the  ratio  of  any 1  value  of  x  to  the  corresponding  value  of  y 


1  It  is  perhaps  needless  to  remark  that  what  is  true  of  any  value  is  true  for  all. 
If  any  apple  in  a  barrel  is  bad,  all  are  bad. 


§  10.  THE  DIFFERENTIAL  CALCULUS.  23 

is  always  the  same,  it  follows  at  once  that  xjy  is  a  constant ;  and 
xy  is  a  constant  when  x  varies  inversely  as  y,  as  p  and  v  in 
"  Boyle's  law  ".     In  symbols,  if 

1  k 

x  oc  y,  x  =  ky ;  and  if  x  cc  -,  x  —  -.     .         .         (2) 

This  result  is  of  the  greatest  importance.  It  is  utilized  in  nearly 
every  formula  representing  a  physical  process,  k  is  called  the 
constant  of  proportionality,  or  constant  of  variation. 

We  can  generally  assign  a  specific  meaning  to  the  constant  of 
proportion.  For  example,  if  we  know  that  the  mass,  m,  of  a  sub- 
stance is  proportional  to  its  volume,  v, 

.'.  m  =  kv. 
If  we  take  unit  volume,  v  =  1,  k  =  m,  k  will  then  represent  the 
density,  i.e.,  the  mass  of  unit  volume,  usually  symbolized  by  p. 
Again,  the  quantity  of  heat,  Q,  which  is  required  to  warm  up 
the  temperature  of  a  mass,  m,  of  a  substance  0°  is  proportional  to 
m  x  6.     Hence, 

Q  =  kmO. 

If  we  take  m  —  1,  and  0  =  1°,  k  denotes  the  amount  of  heat  re- 
quired to  raise  up  the  temperature  of  unit  mass  of  substance  1°. 
This  constant,  therefore,  is  nothing  but  the  specific  heat  of  the  sub- 
stance, usually  represented  by  C  or  by  o-  in  this  work.  Finally, 
the  amount  of  heat,  Q,  transmitted  by  conduction  across  a  plate  is 
directly  proportional  to  the  difference  of  temperature,  0,  on  both 
sides  of  the  plate,  to  the  area,  s,  of  the  plate,  and  to  the  time,  t ; 
Q  is  also  inversely  proportional  to  the  thickness,  n,  of  the  plate. 
Consequently, 

By  taking  a  plate  of  unit  area,  and  unit  thickness  ;  by  keeping  the 
difference  of  temperature  on  both  sides  of  the  plate  at  1° ;  and  by 
considering  only  the  amount  of  heat  which  would  pass  across  the 
plate  in  unit  time,  Q  =  k ;  k  therefore  denotes  the  amount  of  heat 
transmitted  in  unit  time  across  unit  area  of  a  plate,  of  unit  thick- 
ness when  its  opposite  faces  are  kept  at  a  temperature  differing 
by  1°.  That  is  to  say,  k  denotes  the  coefficient  of  thermal  con- 
ductivity. 

The  constants  of  variation  or  proportion  thus  furnish  certain 
specific  coefficients  or  numbers  whose  numerical  values  usually 
depend  upon  the  nature  of  the  substance,  and  the  conditions  under 


24  HIGHER  MATHEMATICS.  §  11. 

which  the  experiment  is  performed.  The  well-known  constants : 
specific  gravity,  electrical  resistance,  the  gravitation  constant,  -n, 
and  the  gas  constant,  B,  are  constants  of  proportion. 

Let  a  gas  be  in  a  state  denoted  by  pv  pv  and  Tv  and  suppose 
that  the  gas  is  transformed  into  another  state  denoted  by  p2,  p2, 
and  T2.     Let  the  change  take  place  in  two  stages  : — 

First,  let  the  pressure  change  from  px  to  p2  while  the  tempera- 
ture remains  at  Tv  Let  pv  in  consequence,  become  x.  Then, 
according  to  Boyle's  law, 

a  ja.  ■...,_«&.    ...     (3) 

Second,  let  the  pressure  remain  constant  at  p.2  while  the  tem- 
perature changes  from  TY  to  T2.  Let  x,  in  consequence,  become 
p2.     Then,  by  Charles'  law, 

PJFlm*Pi •         W 

Substituting  the  above  value  of  x  in  this  equation,  we  get 

~fT  =  "%-  =  constant,  say,  R.       .-.  -  =  BT. 

Pl-Ll  P<LL<L  P 

We  therefore  infer  that  if  x}  y,  z,  are  variable  magnitudes  such 
that  xazy,  when  z  is  constant,  and  x  oc  z,  when  y  is  constant,  then, 
x  oc  yz,  when  y  and  z  vary  together ;  and  conversely,  it  can  also 
be  shown  that  if  x  varies  as  y,  when  z  is  constant,  and  x  varies  in- 
versely as  z,  when  y  is  constant,  then,  x  cc  y/z  when  y  and  z  both 
vary. 

Examples. — (1)  If  the  volume  of  a  gas  varies  inversely  as  the  pressure 
and  directly  as  the  temperature,  show  that 

"^  =  ^  ;  and  that  pv  =  RT.  .  (5) 

(2)  If  the  quantity  of  heat  required  to  warm  a  substance  varies  directly 
as  the  mass,  m,  and  also  as  the  range  of  temperature,  6,  show  that  Q  =a  amd. 

(3)  If  the  velocity,  V,  of  a  chemical  reaction  is  proportional  to  the  amount 
of  each  reacting  substance  present  at  the  time  t,  show  that 

F=feC1C203...Cn, 
where  Clt  C2,  C3,  . . . ,  C„,  respectively  denote  the  amount  of  each  of  the  n 
reacting  substances  at  the  time  t,  k  is  constant. 

§  11.  The  Laws  of  Indices  and  Logarithms. 

We  all  know  that 

4x4=    16,  is  the  second  power  of  4,  written  42 ; 
4x4x4=    64,  is  the  third  power  of  4,  written  43 ; 
4.x  4x4x4  =  256,  is  the  fourth  power  of  4,  written  44; 


§  ll.  THE  DIFFERENTIAL  CALCULUS.  25 

and  in  general,  the  nth  power  of  any  number  &,  is  denned  as  the 
continued  product 

a  x  a  x  a  x  ...  n  times  =  an,         .        .         (1) 
where  n  is  called  the  exponent,  or  index  of  the  number. 

By  actual  multiplication,  therefore,  it  follows  at  once  that 
102  x  103  =  10  x  10  x  10  x  10  x  10  =  102  + 3  =  105  =  100,000  ; 
or,  in  general  symbols, 

am  x  an  =  am  +  n;  or,  ax  x  a?  x  a1  x  . . .  =  ax+v+t+-"  ,         (2) 
a  result  known  as  the  index  law. 

All  numbers  may  be  represented  as  different  powers  of  one 
fundamental  number.     E.g., 

1  =  10°;  2  =  10-301;  3  =  10*477;  4  =  10-602;  5  -  10-699;  ... 
The  power,  index  or  exponent  is  called  a  logarithm,  the  funda- 
mental  number  is  called  the  base  of  the  system  of  logarithms, 
Thus  if 

an  =  b, 
n  is  the  logarithm  of  the  number  b  to  the  base  a,  and  is  written 

n  =  loga6. 
For  convenience  in  numerical  calculations  tables  are  generally  used 
in  which  all  numbers  are  represented  as  different  powers  of  10. 
The  logarithm  of  any  number  taken  from  the  table  thus  indicates 
what  power  of  10  the  selected  number  represents.     Thus  if 

103  =  1000;    and  101  '0413927  =  Q , 

then  3  =  log101000 ;  and  1-0413927  =  log10ll. 

We  need  not  use  10.  Logarithms  could  be  calculated  to  any 
other  number  employed  as  base.  If  we  replace  10  by  some  other 
number,  say,  2-71828,  which  we  represent  by  e,  then 

l  =  e°;  2  =  e0*693;  3  =  e1*099;  4  =  e1>386;   5  =  e1*609;  ... 

Logarithms  to  the  base  e  =  2*71828  are  called  natural,  hyper- 
bolic, or  Napierian  logarithms.  Logarithms  to  the  base  10  are 
called  Briggsian,  or  common  logarithms. 

Again, 

3x5  =  (10°'4771)  x  (10°-6990)  =  101'1761  =  15, 
because,  from  a  table  of  common  logarithms, 

log103  =  0-4771;  log105  =  0-6990;  log1015  =  1-1761. 
Thus  we  have  performed  arithmetical  multiplication  by  the  simple 
addition  of  two  logarithms.     Generalizing,  to  multiply  two  or  more 


26  HIGHER  MATHEMATICS.  §  11. 

numbers,  add  the  logarithms  of  the  numbers  and  find  the  number 
whose  logarithm  is  the  sum  of  the  logarithms  just  obtained. 
Example. — Evaluate  4  x  80, 

log10  4  =  0-6021 
log1080  =  1-9031 

Sum  =  2-5052  =  log10320. 
This  method  of  calculation  holds  good  whatever  numbers  we  employ  in  place 
of  3  and  5  or  4  and  80.     Hence  the  use  of  logarithms  for  facilitating  numerical 
calculations.     We  shall  shortly  show  how  the  operations  of  division,  involu- 
tion, and  evolution  are  as  easily  performed  as  the  above  multiplication. 

From  what  has  just  been  said  it  follows  that 

103  am 

jp  =  103  ~  2  =  101  =  10  J  or  generally,  —  =  CLm  ~  n.      .  (3) 

Hence  the  rule  :  To  divide  two  numbers,  subtract  the  logarithm 
of  the  divisor  (denominator  of  a  fraction)  from  the  logarithm  of  the 
dividend  (numerator  of  a  fraction)  and  find  the  number  correspond- 
ing to  the  resulting  logarithm. 

Examples. — (1)  Evaluate  60  -f  3. 

log1060  =  1-7782 
log10  3  =  0-4771 

Difference  =  1-3011  =  log1020. 

(2)  Show  that  2"2  =  £;  H)-2  =  TU;  33  x  3~3=1. 

i_  i 

(3)  Show  that  ax  x  a1-*  =  a;  p+  px  =  pl~ x\  a*  +  a  =  a-*1-*). 

The  general  symbols  a,  b,  ...  m,  n,  ...  x,  y,  ...  in  any- 
general  expression  may  be  compared  with  the  blank  spaces  in  a 
bank  cheque  waiting  to  have  particular  values  assigned  to  date, 
amount  (£  s.  d.),  and  sponsor,  before  the  cheque  can  fulfil  the 
specific  purpose  for  which  it  was  designed.  So  must  the  symbols, 
a,  b,  ...  of  a  general  equation  be  replaced  by  special  numerical 
values  before  the  equation  can  be  applied  to  any  specific  process 
or  operation. 

It  is  very  easy  to  miss  the  meaning  of  the  so-called  "  properties 
of  indices,"  unless  the  general  symbols  of  the  text-books  are 
thoroughly  tested  by  translation  into  numerical  examples.  The 
majority  of  students  require  a  good  bit  of  practice  before  a  general 
expression  appeals  to  them  with  full  force.  Here,  as  elsewhere, 
it  is  not  merely  necessary  for  the  student  to  think  that  he  "  under- 
stands the  principle  of  the  thing,"  he  must  actually  work  out 
examples  for  himself.     • '  In  scientiis  ediscendis  prosunt  exempla 


§  11.  THE  DIFFEKENTIAL  CALCULUS.  27 

magis  quam  praecepta  " 1  is  as  true  to-day  as  it  was  in  Newton's 
time.  For  example,  how  many  realise  why  mathematicians  write 
e°  =  1,  until   some   such   illustration  as.  the  following  has  been 

worked  out? 

22  x  2°  =  22  +  °  =  22  =  4  =  22  x  1. 

The  same  result,  therefore,  is  obtained  whether  we  multiply  22  by 
2°  or  by  1,  i.e., 

'      22  x  20  =  22  x  1  =  22  =  4. 

Hence  it  is  inferred  that 2 

2°  =  1,  and  generally  that  a0  =  1.  .  .  (4) 

Example. — From  the  Table  on  page  628,  show  that 

loge3  =  1-0986 ;  loge2  =  0-6932 ;  log.l  =0.  .  (5) 

And,  since 

e  x  e  x  e  x  ...  n  times  =  en ;  . . . ;  e  x  e  x  e  —  e3 ;  e  x  e  =  e2 ;  e  =  el ; 

.-.  logeen  =  n ;  ...  log^3  =  3 ;  log«e2  =  2 ;  log^1  =  1  =  logee.     .        (6) 

I  am  purposely  using  the  simplest  of  illustrations,  leaving  the 
reader  to  set  himself  more  complicated  numbers.  No  pretence  is 
made  to  rigorous  demonstration.  We  assume  that  what  is  true  in 
one  case,  is  true  in  another.  It  is  only  by  so  collecting  our  facts 
one  by  one  that  we  are  able  to  build  up  -a  general  idea.  The  be- 
ginner should  always  satisfy  himself  of  the  truth  of  any  abstract 
principle  or  general  formula  by  applying  it  to  particular  and  simple 
cases. 

To  find  the  relation  between  the  logarithms  of  a  number  to  dif- 
ferent bases.  Let  n  be  a  number  such  that  aa^=n;  or,  a  =  logan; 
and  (5b  =  n,  or,  b  =  log^n.     Hence  aa  =  flb.     "  Taking  logs  "  to  the 

base  a,  we  obtain 

a  =  b  loga/?, 

since  logaa  is  unity.     Substitute  for  a  and  b,  and  we  get 

logan  =  log/sw .  loga/?.  .  .  .  (7) 
In  words,  the  logarithm  of  a  number  to  the  base  ft  may  be  obtained 
from  the  logarithm  of  that  number  to  the  base  a  by  multiplying  it 
by  l/loga/?.     For  example,  suppose  a  =  10  and  (3  =  e, 

"«--iss « 

1  Which  may  be  rendered:  "In  learning  we  profit  more  by  example  than  by 
precept ". 

2  Some  mathematicians  define  a«asl  *  ax  ax  a  . . .  n  times  ;  a3  =  lxa  x  ax  a; 
a2  =  \xaxa\  a1  =  1  x  a  ;  and  a0  as  1  x  a  no  times,  that  is  unity  itself.  If  so,  then 
I  suppose  that  0°  must  mean  1  x  0  no  times,  i.e.,  1 ;  and  1/0°  must  mes.n  ?/(]  x  0  no 
times),  i.e.,  unity. 


28  HIGHER  MATHEMATICS.  §  11. 

i 

where  the  subscript  in  logen  is  omitted.     It  is  a  common  practice 

to   omit  the  subscript  of  the  "  log  "  when  there  is  no  danger  of 
ambiguity.    Hence,  since  log102'71828  =  0-4343,  and  loge10  =  2-3026, 
where  2-71828  is  the  nominal  value  of  e  (page  25)  : — 
To  pass  from  natural  to  common  logarithms 

Common  log  =  natural  log  x  0-43431  /q<> 

log10a  =  logea  x  0*4343            J   '  *  ' 
To  pass  from  common  to  natural  logarithms 

Natural  log  =  common  log  x  2-30261  ,*^ 

logea  =  log10a  x  2-3026            J  {     } 

The  number  0*4343  is  called  the  modulus  of  the  Briggsian  or 
common  system  of  logarithms.  When  required  it  is  written  M. 
or  fx.  It  is  sufficient  to  remember  that  the  natural  logarithm  of  a 
number  is  2*3026  times  greater  than  the  common  logarithm. 

By  actual  multiplication  show  that 

(100)3  =  (102)3  =  102  x  3  =  106, 

and  hence,  to  raise  a  number  to  any  power,  multiply  the  loga- 
rithm of  the  number  by  the  index  of  the  power  and  find  the  number 
corresponding  to  the  resulting  logarithm. 

(am)n  =  amn (11) 

Example. — Evaluate  52. 

52  =  (5)2  =  (lO0"6990)2  =  101B98°  =  25, 
since  reference  to  a  table  of  common  logarithms  shows  that 
log105  =  0-6990 ;  log1025  =  1-3980. 

From  the  index  law,  above 

10*  x  10i  =  10*  +  *  =  101  =  10. 
That  is  to  say,  10*  multiplied  by  itself  gives  10.     But  this  is  the 
definition  of  the  square  root  of  10. 

.*.  ( JlO)2  =  JlO  Wl0=  10*  x  10*  =  10. 

A  fractional  index,  therefore,  represents  a  root  of  the  particular 
number  affected  with  that  exponent.     Similarly, 

4/8  =  8*,  because  ^8  x  ^8  x  J/8  =  Sh  x  8*  x  8*  =  8. 

Generalizing  this  idea,  the  nth  root  of  any  number  a,  is 

nJ'a  =  a\      .         .         .         .         (12) 
To  extract  the  root  of  any  number,  divide  the  logarithm  of 


§  12.  THE  DIFFERENTIAL  CALCULUS.  29 

the  number  by  the  index  of  the  required  root  and  find  the  number 
corresponding  to  the  resulting  logarithm. 

Examples.— (1)  Evaluate  #8  and  V93". 
#8  =  (8)*  =  (100'9031)*  =  100'3010  =  2 ;  Z/93  =  (93)f  =  (1019685)*  =  1002812  =  1-91, 
since,  from  a  table  of  common  logarithms, 

log102  =  0-3010 ;  log108  =  0-9031 ;  log10l-91  =  0-2812 ;  log1093  =  1-9685. 

(2)  Perhaps  this  will  amuse  the  reader  some  idle  moment.  Given  the 
obvious  facts  log  J  =  log£,  and  3>2  ;  combining  the  two  statements  we  get 
3  log  £>  2  log  J;  .-.  log(J)8>log(J)»;  .-.  fe)3>(£)2;  .'.  4>i;  .'-  1  is  greater 
than  2.     Where  is  the  fallacy  ? 

The  results  of  logarithmic  calculations  are  seldom  absolutely 
correct  because  we  employ  approximate  values  of  the  logarithms 
of  the  particular  numbers  concerned.  Instead  of  using  logarithms 
to  four  decimal  places  we  could,  if  stupid  enough,  use  logarithms 
accurate  to  sixty-four  decimal  places.  But  the  discussion  of  this 
question  is  reserved  for  another  chapter.  If  the  student  has  any 
difficulty  with  logarithms,  after  this,  he  had  better  buy  F.  G. 
Taylor's  An  Introduction  to  the  Practical  Use  of  Logarithms, 
London,   1901. 

§  12.  Differentiation,  and  its  Uses. 

The  differential  calculus  is  not  directly  concerned  with  the 
establishment  of  any  relation  between  the  quantities  themselves, 
but  rather  with  the  momentary  state  of  the  phenomenon.  This 
momentary  state  is  symbolised  by  the  differential  coefficient,  which 
thus  conveys  to  the  mind  a  perfectly  clear  and  definite  conception 
altogether  apart  from  any  numerical  or  practical  application.  I 
suppose  the  proper  place  to  recapitulate  the  uses  of  the  differential 
calculus  would  be  somewhere  near  the  end  of  this  book,  for  only 
there  can  the  reader  hope  to  have  his  faith  displaced  by  the  certainty 
of  demonstrated  facts.  Nevertheless,  I  shall  here  illustrate  the 
subject  by  stating  three  problems  which  the  differential  calculus 
helps  us  to  solve. 

In  order  to  describe  the  whole  history  of  any  phenomenon  it  is 
necessary  to  find  the  law  which  describes  the  relation  between  the 
various  agents  taking  part  in  the  change  as  well  as  the  law  describ- 
ing the  momentary  states  of  the  phenomenon.  There  is  a  close 
connection  between  the  two.  The  one  is  conditioned  by  the  other. 
Starting  with  the  complete  law  it  is  possible  to  calculate  th6 
momentary  states  and  conversely. 


30  HIGHER  MATHEMATICS.  §  12. 

I.  The  calculation  of  the  momentary  states  from  the  complete 
law.  Before  the  instantaneous  rate  of  change,  dyjdx,  can  be  deter- 
mined it  is  necessary  to  know  the  law,  or  form  of  the  function 
connecting  the  varying  quantities  one  with  another.  For  instance, 
Galileo  found  by  actual  measurement  that  a  stone  falling  vertically 
downwards  from  a  position  of  rest  travels  a  distance  of  s  =  \gt2 
feet  in  t  seconds.  Differentiation  of  this,  as  we  shall  see  very 
shortly,  furnishes  the  actual  velocity  of  the  stone  at  any  instant 
of  time,  V  =  gt.  In  the  same  manner,  Newton's  law  of  inverse 
squares  follows  from  Kepler's  third  law ;  and  Ampere's  law,  from 
the  observed  effect  of  one  part  of  an  electric  circuit  upon  another. 

II.  The  calculation  of  the  complete  law  from  the  momentary 
states.  It  is  sometimes  possible  to  get  an  idea  of  the  relations 
between  the  forces  at  work  in  any  given  phenomenon  from  the 
actual  measurements  themselves,  but  more  frequently,  a  less  direct 
path  must  be  followed.  The  investigator  makes  the  most  plausible 
guess  about  the  momentary  state  of  the  phenomenon  at  his  com- 
mand, and  dresses  it  up  in  mathematical  symbols.  Subsequent 
progress  is  purely  an  affair  of  mathematical  computation  based 
upon  the  differential  calculus.  Successful  guessing  depends  upon 
the  astuteness  of  the  investigator.  This  mode  of  attack  is  finally 
justified  by  a  comparison  of  the  experimental  data  with  the  hypo- 
thesis dressed  up  in  mathematical  symbols,  and  thus 

The  golden  guess 
Is  morning  star  to  the  full  round  of  truth. 

Fresnel's  law  of  double  refraction,  Wilhelmy's  law  of  mass  action, 
and  Newton's  law  of  heat  radiation  may  have  been  established  in  this 
way.  The  subtility  and  beauty  of  this  branch  of  the  calculus  will  not 
appear  until  the  methods  of  integration  have  been  discussed. 

III.  The  educti/m  of  a  generalization  from  particular  cases. 
A  natural  law,  deduced  directly  from  observation  or  measurement, 
can  only  be  applied  to  particular  cases  because  it  is  necessarily 
affected  by  the  accidental  circumstances  associated  with  the  con- 
ditions under  which  the  measurements  were  made.  Differentiation 
will  eliminate  the  accidental  features  so  that  the  essential  circum- 
stances, common  to  all  the  members  of  a  certain  class  of  phenomena, 
alone  remain.  Let  us  take  one  of  the  simplest  of  illustrations,  a 
train  travelling  with  the  constant  velocity  of  thirty  miles  an  hour. 
Hence,  V  =  30.     From  what  we  have  already  said,  it  will  be  clear 


§12. 


THE  DIFFERENTIAL  CALCULUS. 


31 


that  the  rate  of  change  of  velocity,  at  any  moment,  is  zero.  Other- 
wise expressed,  dV/dt  =  0.  The  former  equation,  V  =  30,  is  only 
"true  of  the  motion  of  one  particular  object,  whereas  dV/dt  —  0,  is 
true  of  the  motion  of  all  bodies  travelling  with  a  constant  velocity. 
In  this  sense  one  reliable  observation 
might  give  rise  to  a  general  law. 

The  mechanical  operations  of  finding 
the  differential  coefficient  of  one  vari- 
able with  respect  to  another  in  any 
expression  are  no  more  difficult  than 
ordinary  algebraic  processes.  Before  de- 
scribing the  practical  methods  of  differ- 
entiation it  will  be  instructive  to  study 
a  geometrical  illustration  of  the  process. 

Let  x  (Fig.  5)  be  the  side  of  a  square,  and  let  there  be  an  incre- 
ment1 in  the  area  of  the  square  due  to  an  increase  of  h  in  the 
variable  x. 


Fig.  5. 


The  original  area  of  the  square  =  x2. 
The  new  area  =  (x  +  h)2 

The  increment  in  the  area  =  (a;  +  h)2 


x2  +  2xh  +  h2. 


=  2xh  +  h2.  .         (3) 

This  equation  is  true,  whatever  value  be  given  to  h.  The 
Bmaller  the  increment  h  the  less  does  the  value  of  h2  become. 
If  this  increment  h  ultimately  become  indefinitely  small,  then  h2, 
being  of  a  very  small  order  of  magnitude,  may  be  neglected.  For 
example,  if  when  x  =  1, 

h  =  1,  increment  in  area  =  2  +  1 ; 
fr  =  xV>        H  »       =0-2  +  ^; 


mnr> 


0-002  + 


1.0OQ.000' 


etc. 


If,  therefore,  dy  denotes  the  infinitely  small  increment  in  the 
area,  y,  of  the  square  corresponding  to  an  infinitely  small  incre- 
ment dx  in  two  adjoining  sides,  x,  then,  in  the  language  of 
differentials, 

Increment  y  =  2xh,  becomes,  dy  =  2# .  dx.  .  (4) 

The  same  result  can  be  deduced  by  means  of  limiting  ratios. 
For  instance,  consider  the  ratio  of  any  increment  in  the  area,  y, 
to  any  increment  in  the  length  of  a  side  of  the  square,  x. 


1  When  any  quantity  is  increased,  the  quantity  by  which  it  is  increased  is  called 
its  increment,  abbreviated  "incr."  ;  a  decrement  is  a  negative  increment. 


,32  HIGHER  MATHEMATICS.  §  13. 

Increment  y  _  Incr.  y        hy        2xh  +  h2 

and  when  the  value  of  h  is  zero 

£  =  Uh  =  0£  =  2x.        .         .         .         (5) 

To  measure  the  rate  of  change  of  any  two  variables,  we  fix  upon 
one  variable  as  the  standard  of  reference.  When  x  is  the  standard 
of  reference  for  the  rate  of  change  of  the  variable  y,  we  call  dy/dx 
the  Aerate  of  y.  In  practical  work,  the  rate  of  change  of  time,  t, 
is  the  most  common  standard  of  reference.  If  desired  we  can 
interpret  (4)  or  (5)  to  mean 

dy  dx 

Tt  =  2x  ~df 

In  words,  the  rate  at  which  y  changes  is  2x  times  the  rate  at  which 
x  changes. 

Examples. — (1)  Show,  by  similar  reasoning  to  the  above,  that  if  the  three 

By 

adjoining  sides,  x,  of  a  cube  receive  an  increment  h,  then  Lt»  =  o  g-  =  3x2. 

(2)  Prove  that  if  the  radius,  r,  of  a  circle  be  increased  by  an  amount  h, 
the  increment  in  the  area  of  the  circle  will  be  (2rh  +  h2)ir.  Show  that  the 
limiting  ratio,  dy/dx,  in  this  case  is  lirr.     Given,  area  of  circle  =  irr2. 

The  former  method  of  differentiation  is  known  as  "  Leibnitz's 
method  of  differentials,"  the  latter,  "  Newton's  method  of  limits  ". 
It  cannot  be  denied  that  while  Newton's  method  is  rigorous, 
exact,  and  satisfying,  Leibnitz's  at  once  raises  the  question : 

§  13.  Is  Differentiation  a  Method  of  Approximation  only  ? 

The  method  of  differentiation  might  at  first  sight  be  regarded 
as  a  method  of  approximation,  for  these  small  quantities  appear 
to  be  rejected  only  because  this  may  be  done  without  committing 
any  sensible  error.  For  this  reason,  in  its  early  days,  the  calculus 
was  subject  to  much  opposition  on  metaphysical  grounds.  Bishop 
Berkeley x  called  these  limiting  ratios  "  the  ghosts  of  departed 
quantities  ".  A  little  consideration,  however,  will  show  that  these 
small  quantities  must  be  rejected  in  order  that  no  error  may  be 
committed  in  the  calculation.  The  process  of  elimination  is 
essential  to  die  operation. 


i  G.  Berkeley,  Collected  Works,  Oxford,  3,  44,  1901. 


§  13.  THE  DIFFERENTIAL  CALCULUS.  S3 

There  has  been  a  good  bit  of  tinkering,  lately,  at  the  founda- 
tions of  the  calculus  as  well  as  other  branches  of  mathematics,  but 
we  cannot  get  much  deeper  than  this  :  assuming  that  the  quantities 
under  investigation  are  continuous,  and  noting  that  the  smaller  the 
differentials  the  closer  the  approximation  to  absolute  accuracy,  our 
reason  is  satisfied  to  reject  the  differentials,  when  they  become  so 
small  as  to  be  no  longer  perceptible  to  our  senses.  The  psycho- 
logical process  that  gives  rise  to  this  train  of  thought  leads  to  the 
inevitable  conclusion  that  this  mode  of  representing  the  process  is 
the  true  one.  Moreover,  if  this  be  any  argument,  the  validity  of 
the  reasoning  is  justified  by  its  results. 

The  following  remarks  on  this  question  are  freely  translated 
from  Carnot's  Beflexions  sur  la  Mdtaphysique  du  Galcul  In- 
finitesimal.1 "The  essential  merit,  the  sublimity,  one  may  say,  of 
the  infinitesimal  method  lies  in  the  fact  that  it  is  as  easily  performed 
as  a  simple  method  of  approximation,  and  as  accurate  as  the  results 
of  an  ordinary  calculation.  This  immense  advantage  would  be 
lost,  or  at  any  rate  greatly  diminished,  if,  under  the  pretence  of 
obtaining  a  greater  degree  of  accuracy  throughout  the  whole  pro- 
cess, we  were  to  substitute  for  the  simple  method  given  by  Leibnitz 
one  less  convenient  and  less  in  accord  with  the  probable  course  of 
the  natural  event.  If  this  method  is  accurate  in  its  results,  as  no 
one  doubts  at  this  day ;  if  we  always  have  recourse  to  it  in  difficult 
questions,  what  need  is  there  to  supplant  it  by  complicated  and 
indirect  means  ?  Why  content  ourselves  with  founding  it  on  in- 
ductions and  analogies  with  the  results  furnished  by  other  means 
when  it  can  be  demonstrated  directly  and  generally,  more  easily, 
perhaps,  than  any  of  these  very  methods  ?  The  objections  which 
have  been  raised  against  it  are  based  on  the  false  supposition  that 
the  errors  made  by  neglecting  infinitesimally  small  quantities 
during  the  actual  calculation  are  still  to  be  found  in  the  result  of 
the  calculation,  however  small  they  may  be.  Now  this  is  not  the 
case.  The  error  is  of  necessity  removed  from  the  result  by  elimi- 
nation. It  is  indeed  a  strange  thing  that  every  one  did  not  from 
the  very  first  realise  the  true  character  of  infinitesimal  quantities, 
and  see  that  a  conclusive  answer  to  all  objections  lies  in  this  indis- 
pensable process  of  elimination." 

The  beginner  will  have  noticed  that,  unlike  algebra  and  arith- 


i  Paris,  215,  1813. 
0 


34  HIGHER  MATHEMATICS.  §  13. 

metic,  higher  mathematics  postulates  that  number  is  capable  of 
gradual  growth.  The  differential  calculus  is  concerned  with  the 
rate  at  which  quantities  increase  or  diminish.  There  are  three 
modes  of  viewing  this  growth  : — 

i".  Leibnitz's  ■'  method  of  infinitesimals  or  differentials  ".  Accord- 
ing to  this,  a  quantity  is  supposed  to  pass  from  one  degree  of  mag- 
nitude to  another  by  the  continual  addition  of  infinitely  small  parts, 
called  infinitesimals  or  differentials.  Infinitesimals  may  have 
different  orders  of  magnitude.  Thus,  the  product  dx.dy  is  an  in- 
finitesimal of  the  second  order,  infinitely  small  in  comparison  with 
the  product  y.dx,  or  x.dy. 

In  a  preceding  section  it  was  shown  that  when  each  of  two 
sides  of  a  square  receives  a  small  increment  h,  the  corresponding 
increment  in  the  area  is  2xh  +  h2.  When  h  is  made  indefinitely 
small  and  equal  to  say  dx,  then  (dx)2  is  vanishingly  small  in  com- 
parison with  x.dx.     Hence, 

dy  =  Zx.dx. 

In  calculations  involving  quantities  which  are  ultimately  made 
to  approach  the  limit  zero,  the  higher  orders  of  infinitesimals  may 
be  rejected  at  any  stage  of  the  process.  Only  the  lowest  orders  of 
infinitesimals  are,  as  a  rule,  retained. 

II.  Newton's  "  method  of rates  or  fluxions  ".  Here,  the  velocity 
or  rate  with  which  the  quantity  is  generated  is  employed.  The 
measure  of  this  velocity  is  called  a  fluxion.  A  fluxion,  written  x,  y, 
. . .  ,  is  equivalent  to  our  dx/dt,  dy/dt} . . . 

These  two  methods  are  modifications'  of  one  idea.  It  is  all  a 
question  of  notation  or  definition.  While  Leibnitz  referred  the 
rate  of  change  of  a  dependent  variable  y,  to  an  independent  variable 
x,  Newton  referred  each  variable  to  "  uniformly  flowing  "  time. 
Leibnitz  assumed  that  when  x  receives  an  increment  dx,  y  is  in- 
creased by  an  amount  dy.  Newton  conceived  these  changes  to 
occupy  a  certain  time  dt,  so  that  y  increases  with  a  velocity  $,  as  x 
increases  with  a  velocity  x.  This  relation  may  be  written  sym- 
bolically, 

dy 

dx  =  xdt,  dy  =  ydt ;  .-.  ?  =  |L^. 
x    ax    U/X 

~di 
The  method  of  fluxions  is  not  in  general  use,  perhaps   because, 


§  14.  THE  DIFFERENTIAL  CALCULUS.  35 

of  its  more*  abstruse  character.  It  is  occasionally  employed  iii 
mechanics. 

III.  Neivton's  "  method  of  limits  ".  This  has  been  set  forth  in 
§  2.  The  ultimate  limiting  ratio  is  considered  as  a  fixed  quantity 
to  which  the  ratio  of  the  two  variables  can  be  made  to  approximate 
as  closely  as  we  please.  "The  limiting  ratio,"  says  Carnot,  "is 
neither  more  nor  less  difficult  to  define  than  an  infinitely  small 
quantity. ...  To  proceed  rigorously  by  the  method  of  limits  it  is 
necessary  to  lay  down  the  definition  of  a  limiting  ratio.  But  this 
is  the  definition,  or  rather,  this  ought  to  be  the  definition,  of  an 
infinitely  small  quantity."  "  The  difference  between  the  method 
of  infinitesimals  and  that  of  limits  (when  exclusively  adopted)  is, 
that  in  the  latter  it  is  usual  to  retain  evanescent  quantities  of  higher 
orders  until  the  end  of  the  calculation  and  then  neglect  them.  On 
the  otber  hand,  such  quantities  are  neglected  from  the  commence- 
ment in  the  infinitesimal  method  from  the  conviction  that  they 
cannot  affect  the  final  result,  as  they  must  disappear  when  we 
proceed  to  the  limit "  (Encyc.  Brit.).  It  follows,  therefore,  that  the 
psychological  process  of  reducing  quantities  down  to  their  limiting 
ratios  is  equivalent  to  the  rejection  of  terms  involving  the  higher 
orders  of  infinitesimals.  These  operations  have  been  indicated  side 
by  side  on  pages  31  and  32. 

The  methods  of  limits  and  of  infinitesimals  are  employed  in- 
discriminately in  this  work,  according  as  the  one  or  the  other 
appeared  the  more  instructive  or  convenient.  As  a  rule,  it  is  easier 
to  represent  a  process  mathematically  by  the  method  of  infinit- 
esimals. The  determination  of  the  limiting  ratio  frequently 
involves  more  complicated  operations  than  is  required  by  Leibnitz's 
method. 

§  1*.  The  Differentiation  of  Algebraic  Functions. 

We  may  now  take  up  the  routine  processes  of  differentiation. 
It  is  convenient  to  study  the  different  types  of  functions — alge- 
braic, logarithmic,  exponential,  •  and  trigonometrical — separately. 
An  algebraic  function  of  x  is  an  expression  containing  terms 
which  involve  only  the  operations  of  addition,  subtraction,  multi- 
plication, division,  evolution  (root  extraction),  and  involution.  For 
instance,  x2y  +  *Jx  +  yh  -  ax  =  1  is  an  algebraic  function.  Func- 
tions  that   cannot   be   so  expressed  are  termed  transcendental 

n  * 


36  HIGHER  MATHEMATICS.  §  14. 

functions.  Thus,  sin  x  =  y,  log  x  =  y,  ex  =  y  are  transcendental 
functions. 

On  pages  31  and  32  a  method  was  described  for  finding  the 
differential  coefficient  of  y  =  x2,  by  the  following  series  of  opera- 
tions : — (1)  Give  an  arbitrary  increment  h  to  x  in  the  original 
function ;  (2)  Subtract  the  original  function  x2  from  the  new  value 
of  (x  +  h)2  found  in  (1) ;  (3)  Divide  the  result  of  (2)  by  h  the  in- 
crement of  x ;  and  (4)  Find  the  limiting  value  of  this  ratio  when 
h  =  0. 

This  procedure  must  be  carefully  noted ;  it  lies  at  the  basis  of 
all  processes  of  differentiation.     In  this  way  it  can  be  shown  that 

if  y  =  *2> Tx =  2x;  {iy  =  x">Tx  =  ^>'liy  =  ^Tx  =  *x*> etc- 

By  actual  multiplication  we  find  that 

(x  +  hf  =  (x  +  h)  (x  +  h)  =  x2  +  2hx  +  h2 ; 

(x  +  hf  =  (x  +  h)2(x  +  h)  =  x*  +  3hx2  +  3h2x  +  /t3 ; 

Continuing  this  process  as  far  as  we  please,  we  shall  find  that 

(x  +  h)n  =  xn  +~xn~1h  +  n(n  ~^xn~2h2  +  ...  +  ^xhn-l  +  h«.     (1) 
1  1 .  A  1 

This  result,  known  as  the  binomial  theorem,  enables  us  to  raise  any 
expression  of  the  type  x  +  h  to  any  power  of  n  (where  n  is  positive 
integer,  i.e.,  a  positive  whole  number,  not  a  fraction)  without  going 
through  the  actual  process  of  successive  multiplication.  A  similar 
rule  holds  for  (x  -  h)n.  Now  try  if  this  is  so  by  substituting  n  =  1, 
2,  3,  4,  and  5  successively  in  (1),  and  comparing  with  the  results 
obtained  by  actual  multiplication. 

It  is  convenient  to  notice  that  the  several  sets  of  binomial  co- 
efficients obey  the  law  indicated  in  the  following  scheme,  as  n 
increases  from  0,  1,  2,  3 


(a  +  by  =  1 

m 

(a  +  bf  =  1     1 

(a  +  b)2  =  1     2 

1 

(a  +  by  =  1     3 

3      1 

(a  +  by  =  1     4 

6       4       1 

(a  +  b)6  =  15 

10     10       5     1 

(a  +  b)6  ^16 

15     20     15     6     1 

I.  Th  e  differential  coefficient 

of  any  power  of  a  variable. 

find  the  differential  coefficient  of 

y- 

=  X\ 

To 


§  14.  THE  DIFFERENTIAL  CALCULUS.  37 

Let  each  side  of  this  expression  receive  a  small  increment  so  that 
y  becomes  y  +  h'  when  x  becomes  x  +  h  ; 

.*.  (y  +  h')  -  y  =  incr.  y  =  (a?  +  ft)"  -  #n. 
From  the  binomial  theorem,  (1)  above, 

incr.  y  =  nxn  ~  lh  +  \n(n  -  l)xn  ~  2h2  +  . .  , 
Divide  by  increment  x,  namely  h. 

Incrj,  _  Incrj,  _  ^  .  ,  +  j  V    _  1)x„_2}l  +       _ 
/j  Incr.  £C 

Hence  when  h  is  made  zero 

T  ,        Incr.  y        T .    ,A         (x  +  h)n-Xn  „  _  , 

LtA=0  — JL  =  Limit*  .  0V i =  nr     K 

Incr.  X  'I 

That  is  to  say 

ft.^J-.^-i.        ...        (2) 

a#        arc 

Hence  the  rule  : — The  differential  coefficient  of  any  power  of  x 
is  obtained  by  diminishing  the  index  by  unity  and  multiplying  the 
power  of  x  so  obtained  by  the  original  exponent  (or  index). 

Examples.— (1)  If  y  =  x6 ;  show  that  dyfdx  =  6x*.  This  means  that  y 
changes  6x5  times  as  fast  as  x.  If  x  =  1,  y  increases  6  times  as  fast  as  x  ;  if 
x=  -  2,  y  decreases -6  x  32=  -  192  times  as  fast  as  x. 

(2)  If  y  =  x™  ;  show  that  dy\dx  =  20a?19. 

(3)  If  y  =  x5  ;  show  that  dyfdx  =  5aA 

(4)  If  y  =  <c3  ;  show  that  d?//<i£  =  300,  when  x  =  10. 

Later  on  we  deduce  the  binomial  theorem  by  differentiation. 
The  student  may  think  wTe  have  worked  in  a  vicious  circle.  This 
need  not  be.  The  differential  coefficient  of  xn  may  be  established 
without  assuming  the  binomial  theorem.     For  instance,  let 

y  =  xn, 

and  suppose  that  when  x  becomes  xx  =  x  +  h,y  becomes  yx  ;  then 
we  have 

Ux^y  ==ocln-xn  =  x^n_1  +  xx^n_2  +  ^  +  x,t_^ 

X-i      X         X-i  ~~  x 

by  division.     But  LtA=0a;1  =  x  ; 

dn 
.  \  -/    =  xn  ~  x  +  Xn  ~  l  +  ...  to  W  terms  =  1lXn "  l. 

ax 

II.  The  differential  coefficient  of  the  sum  or  difference  of  any 
number  of  functions.     Let  u,  v,  w . .  .  be  functions  of  x  ;  y  their 


38  HIGHER  MATHEMATICS.  §  14. 

sum.      Let  uv  vv  wv  . . . ,  yv  be  the  respective  values  of  these 
functions  when  x  is  changed  to  x  +  h,  then 

y  =  u  +  v  +  w  +  . . .  \  yY  =  uY  +  vx  +  wY  +  . . . 
.:  y1  -  y  =  {uY  -  u)  +  (vx  -  v)  +  (w1  -  w)  +  .  . . , 
by  subtraction  ;  dividing  by  h, 

Incr.  y        Incr.  U        Incr.  ^        Incr.  w 
h       ~       h  h  h 

T        Incr.  y  _  dy  _  du      dv      dw  ,«* 

;t=0  Incr.  x       dx  ~  dx       dx       dx 
If  some  of  the  symbols  have  a  minus,  instead  lof  a  plus,  sign  a 
corresponding  result  is  obtained.     For  instance,  if 

y  =  U-V-W-..., 

,,  dy  _  du      dv      dw 

dx  ~  dx      dx      dx  ~  " '     '  *  ' 

Hence  the  rule  :• — The  differential  coefficient  of  the  sum  or 
difference  of  any  number  of  functions  is  equal  to  the  sum  or  differ- 
ence of  the  differential  coefficients  of  the  several  functions. 

III.  The  differential  coefficient  of  the  product  of  a  variable  and 
a  constant  quantity.     Let 

y  =  axn ; 

Incr.  y  =  a(x  +  h)n  -  axn  =  anxn  "  lh  +  an\n~~    >xn  -  2h2  +  . . . 
Therefore 

^»^F=2  — ■    •    •    <s) 

Hence  the  rule : — The  differential  coefficient  of  the  product  of 
a  variable  quantity  and  a  constant  is  equal  to  the  constant  multi- 
plied by  the  differential  coefficient  of  the  variable. 

IV.  The  differential  coefficient  of  any  constant  term  is  zero. 
Since  a  constant  term  is  essentially  a  quantity  that  does  not  vary, 
if  y  be  a  constant,  say,  equal  to  a ;  then  da/dt  must  be  absolute 
zero.     Let 

y  -  (aj»  +  a) ; 
then,  following  the  old  track, 

incr.  y  =  (x  +  h)n  +  a  -  (xn  +  a) ; 

n      n-M       ,      n(n    -    1)       «-9L9      , 

.\  Incr.  y  =  -  xn    lh  +   -^— ^ — '-  xn    2h2  +  . . . 
1  A\ 

Incr.  X        dx  v   ' 

where  the  constant  term  has  disappeared. 


§  14.  THE  DIFFERENTIAL  CALCULUS.  30 

For  the  sake  of  brevity  we  have  written  1!  =  1 ;  2!  =  1  x  2 ; 
3!  =  1  x  2  x  3  ;  n\  =  1  x  2  x  3  x  . . .  x  (n  -  2)  x  (n  -  1)  x  n.  Strictly 
speaking,  0!  has  no  meaning  ;  mathematicians,  however,  find  it  con- 
venient to  make  0!  =  1.  This  notation  is  due  to  Kramp.  "  n\  " 
is  read  "  factorial  n". 

V.  The  differential  coefficient  of  a  polynomial l  raised  to  any 
'power.     Let 

y  =  (ax  +  x2)n. 

If  we  regard  the  expression  in  brackets  as  one  variable  raised  to 
the  power  of  n}  we  get 

dy  =  n(ax  +  x2)n  ~ 1  d(ax  +  x2). 
Differentiating  the  last  term,  we  get 

^  =  n{ax  +  x2)n~l(a  +  2x).  .         .         (7) 

Hence  the  rule :  —  The  differential  coefficient  of  a  polynomial 
raised  to  any  power  is  obtained  by  diminishing  the  exponent  of  the 
power  by  unity  and  multiplying  the  expression  so  obtained  by  the 
differential  coefficient  of  the  polynomial  and  the  original  exponent. 

Examples. — (1)  If  y  =  x  -  2x2,  show  that  dyjdx  =  1  -  4tx. 

(2)  If  y  =  (1  -  a;2)3,  show  that  dyjdx  =  -  6x{l-x2)2>  This  means  that 
y  changes  at  the  rate  of  -  6a-(l  -  a;2)'2  for  unit  change  of  x ;  in  other  words, 
y  changes  -  6a*(l  -  x2)2  times  as  fast  as  x. 

(3)  If  the  distance,  s,  traversed  by  a  falling  body  at  the  time  t,  is  given  by 
the  expression  s  =  \gt2,  show  that  the  body  will  be  falling  with  a  velocity 
ds/dt  =  gt,  at  the  time  t. 

(4)  Young's  formula  for  the  relation  between  the  vapour  pressure  p  and 
the  temperature  6  of  isopentane  at  constant  volume  is,  p  =  bO  -  a,  where  a 
and  b  are  empirical  constants.  Hence  show  that  the  ratio  of  the  change  of 
pressure  with  temperature  is  constant  and  equal  to  b. 

(5)  Mendeleeff's  formula  for  the  superficial  tension  s  of  a  perfect  liquid  at 
any  temperature  d  is,  s  =  a  -  bd,  where  a  and  b  are  constants.  Hence  show 
that  rate  of  change  of  s  with  Q  is  constant.     Ansr.  -  b. 

(6)  One  of  Callendar's  formulae  for  the  variation  of  the  electrical  resistance 
B  of  a  platinum  wire  with  temperature  6  is,  B  =  B0(l  +  a.6  +  &02),  where  a  and 
0  and  B0  are  constants.  Find  the  increase  in  the  resistance  of  the  wire  for  a 
small  rise  of  temperature.     Ansr.  dB  =  B0(a  +  2/30)d0. 

(7)  The  volume  of  a  gram  of  water  is  nearly  1  +  a  (0  -  4)2  ccs.  where  6 
denotes  the  temperature,  and  a  is  a  constant  very  nearly  equal  to  8*38  x  10  ~|6. 
Show  that  the  coefficient  of  cubical  expansion  of  water  at  any  temperature  6  is 
equal  to  2a(0  -  4).  Hence  show  that  the  coefficients  of  cubical  expansion  of 
water  at  0°  and  10°  are  respectively  -  67*04  x  10  -  6,  and  +  100'56  x  10  -  6. 

1 A  polynomial  is  an  expression  containing  two  or  more  terms  connected  by  plus  or 
minus  signs.     Thus,  a  +  bx  ;  ax  +  by  +  z,  etc.     A  binomial  contains  two  such  terms. 


40  HIGHER  MATHEMATICS.  §  14. 

(8)  A  piston  slides  freely  in  a  circular  cylinder  (diameter  6  in.).  At  what 
rate  is  the  piston  moving  when  steam  is  admitted  into  the  cylinder  at  the 
rate  of  11  cubic  feet  per  second?  Given,  volume  of  a  cylinder  =  irr2h.  Hint. 
Let  v  denote  the  volume,  x  the  height  of  the  piston  at  any  moment.  Hence, 
v  =  ir($)2x  ;  .*.  dv  =  ir(£)2dx.  But  we  require  the  value  of  dxfdt.  Divide  the 
last  expression  through  with  dt,  let  ir  =  ^, 

dx      dv  7        -„  ,, 

•••^=dlxl6><22  =  56ft-perSe0- 
(9)  If  the  quantity  of  heat,  Q,  necessary  to  raise  the  temperature  of  a 
gram  of  solid  from  0°  to  6°  is  represented  by  Q  =  aQ  +  &02  +  c03  (where  a,  b,  c, 
are  constants),  what  is  the  specific  heat  of  the  substance  at  9°.  Hint.  Com- 
pare the  meaning  of  dQ/dd  with  your  definition  of  specific  heat.  Ansr. 
a  +  2be  +  Scd2. 

(10)  If  the  diameter  of  a  spherical  soap  bubble  increases  uniformly  at  the 
rate  of  0*1  centimetre  per  second,  show  that  the  capacity  is  increasing  at  the 
rate  of  0'2ir  centimetre  per  second  when  the  diameter  becomes  2  centimetres. 
Given,  volume  of  a  sphere,  v  =  \tzU\ 

.-.  dv  =  frrD2dD,  .-.  dvjdt  =  £  x  tr  x  22  x  O'l  =  0-2tt. 

(11)  The  water  reservoir  of  a  town  has  the  form  of  an  inverted  conical 
frustum  with  sides  inclined  at  an  angle  of  45°  and  the  radius  of  the  smaller 
base  100  ft.  If  when  the  water  is  20  ft.  deep  the  depth-  of  the  water  is  de- 
creasing at  the  rate  of  5  ft.  a  day,  show  that  the  town  is  being  supplied  with 

water  at  the  rate  of  72,000  ir  cubic  ft.  per 
diem.  Given,  frustum,  y,  of  cone  =  frr  x 
height  x  (a2  +  ab  +  b7),  where  a,  and  b  are 
the  radii  of  the  circular  ends.  Hint.  Let  a 
(Fig.  6)  denote  the  radius  of  the  smaller  end, 
x  the  depth  of  the  water.  First  show  that 
a  +  x  is  the  radius  of  the  reservoir  at  the 

surface  of  the  water.     Hence,  y  =  \ir{{a  +  x)2  +  a(a  +  x)  +  a2)x  ;  .*.  dy  = 

ir(a2  +  lax  +  x2)dx,  etc. 

(12)  If  a,  b,  c  are  constants,  show  that  dy/dx  =  b,  when  x  =0,  given  that 
y  =  a  +  bx  +  ex2.     Hint.  Substitute  x  =  0  after  the  differentiation. 

(13)  The  area  of  a  circular  plate  of  metal  is  expanding  by  heat.  When 
the  radius  passes  through  the  value  2  cm.  it  is  increasing  at  the  rate  of  0*01 
cm.  per  second.  How  fast  is  the  area  changing  ?  Ansr.  0-047r  sq.  cm.  per 
second.  Hint.  Radius  =  x  cm.  ;  area=?/  sq.  cm. ;  .*.  area  of  circle  =  y=wx2. 
Hence,  dy\dt  =  2irx  .  dxfdt ;  when  x  =  2,  dxjdt  =  0-01,  etc. 

VI.  The  differential  coefficient  of  the  product  of  any  number  of 
functions.     Let 

y  =  uv 

where  u  and  v  are  functions  of  x.  When  x  becomes  x  +  h,  let  u,  v, 
and  y  become  ul9  vv  and  yv  Then  y1  =  ulv1 ;  yx  -  y  =  u1v1  -  uv, 
add  and  subtract  uvl  from  the  second  member  of  this  last  equation, 
and  transpose  the  terms  so  that 

Vi  ~  V  =  u(vi  -  v)  +  v^  -  u). 


§14. 


THE  DIFFERENTIAL  CALCULUS. 


41 


In  the  language  of  differentials  we  may  write  this  relation 

dy  =  d(uv)  =  udv  +  vdu.    ...         (8) 
Or,  divide  by  hx,  and  find  the  limit  when  8x  =  0,  thus, 
j,       by  _    dv        du 
x=0  8x        dx        dx9 

dx        dx  dx        dx  v  ' 

Similarly,  by  taking  the  product  of  three  functions,  say, 

y  =  uvw. 
Let  vw  =  z  ;  then  y  =  uz.     From  (8) ,  therefore 

dy  =  z.du  +  u.dz  =  vw.du  +  u.d(vw) ; 

.-.  dy  =  vw.du  +  u{w.dv  +  v.div) ; 

.*.  dy  =  vw.du  +  uw.dv  +  uv.dw,       .         .       (10) 

in  differential  notation.     To  pass  into  differential  coefficients,  divide 

by  dx.     This  reasoning  may  obviously  be  extended  to  the  product 

of  a  greater  number  of  functions. 

Hence  the  rule  : — The  differential  coefficient  of  any  number  of 
functions  is  obtained  by  multiplying  the  differential  coefficient  of 
each  separate  function  by  the  product  of  all  the  remaining  func- 
tions and  then  adding  up  the  results. 

Examples. — (1)  If  the  volume,  v,  of  gas  enclosed  in  a  vessel  at  a  pressure 
p,  be  compressed  or  expanded  without  loss  of  heat,  it  is  known  that  the 
relation  between  the  pressure  and  volume  is  pvy  =  constant ;  y  is  also  a 
constant.     Hence,  prove  that  for  small  changes  of  pressure,  dvfdp  =  -  vjyp. 

(2)  liy  =  {x-  1)  {x  -  2)  (x  -  3),  dy/dx  =  Bx2  -  12x  +  11. 

(3)  liy  =  x2(l  +  ax2)  (1  -  ax2),  dy/dx  =  2x  -  6a2x5. 

(4)  Show  geometrically  that  the  differential  of  a  small  increment  in  the 
capacity  of  a  rectangular  solid  figure  whose  unequal  sides  are  x,  y,  z  is 
denoted  by  the  expression  xydz  +  yzdx  +  zxdy.  Hence,  show  that  if  an  ingot 
of  gold  expands  uniformly  in  its  linear  dimensions  at  the  rate  of  0*001  units 
per  second,  Its  volume,  v,  is  increasing  at  the  rate  of  dvfdt  =  0-110  units  per 
second,  when  the  dimensions  of  the  ingot  are  4  by  5  by  10  units. 

The  process  may  be  illustrated  by  a  geometrical  figure  similar 

to  that  of  page  31.      In  the       &  dx 

rectangle  (Fig.  7)  let  the  un- 


equal sides  be  represented  by 
x  and  y.     Let  x  and  y  be  in- 
creased  by  their  differentials   «V 
dx  and  dy.     Then  the  incre- 
ment of  the  area  will  be  re-  Fig.  7. 
presented  by  the  shaded  parts,  which  are  in  turn  represented  by 


42  HIGHER  MATHEMATICS.  §  14. 

the  areas  of  the  parallelograms  xdy  +  ydx  +  dxdy,  but  at  the  limit 
dx.dy  vanishes,  as  previously  shown. 

VII.  The  differential  coefficient  of  a  fraction,  or  quotient.     Let 

u 
y  =  — , 

where  u  and  v  are  functions  of  x.     Hence,  u  =  vy,  and  from  (9) 

du  =  vdy  +  ydv;  .-.  du  =  vd(-)  +  -dv9 

on  replacing  y  by  its  value  ujv.     Hence,  on  solving, 

jfu\              ~  v                1fu\       vdu  -  udv  ,:__ 

d(v)  -  —^—  '  •••   %)  =  — "P •       •       <"> 

in  the  language  of  differentials ;  or,  dividing  through  with  dx  we 
obtain,  in  the  language  of  differential  coefficients, 

du  dv 
dy  vfa  ~  Udx' 
dx~         v*  ...         (14) 

In  words,  to  find  the  differential  coefficient  of  a  fraction  or  of  a 
quotient,  subtract  the  product  of  the  numerator  into  the  differ- 
ential coefficient  of  the  denominator,  from  the  product  of  the 
denominator  into  the  differential  coefficient  of  the  numerator,  and 
divide  by  the  square  of  the  denominator. 

A  special  case  occurs  when  the  numerator  of  the  fraction  a/x 
is  a  constant,  a,  then 

a     j        x  .da  -  a  .dx       -  a  .dx         dy  a       , .  • 

y~x'dV ? --S----S"  -3-      <13> 

In  words,  the  differential  coefficient  of  a  fraction  a/x  whose 
numerator  is  constant  is  minus  the  constant  divided  by  the  square 
of  the  denominator. 

Examples. — (1)  If  y  =  x/(l  -  x) ;  show  that  dyfdx  =  1/(1  -  x)2. 

(2)  If  x  denotes  the  number  of  gram  molecules  of  a  substance  A  trans- 
formed by  a  reaction  with  another  substance  B,  at  the  time  t,  experiment 
shows  that  xj(a  -  x)  =  akt,  when  k  is  constant.  Hence,  show  that  the 
velocity  of  the  reaction. is  proportional  to  the  amounts  of  A  and  B  present  at 
the  time  t.  Let  a  denote  the  number  of  gram  molecules  of  A,  and  of  B 
present  at  the  beginning  of  the  reaction.  Hint.  Show  that  the  velocity  of 
the  reaction  is  equal  to  k(a  -  x)2,  and  interpret  the  result. 

(3)  If  y  =  (1  +  s2)/(l  -  x2),  show  that  dyfdx  =  4sc/(l  -  x2)\ 

(4)  If  y  =  a/a;w,  show  that  dyfdx  =  -  nafxn  + 1. 


§  14.  THE  DIFFERENTIAL  CALCULUS.  43 

(5)  The  refractive  index,  /*,  of  a  ray  of  light  of  wave-length  A  is,  according 
to  Christoffel's  dispersion  formula 

H  =  p0J2/{*jr+~\J\  +  \/l  -  a0/a}> 
where  ^  and  A0  are  constants.     Find   the  change  in  the  refractive  index 
corresponding  to  a  small  change  in  the  wave-length  of  the  light.      Ansr. 
dfifd\  =  -  ^3a02/{2a3/x0V(1  -  V/*2)K     It;  is  not  often  so  difficult  a  differentia- 
tion occurs  in  practice.     The  most  troublesome  part  of  the  work  is  to  reduce 

djL_  s/frW  n/(1  +  Ap/X)  -  V(t  -  A0/a)}/A* 

dK  -  "  V(l  -"AoW{  V(l  +  Afl/A)  +  V(i  -  a0/aH2' 
to  the  answer  given.  Hint.  Multiply  the  numerator  and  denominator  of  the 
right  member  with  the  factor  4yu02(  sll  +  aJa  +  s/l  -  a0/a).  and  take  out  the 
terms  which  are  equal  to  /x  of  the  original  equation  to  get  /**.  Of  course  the 
student  is  not  using  this  abbreviated  symbol  of  division.  See  footnote,  page 
14.  I  recommend  the  beginner  to  return  to  this,  and  try  to  do  it  without  the 
hints.     It  is  a  capital  exercise  for  revision. 

VIII.  The  differential  coefficient  of  a  function  affected  with  a 
fractional  or  negative  exponent.  Since  the  binomial  theorem  ia 
true  for  any  exponent  positive  or  negative,  fractional  or  integral, 
formula  (2)  may  be  regarded  as  quite  general.  The  following  proof 
for  fractional  and  negative  exponents  is  given  simply  as  an  exercise. 
Let 

y  =  xn. 

First.  When  n  is  a  positive  fraction.  Let  n  =  p/q,  where  p 
and  q  are  any  integers,  then 

y  =  xq (14) 

Raise  each  term  to  the  5th  power,  we  obtain  the  expression  yq  =  xp. 
By  differentiation,  using  the  notation  of  differentials,  we  have 

qy9  ~  ldy  =  pxp  ~  ldx. 
Now   raise   both   sides   of   the   original   expression,   (14),   to   the 
(q  -  l)th  power,  and  we  get 

P7-P 

%f  -  l  =  x    q    . 
Substitute  this  value  of  y7  ~  l  in  the  preceding  result,  and  we  get 

dy       p  xp~1xplq  dy      p  f _1  /Jn, 


which  has  exactly  the  same  form  as  if  n  were  a  positive  integer. 

Second.     When  n  is  a  negative  integer  or  a  negative  fraction. 
Let 

y  =  x~n; 
then  y  —  l/xn.     Differentiating  this  as  if  it  were  a  fraction,  (13) 


44  HIGHER  MATHEMATICS.  §  14. 

above,  we  get  dyldx  =  -  nxn  ~  1/x2n,  which  on  reduction  to  its 
simplest  terms,  assumes  the  form 

dy_d(x-)_      TO_n_1- 
dx  dx 

Thus  the  method  of  differentiation  first  given  is  quite  general. 
A  special  case  occurs  when  y  =  Jx,  in  that  case  y  =  xh- ; 

^=^  =  J_=i*-*.         .         .         (16) 
dx        dx        2Jx      2  V     ; 

In  words,  the  differential  coefficient  of  the  square  root  of  a  variable 
is  half  the  reciprocal  of  the  square  root  of  the  variable. 

Examples. — (1)  Matthiessen's  formula  for  the  variation  of  the  electrical 
resistance  R  of  a  platinum  wire  with  temperature  0,  between  0°  and  100°  is 
R  —  jR0(1  -  ad  +  be2)  ~  K  Find  the  increase  in  the  resistance  of  the  wire  for 
a  small  change  of  temperature.  Ansr.  dRjde  =  R2(a  -  2b6)jRQ.  Note  a  and 
b  are  constants  ;  dR  =  -  R0(l  -  a6  +  bff2)  ~2d(l  -  ad  +  b82)  ;  multiply  and 
divide  by  R0  ;  substitute  for  R  from  the  original  equation,  etc 

(2)  Siemens'  formula  for  the  relation  between  the  electrical  resistance  of 
a  metallic  wire  and  temperature  is,  R  =  R0{1  +  ae  +  b  sfe)  Hence,  find  the 
rate  of  change  of  resistance  with  temperature.     Ansr.  R0{a  +  \be  ~  h). 

(3)  Batschinski  (Bull.  Soc.  Imp.  Nat.  Moscow,  1902)  finds  that  the  pro- 
duct 7](e  +  273)3  is  constant  for  many  liquids  of  viscosity  77,  at  the  temperature 
e.     Hence,  show  that  if  A  is  the  constant,  drj/de  =  -  3^/(0  +  273). 

(4)  Batschinski  (I.  c.)  expresses  the  relation  between  the  "  viscosity  para- 
meter," 7],  of  a  liquid  and  the   critical  temperature,  6,  by  the  expression 

M^e^rjmJ  =  B,  where  B,   M,  and  m  are  constants.       Hence  show  that 

d-nfde  =  -  &n[e. 

IX.    The  differential  coefficient  of  a  function  of  a  function. 

Let 

u  =  <f>(y) ;  and  y  =  f(x). 

It  is  required  to  find  the  differential  coefficient  of  u  with  respect  to 
x.  Let  u  and  y  receive  small  increments  so  that  when  u  becomes 
uv  y  becomes  yx  and  x  becomes  xv     Then 

%  ~  u  =  u>i-u    Vi~  y 

xi  -  x    V\  -  y  '  xi  -  x' 

which  is  true,  however  small  the  increment  may  be.  At  the  limit, 
therefore,  when  the  increments  are  infinitesimal 

du  _  du    dy 

dx~~~3y'dx  '  '  '  *  (17> 
I  may  add  that  we  do  not  get  the  first  member  by  cancelling  out 
the  dy's  of  the  second      The  operations  are 


§  14.  THE  DIFFERENTIAL  CALCULUS.  45 

In  words,  (17)  may  be  expressed :  the  differential  coefficient  of  a 
function  with  respect  to  a  given  variable  is  equal  to  the  product  of 
the  differential  coefficient  of  the  function  with  respect  to  a  second 
function  and  the  differential  coefficient  of  the  second  function  with 
respect  to  the  given  variable.  We  can  get  a  physical  meaning  of 
this  formula  by  taking  x  as  time.  In  that  case,  the  rate  of  change 
of  a  function  of  a  variable  is  equal  to  the  product  of  the  rate  of 
change  of  that  function  with  respect  to  the  variable,  and  the  rate 
of  change  of  the  variable. 

The  extension  to  three  or  more  variables  will  be  obvious.     If 
u  =  <f}(w),  w  =  \p{y),  y  =  f(x),  it  follows  that 
du      du   dw   dy 

=  , .  _£  (181 

dx      dw    dy   dx  v     ; 

With  the  preceding  notation,  it  is  evident  that  the  relation 

yi  -  y'x1-  x 
is  true  for  all  finite  increments,  we  assume  that  it  also  holds  when 
the  increments  are  infinitely  small ;  hence,  at  the  limit, 

dx  dy  <te_±  mq\ 

dy'dx"  L>  0T>  dy~  dy  '         <iy' 

dx 

We  have  seen  that  if  y  is  a  function  of  x  then  #  is  a  function 

of  y  ;    the   latter,  however,   is   frequently  said  to  be  an   inverse 

function  of  the  former,  or  the  former  an  inverse  function  of  the 

latter.     This  is  expressed  as  follows  :  If  y  =  f(x),  then  x  =  f~\y), 

or,  \ix=  f(y),  then  y  =f~\x). 

Examples.— (1)  If  y  =  xn/(l  +  x)n,  show  that  dy/dx  =  nxn  -  7(1  +  x)n  + l. 

(2)  If  y  =  1/^/(1  -  z2),  show  that  dy/dx  =  x/ ^(1  -  jb2)3. 

(3)  The  use  of  formula  (17)  often  simplifies  the  actual  process  of  differentia- 
tion ;  for  instance,  it  is  required  to  differentiate  the  expression  u  =  J  (a2  -  x2). 
Assume y  =  a2-  x2.  Then, u  =  \/y,y  =  a2-x2;  &n& dy/dx=  -2x;  du/dy=^y~i, 
from  (16) ;  hence,  from  (17),  du/dx=  -x(a?-x2)~%.  This  is  an  easy  example 
which  could  be  done  at  sight ;  it  is  given  here  to  illustrate  the  method. 

By  the  application  of  these  principles  any  algebraic  function 
which  the  student  will   encounter  in   physical  science,1  may  be 

1  K.  Weierstrass  has  shown  that  there  are  some  continuous  functions  which  have 
not  yet  been  differentiated,  but,  as  yet,  they  have  no  physical  application  except 
perhaps  to  vibrations  of  very  great  velocity  and  small  amplitude.  See  J.  JJarkness 
and  F.  Morley's  Theory  of  functions^  London,  65,  1893, 


46  HIGHER  MATHEMATICS.  §  15. 

differentiated.  Before  proceeding  to  transcendental  functions,  that 
is  to  say,  functions  which  contain  trigonometrical,  logarithmic  or 
other  terms  not  algebraic,  we  may  apply  our  knowledge  to  the 
well-known  equations  of  Boyle  and  van  der  Waals. 

§  15.  The  Gas  Equations  of  Boyle  and  van  der  Waals. 

In  van  der  Waals'  equation,  at  a  constant  temperature, 

*  [P  +  f§)  (V  -  b)  =  constant,    .  .  .  .(1) 

where  b  is  a  constant  depending  on  the  volume  of  the  molecule,  a 
is  a  constant  depending  on  intermolecular  attraction.  Differenti- 
ating with  respect  to  p  and  v,  we  obtain,  as  on  pages  40  and  41, 

(v  -  b)d(p  +  £2)  +  (p  +  fyd(v  -  b)  =  0, 

and  therefore 

dv  v  -  b 


dp  a       2ab 


(2) 


The  differential  coefficient  dv/dp  measures  the  compressibility  of 
the  gas.  If  the  gas  strictly  obeyed  Boyle's  law,  a  =  b  =  0,  and 
we  should  have 

—  =  -  -  (S) 

dp  p'  v  ' 

The  negative  sign  in  these  equations  means  that  the  volume  of 
the  gas  decreases  with  increase  of  pressure.  Any  gas,  therefore, 
will  be  more  or  less  sensitive  to  changes  of  pressure  than  Boyle's 
law  indicates,  according  as  the  differential  coefficient  of  (2)  is 
greater  or  less  than  that  of  (3),  that  is  according  as 

v  -  b        ^v                      ,_            a      2ab  ,  _  a      lab 

pv  -  pb>pv +— o-;    .'.jpb%  — 


_  a_      2ab^p'   *   r        *^"r      v   '    v2   '        r    ^v        v*   ' 

V      v2  +    vz 

a      2a 
•'•^56-V (*) 

If  Boyle's  law  were  strictly  obeyed. 

pv  =  constant,        ....  (5) 

but  if  the  gas  be  less  sensitive  to  pressure  than  Boyle's  law 
indicates,  so  that,  in  order  to  produce  a  small  contraction,  the 
pressure  has  to  be  increased  a  little  more  than  Boyle's  law 
demands,  pv  increases  with  increase  of  pressure ;  while  if  the  gas 


§16.  THE  DIFFERENTIAL  CALCULUS.  47 

be  more  sensitive  to  pressure  than  Boyle's  law  provides  for,  pv 
decreases  with  increase  of  pressure. 

Some  valuable  deductions  as  to  intermolecular  action  have  been 
drawn  by  comparing  the  behaviour  of  gases  under  compression  in 
the  light  of  equations  similar  to  (4)  and  (5).  But  this  is  not  all. 
From  (5),  if  c  =  constant,  v  =  c/p,  which  gives  on  differentiation 

dv  c 

dp  =  ~  p* 

or  the  ratio  of  the  decrease  in  volume  t .  ...  of  pressure,  is 

inversely  as  the  square  of  the  pressure.     By  substituting  p  =  2,  3, 

4,  ...  in  the  last  equation  we  obtain 

dv  _  1     1     JL_ 

dp  ~  4 '  9 '  16 ;  ' ' ' 
where  c  =  unity.      In  other  words,  the  greater  the  pressure  to 
which  a  gas  is  subjected  the  less  the  corresponding  diminution  in 
volume  for  any  subsequent  increase  of  pressure.     The  negative 
sign  means  that  as  the  pressure  increases  the  volume  decreases. 

§  16.  The  Differentiation  of  Trigonometrical  Functions. 

Any  expression  containing  trigonometrical  ratios,  sines,  cosines, 
tangents,  secants,  cosecants,  or  cotangents  is  called  a  trigono- 
metrical function.  The  elements  of  trigonometry  are  discussed 
in  Appendix  I.,  on  page  606  et  seq.,  and  the  beginner  had  better 
glance  through  that  section.  We  may  then  pass  at  once  in  medias 
res.     There  is  no  new  principle  to  be  learned. 

I.  The  differential  coefficient  of  sin  x  is  cos  x.  Let  y  become  yv 
when  x  changes  to  x  +  h,  consequently, 

y  =  sin  x ;  and  yx  =  sin  (x  +  h) ;  .'.  yx  -  y  =  sin  (x  +  h)  -  sin  x. 
By  (39),  page  612, 

Vi  ~  V  =  2 sin g  cos  (x  +  ^j. 
Divide  by  h  and 

sin  x 

But  the  value  of approaches  unity,  page  611,  as  x  approaches 

x 


h  \h 

.  sin  a: 

of 

x 

zero,  therefore, 


Vi  -  V  dy      dismx) 

Ut.-Pt1  - ooax<  •■•!" -V- =  C0BX  ■    (li 


48  HIGHER  MATHEMATICS.  §  16. 

The  rate  of  change  of  the  sine  of  an  angle  with  respect  to  the 
angle  is  equal  to  the  cosine  of  the  angle.  When  x  increases  from 
0  to  Jtt,  the  rate  of  increase  of  sine  x  is  positive  because  cos  x  is 
then  positive,  as  indicated  on  page  610 ;  and  similarly,  since  cos  x 
is  negative  from  ^ir  to  7r,  as  the  angle  increases  from  ^ir  to  irt  sine  x 
decreases,  and  the  rate  of  increase  of  sin  x  is  negative. 
If  x  is  measured  in  degrees,  we  must  write 

d(smx°)  _  <*(siniJ5^C)       _tt_        ttV        _*■_ 

dx  dx  180 C0S  180      180  cos  x  ' 

since  the  radian  measure  of  an  angle  =  angle  in  degrees  x  T|o7r, 
where  it  =  3-1416,  as  indicated  on  page  606. 

Numerical  Illustration. — You  can  get  a  very  fair  approximation  to 
the  fact  stated  in  (1),  by  taking  h  small  and  finite.  Thus,  if  x  =  42°  6' ; 
and  h  =  V  ;  x  +  h  =  42°  7' ; 

.  Incr.  y  _  sin  (x  +  h)  -  sin  x  _  Q-Q002158 
*  .*  Incr.  x  ~         hih  radians        ~  0-0002909  =  °'74183* 
But  cos  x  =  0-74198 ;  cos  (x  +  h)  =  0-74178  ;  so  that  when  h  is  -^°,  dyjdx  lies 
somewhere  between  cos  a;  and  cos  (a;  +  h).      By  taking  smaller  and  smaller 
values  of  h,  dy/dx  approaches  nearer  and  nearer  in  value  to  cos  x. 

II.  The  differential  coefficient  of  cos  x  is  -  sinx.  Let  us  put 
y  =  cos  x ;  and  yx  =  cos  (x  +  h) ;  yl  -  y  =  cos  (x  +  h)  -  cos  x. 
From  the  formula  (41)  on  page  612,  it  follows  that 

«'.-£.'/         h\         y,  -  y          sinhh  .    /         h\ 
y1  -  y  =  -  2  sin  ^  sin  [x  +  ^J  ;  or       h      = IfiT  sm  \x  +  2)  ' 

and  at  the  limit  when  h  =  0, 

Vi  -  V  •  %      c?(cosic) 

Lt„  =  0— 5- --am*;  a^-  -^- -  -  ana  (2) 

The  meaning  of  the  negative  sign  can  readily  be  deduced  from 
the  definition  of  the  differential  coefficient.  The  differential  co- 
efficient of  cos  x  with  respect  to  x  represents  the  rate  at  which 
cosa:  increases  when  x  is  slightly  increased.  The  negative  sign 
shows  that  this  rate  of  increase  is  negative,  in  other  words,  cos  x 
diminishes  as  x  increases  from  0  to  ^tt.  When  x  passes  from  ^tt  to 
7r  cos  x  increases  as  x  increases,  the  differential  coefficient  is  then 
positive. 

III.  The  differential  coefficient  of  tan  x  is  sec2x.  Using  the  re- 
sults already  deduced  for  sin  x  and  cos  x,  and  remembering  that 
sin  #/cos  x  is,  by  definition,  equal  to  tan  x,  let  y  =  tan  x,  then 


§  17.  THE  DIFFERENTIAL  CALCULUS.  49 

,/sin#\              d(smx)        .       d(cosx) 
ilL       x       d\  — —  I     cosoj-1-^ — -  -  sin#— ^ — -  9         .  9 

d(ta,nx)  _     Vcosic/ cfa  <ja;       _  cos2a;  +  sm2a; 

dx  dx  cos2#  cos2# 

But  the  numerator  is  equal  to  unity  (19),  page  611.     Hence 

ri(tan«)_      1     =sec2g-      ...         (3) 
a#  cos2# 

In  the  same  way  it  can  be  shown  that 

-i-__ — '  =  -  cosec2#.  (3) 

arc 

The  remaining  trigonometrical  functions  may  be  left  for  the 

reader  to  work  out  himself.     The  results  are  given  on  page  193. 

Examples. — (1)  If  y  =  cosn#  ;  dy\dx  —  -  n  cosw  —  xx  .  sin  x. 

(2)  If  y  =  sinnx ;  dy/dx  =  n  sinn  —  lx .  cos  x. 

(3)  If  a  particle  vibrates  according  to  the  equation  y  =  a  sin  (qt  -  e),  what 
is  the  velocity  at  any  instant  when  a,  q  and  c  are  constant  ?  The  answer  is 
aq  cos  (qt  -  e). 

(4)  If  y  =  sin2(naj  -  a) ;  dy/da;  =  2n  sin  (nx  -  a)  cos  {nx  -  a). 

(5)  Differentiate  tan  6  =  y/a?.    Ansr.  d$  =  (xdy  -  ydx)  -f  (a:2  4-  y2).    Hint. 

sec20 .  de  =  (1  +  tan20)tf0  ; .-.  ^_tde=xdy  -Vdx^  etc< 

(6)  If  the  point  P  moves  upon  a  circle,  with  a  centre  0  and  radius  18  cm., 
AB  is  a  diameter  ;  MP  is  a  perpendicular  upon  ^IP,  show  that  the  speed  of  M 
on  the  line  AB  is  22G  cm.  per  second  when  the  angle  BOP  =  a  =  30°  ;  and 
P  travels  round  the  perimeter  four  times  a  second.  Sketch  a  diagram.  Here 
OM  =  y  =  r  cos  a  =  18  cos  a ;  .'.  dyjdt  =  -  (18  sin  a)dafdt.  But  dajdt  =  4  x  2?r 
since  7r  =  half  the  circumference  ;  sin  30°  =  £  ; 

.-.  dyjdt  =  -  18  x  £  x  8tt  =  -9x8x3-  1416  =  -  226  cm.  per  sec.  (nearly). 

§  17.    The   Differentiation  of  Inverse    Trigonometrical 
Functions.     The   Differentiation  of  Angles. 

The  equation,  sin  y  =  x,  means  that  y  is  an  angle  whose  sine  is 
x.     It  is  sometimes  convenient  to  write  this  another  way,  viz., 

sin  ~  lx  =  y, 
meaning  that  sin  ~  lx  is  an  angle  whose  sine  is  x.  Thus  if  sin  30°  =  J, 
we  say  that  30°  or  sin_1i  is  an  angle  whose  sine  is  \.  Trigono- 
metrical ratios  written  in  this  reverse  way  are  called  inverse 
trigonometrical  functions.  The  superscript  "  -  1  "  has  no  other 
signification  when  attached  to  the  trigonometrical  ratios.  Note,  if 
tan  45°  =  1,  then  tan_1l  =  45°  ;  .*.  tan  (tan_1l)  =  tan  45°. 
Some  writers  employ  the  symbols  arc  sin  x  ;  arc  tan  x  ;  . . .  for  our 
sin  -  lx  ;  tan  ~  lx ; , . . 

D 


50  HIGHER  MATHEMATICS.    .  §  17. 

The  differentiation  of  the  inverse  trigonometrical  functions  may 
be  illustrated  by  proving  that  the  differential  coefficient  of  sin  ~  lx  is 
1/  J(l  -  x2).     If  y  =  sin  ~  lx,  then  sin  y  =  x,  and 

dx  dy  1 

=  cos  y  ;  or  -/  = , 

ay  ax      cos  y 

But  we  know  from  (19),  page  611,  that 

cos2y  4-  sin'2?/  =  1 ;  or  cos  y=±  .  •'<  '•  -  sin2?/)  =  ±  J(l  -  a?2), 

for  by  hypothesis  sin  y  =  x.     Hence 

d(sin  ~lx)  _  dy  1  1 

dx  dx  ~  cos  y~~  J\  -  x2 

The  fallacy  mentioned  on  page  5  illustrates  the  errors  which 
might  enter  our  work  unsuspectingly  by  leaving  the  algebraic  sign 
of  a  root  extraction  undetermined.  Here  the  ambiguity  of  sign 
means  that  there  are  a  series  of  values  of  y  for  any  assigned  value 
of  x  between  the  limits  +  1.  Thus,  if  n  is  a  positive  integer,  we 
know  that  sin  x  =  sin  {nir  ±  x)  j  the  +  sign  obtains  if  n  is  even ; 
the  negative  sign  if  n  is  odd.  This  means  that  if  x  satisfies  sin  ~  }y, 
bo  will  tv  ±  x  ;  2tt  ±  x  ;  . .  .  If  we  agree  to  take  sin  "  ly  as  the  angle 
between  -  \tt  and  +  \ir ,  then  there  will  be  no  ambiguity  because 
cos  y  is  then  necessarily  positive.  The  differential  coefficient  is 
then  positive,  that  is  to  say, 

d(sin  ~lx)  _  1 


dx  J(l  -  x2) 

Similarly, 


(1) 


•rfCoOB-ls)^  1        _        /  1  \  1 

dx  sin  y         \±  J±  _  x2j         Jl  -  x2' 

The  ambiguity  of  sign  is  easily  decided  by  remembering  that  sin  y 
is  positive  when  y  lies  between  rr  and  0.  Again,  if  y  =  tan  ~  1xf 
x  =  tan  y,  dx/dy  =  1/cos2?/.  But  cos2?/  =  1/(1  +  tan2?/)  =  1/(1  +  x2) 
(page  612).     Hence 

d(tan  ~lx)  9  1 

The  differential  coefficient  of  tan-]#  is  an  important  function, 
since  it  appears  very  frequently  in  practical  formulae.  It  follows  in 
a  similar  manner  that 

dx  1  +  x2'  w 


§  18.  THE  DIFFEEENTIAL  CALCULUS.  51 

The  remaining  inverse  trigonometrical  functions  may  be  left  as  an 
exercise  for  the  student.     Their  values  will  be  found  on  page  193. 

Examples.— (1)  Differentiate  y  =  sin-1  \xj>J(l  +  x2)].  Sin  y=xl\/l  +  x* 
hence  cos  ydy  =  dxj(l  +  x2)i.  But  cos  y  =  ,J(1  -  sin2?/)  =  ^/[l  -  x2/(l  +  a;2)]. 
Substituting  this  value  of  cos  y  in  the  former  result  we  get,  on  reduction, 
dy/dx  =  (1  +  x2)  - l,  the  answer  required.  Note  the  steps  : 
(1  +  a2)  "  \dx  -  x2(l  +  x2  ~  Ux  =  (1  +  x2)  ~  f  (1  +  cc2  -  x2)dx,  etc.  Also 
cos  y  x  (1  +  z2)f  =  (1  +  a2)  "  i  (1  +  s2)S  =  (1  +  a;2). 

(2)  If  y  =  sin  -  ^ ;  dy/dte  =  2<r(l  -  a;4)  ~  i 

(3)  If  V  =  tan  -  »        g        ■  ^jL=       *  See  formula  (22),  page  612. 

§  18.  The   Differentiation  of  Logarithms. 

Any  expression  containing  logarithmic  terms  is  called  a  logar- 
ithmic function.  E.g.,  y  =  logx  +  xz.  To  find  the  differential 
coefficient  of  log  x.     Let 

y  =  logx;  and  yx  =  log(a;  +  h). 
Where  y1  denotes  the  value  of  y  when  x  is  augmented  to  a;  +  h. 
By  substitution, 

Vi  ~  V  _  log(a?  +  fe)  -  loga?. 

but  we  know,  page  26,  that  log  a  -  log  b  =  logr,  therefore 
Incr.  y       1.     /x  +  h\       1 .  ■"'  /,        &\ 

and  t  =  ^log(l  +  g.    .         .         .         (1) 

The  limiting  value  of  this  expression  cannot  be  determined  in  its 
present  form  by  the  processes  hitherto  used,  owing  to  the  nature 
of  the  terms  1/h  and  h/x.  The  calculation  must  therefore  be  made 
by  an  indirect  process.     Let  us  substitute 

h      1 1    m  1  _  u 

x      u'  '  ' h      x 

•"•  \  l0"(1  +  i)  =  I  ■  «Io§(1  +  i) ;-.!  ■  to8(l  +  | 

As  A  decreases  %  increases,  and  the  limiting  value  of  u  when 
A  becomes  vanishingly  small,  is  infinity.  The  problem  now  is  to 
find  what  is  the  limiting  value  of  log(l  +  u  ~  l)u  when  u  is  infinitely 
great.  In  other  words,  to  find  the  limiting  value  of  the  above 
expression  when  u  increases  without  limit. 


52  HIGHER  MATHEMATICS.  §  18. 

-1=^4^  +  3"  •    •   (*> 

According  to  the  binomial  theorem,  page  36, 
1\M      ,       u     1       u(u  -  1)      1 

dividing  out  the  u's  in  each  term,  and  we  get 

1  +  iT  .,(-D,(-S(-S, 

f  u)  -2  +  — 2|—  + 3! +  '-- 

The  limiting  value  of  this  expression  when  u  is  infinitely  great  is 
evidently  equal  to  the  sum  of  the  infinite  series  of  terms 

1+I  +  2!+3!  +  i!  +  -"t0  infinifcy-  •         (3) 

Let  the  sum  of  this  series  of  terms  be  denoted  by  the  symbol  e. 
By  taking  a  sufficient  number  of  these  terms  we  can  approximate 
as  close  as  ever  we  please  to  the  absolute  value  of  e.  If  we  add 
together  the  first  seven  terms  of  the  series  we  get  2-71826 

1  +  |        =  2-00000 

J,  =  0-50000 

h  =  i'h  =  0-16667 


I  •  ± .--  0-04167 


l. 

f.-  J-h  =  0-00833 
|l  =  if,.=  0-00139 
^  =  A  •  i  =  0-00020 


Sum  of  first  seven  terms  =  2*71826 

The  value  of  e  correct  to  the  ninth  decimal  place 

e  =  2-718281828  . . . 
This  number,  like  tr  =  3'14159265  . . .,  plays  an  important  role  in 
mathematics.    Both  magnitudes  are  incommensurable  and  can  only 
be  evaluated  in  an  approximate  way. 

Returning  now  to  (2),  it  is  obvious  that 

dy      d(logx)       1 

Tx=<te-  =  xl°Ze      •         v        •         W 

This   formula  is  true  whatever  base  we  adopt  for  our  system  of 

logarithms.     If  we  use  10,  log10e  =  0*43429  . . .  =  (say)  M, 

:.  dy      d(\oswx)      M  /c. 

and  -f-  =    v   -T10  l  =  — .  .        .        (5) 

dx  dx  x 


§  18,  THE  DIFFERENTIAL  CALCULUS.  53 

Sinco  logaa  =  1,  from  (6),  page  27,  we  can  put  expression  (4)  in  a 
much  simpler  form  by  using  a  system  of  logarithms  to  the  base  e, 
then 

dy  =  d(log^)  =  1  : 

dx  dx  x  •         \  ) 

Continental  writers  variously  use  the  symbols  L,  I,  In,  lg,  for 
"log"  and  "log  nep " ;  "  nat  log,"  or  "hyp  log,"  for  "log/'. 
"  Nep  "  is  an  abbreviation  for  "  Neperian,"  a  Latinized  adjectival 
form  of  Napier's  name. — J.  Napier  was  the  inventor  of  logarithmic 
computation.     You  will  see  later  on  where  "  hyp  log  "  comes  from. 

Examples. — (1)  If  y  =  log  ax*,  show  that  dyjdx  =  4/rc. 
(fi)  If  y  =  xn  log  x,  show  that  dyjdx  =  «»  — *(1  +  n  log  x). 

(3)  What  is  meant  by  the  expression,  2-7182S"  *  2'3026  =  io»  ?  Ansr.  If 
n  is  a  common  logarithm,  then  n  x  2-3026  is  a  natural  logarithm.  Note, 
e  =  2-71828. 

(4)  A.  Dupre  (1869)  represented  the  relation  between  the  vapour  pressures, 
p,  of  a  substance  and  the  absolute  temperature  T  by  the  equation 

a      ..-.-,  d(log  p)     A  +  BT 


log  P  =  7/t  +  b  lo8  T  +  e. 


a  result  resembling  van't  Hoff's  well-known  equation.     Hence  show  that  if 
a,b,c,A,B  are  all  constants,  dpJdT  =  p(A  +  BT)/T*. 

In  seeking  the  differential  coefficient  of  a  complex  function 
containing  products  and  powers  of  polynomials,  the  work  is  often 
facilitated  by  taking  the  logarithm  of  each  member  separately  be- 
fore differentiation.  The  compound  process  is  called  logarithmic 
differentiation. 

Examples. — (1)  Differentiate  y  =  xn/(l  +  x)n. 
Here  log  y  =  n  log  x  -  n  log  (1  +  x),   or  dy\y  =  ndxlx(l  +  x).       Hence 
dy/dx  =  ynfx(l  +  x)  =  nxn~1l(l  +  x)n+K 

(2)  Differentiate  ar*(l  +  x)nj(x3  -  1). 

Ansr.  {(n  +  l)x*  +  x?  -  (n  +  4)jb  -  4}z3(l  +  x)«  -  l(x*  -  1)  - 2. 

(3)  Establish  (10),  page  41,  by  log  differentiation.  In  the  same  way, 
show  that  d(xyz)  =  yzdx  +  zxdy  +  xydz. 

(4)  If  y  =  x{a2  +  x2)  Ja?  -  x*;  dy/dx  =  (a4  +  aW  -  4z4)(a2  -  x2) -J. 

(5)  If  y  =  log  sin  x ;  dyjdx  =  d(sin  a;)/sin  x  =  cot  x. 

(6)  How  much  more  rapidly  does  the  number  x  increase  than  its  log- 
arithm ?  Here  d(log  x)jdx  =  1/x.  The  number,  therefore,  increases  more 
rapidly  *or  more  slowly  than  its  logarithm  according  as  x  >  or  <  1.  If 
x  =  1,  the  rates  are  the  same.  If  common  logarithms  are  employed,  M  will 
have  to  be  substituted  in  place  of  unity.     E.g.,  d(\&g10x)dx  =  M/x. 

(7)  If  the  relation  between  the  number  of  molecules  a;  of  substances  A  and 
B  transformed  in  the  chemical  reaction  :  A  +  B  =  C  +  D,  and  the  time  t  be 


54  HIGHER  MATHEMATICS.  §  19. 

represented  by  the  equation 

where  k  is  constant,  and  a  and  b  respectively  denote  the  amounts  of  A  and  B 
present  when  t  =  0,  show  that  the  velocity  of  the  reaction  is  proportional  to 
the  amounts  of  A  and  B  actually  present  at  the  time  t.  Hint.  Show  that  the 
velocity  of  the  reaction  is  proportional  to  (o  -  x)(b  -  x)  and  interpret. 

§  19.  The  Differential  Coefficient  of  Exponential  Functions. 

Functions  in  which  the  variable  quantity  occurs  in  the  index 
are  called  exponential  functions.  Thus,  ax,  ex  and  (a  +  x)x  are 
exponential  functions.  A  few  words  on  the  transformation  of 
logarithmic  into  exponential  functions  may  be  needed.  It  is  re- 
quired to  transform  log  y  =  ax  into  an  exponential  function. 
Eemembering  that  log  a  to  the  base  a  is  unity,  it  makes  no 
difference  to  any  magnitude  if  we  multiply  it  by  such  expressions 
as  logaa, ;  log1010 ;  and  logee.  Thus,  since  loge(eaz)  =  ax  logBe ;  if 
l°ge2/  =  ax>  we  can  write 

\ogey  =  ax  logee  =  logee°*;  .*.  y  =  eax, 
when  the  logarithms  are  removed.     In  future  "  log  "  will  generally 
be  written  in  place  of  "  loge ".     "  Exp  x  "  is  sometimes  written  for 
"ex"  ;  "Exp(-  x)"  for  "e~x". 

Examples. — (1)  If  y  =  e  l0«  *  ;  show  y  =  x. 

(2)  If  log  I  =  -  an  ;  1=  e~an. 

(3)  If  6  =  be  -  at ;  log  b  -  log  6  =  at. 

(4)  If  loges  =  ad  ;  log10s  =  O'4343a0. 

(5)  Show  that  if  log  y0  -  log  y  =  kct ;  y  =  y0e  -  *<*. 

The  differentiation  of  exponential  functions  may  be  conveniently 

studied  in  three  sections  : 

(i)  Let 

y  =  ex. 

Take  logarithms,  andafchen,  differentiating,  we  get 
log  y  =  x  log  e ;  -^  =  dx,  or  -,-  =  ex ; 

° a  y  dx 

in  other  words,  the  differential  coefficient  of  ex  is  ex  itself,  or, 

«=-.  .  .  .  .  ,„ 

The  simplicity  of  this  equation,  and  of  (6)  in  the  preceding 
section,  explains  the  reason  for  the  almost  exclusive  use  of  natural 
logarithms  in  higher  mathematics. 


§  19.  THE  DIFFERENTIAL  CALCULUS.  55 

(ii)  Let 

y  =  a*. 
As  before,  taking  logarithms,  and  differentiating,  we  get 

log  y  =  x  log  a  ;  -£  =  y  log  a;  .-.  -^  =  a*  logea  .         (2) 

In  words,  the  differential  coefficient  of  a  constant  affected  with 

a  variable  exponent  is  equal  to  the  product  of  the  constant  affected 

with  the  same  exponent  into  the  logarithm  of  the  constant. 

(iii)  Let 

y  =  x% 

where  x  and  z  are  both  variable.  Taking  logarithms,  and  differ- 
entiating 

dy       _         _        zdx 
logy  =  zlogx  ;  —  =  log  xdz  +  — ; 

.-.  dy  =  xz  log  xdz  +  zxz~1dx  .         .         (3) 

If  x  and  z  are  functions  of  t,  we  have 

d(xz)      dy  dz  dx  ... 

-dt=dt  =  xn°zxdt  +  zx'-,di    •      *      <*> 

Examples. — (1)  The  amount,  x,  of  substance  transformed  in  a  chemical 
reaction  at  the  time  t  is  given  by  the  expression  x  =  ae  -  kt,  where  a  denotes 
•the  amount  of  substance  present  at  the  beginning  of  the  reaction,  hence  show- 
that  the  velocity  of  the  chemical  reaction  is  proportional  to  the  amount  of 
substance  undergoing  transformation.  Hint.  Show  that  dxjdt  =  -  kx,  and 
interpret. 

(2)  If  y  =  (a'+x)*,  dy/dx  =  2(ax  +  x)  (ax  log  a  +  1). 

(3)  If  y  =  a**,  dy/dx  =  naM  log  a. 

(4)  From  Magnus'  empirical  formula  for  the  relation  between  the  pres- 
sure of  aqueous  vapour  and  temperature 

o  e 

v-aby  +  o.    ,  dp       aylogb        +9 

where  a,  b,  y  are  constants.  This  differential  coefficient  represents  the  in- 
crease of  pressure  corresponding  with  a  small  rise  of  temperature,  say, 
roughly  from  6°  to  (6  +  1)°. 

(5)  Biot's  empirical  formula  for  the  relation  between  the  pressure  of 
aqueous  vapour,  p,  and  the  temperature,  6,  is 

logp  =  a  +  ba.9  -  c&o  ;  show  tt  =  pbrf  log  a  -  pc^  log  j8. 

(6)  Required  the  velocity  of  a  point  which'  moves  according  to  the 
equation  y  =  ae  -  M  cos  2ir(qt  +  e).  .  Since  velocity  =  dy/dt,  the  answer  is 
-  ae  -  a«{\  cos  2ar{qt  4.  e)  +  2irq  sin  2ir(qt  +  e)f 

(7)  The  relation  between  the  amount,  x,  of  substance  formed  by  two  con- 
secutive unimolecular  reactions  and  the  time  t  or  the  intensity  of  the  "  excited  " 
radioactivity  of  thorium  or  radium  emanations  at  the  time  t,  is  given  by  the 
expression 


56  HIGHER  MATHEMATICS.  §  20. 

ho  k,  dx         k,k0    f  \ 

where  \  and  k2  are  constants.  Show  that  the  last  expression  represents  the 
velocity  of  the  change. 

(8)  The  viscosity,  77,  of  a  mixture  of  non-electrolytes  (when  the  concentra- 
tions of  the  substances  with  viscosity  coefficients  A,  B,  C,  ...  are  x,  y,  z,  .  . . 
respectively)  is  97  =  AxBvCz . . .  Show  that  for  a  small  change  in  x,  y,  z  . . . 
dy\  becomes  ri(adx  +  bdy  +  cdz),  where  log  A  =  a,  log  B  =  b,  log  C  =  c.  Hint. 
Take  logs  before  differentiation. 

§  20.  The  "  Compound  Interest  Law "  in  Nature. 

I  cannot  pass  by  the  function  ex  without  indicating  its  great 
significance  in  physical  processes.  From  the  above  equations  it 
follows  that  if 

y  =  CtT;  then-£=be«*    .         .  (1) 

where  a,  b  and  G  are  constants,  b,  by  the  way,  being  equal  to 
aC  logee.  G  is  the  value  of  y  when  x  =  0.  Why  ?  It  will  be 
proved  later  on  that  this  operation  may  be  reversed  under  certain 
conditions,  and  if 


||  =  be?*9  then  y  =  Cte«    .         .        .        (2) 

where  a,  b  and  C  are  again  constant.  All  these  results  indicate 
that  the  rate  of  increase  of  the  exponential  function  ex  is  ex  itself. 
If,  therefore,  in  any  physical  investigation  xoe  find  some  function, 
say  y,  varying  at  a  rate  proportional  to  itself  (with  or  without 
some  constant  term)  we  guess  at  once  that  we  are  dealing  ivith  an 
exponential  function.     Thus  if 

~T~  =  ±  ay  J   we  may  write   y  =  Ceax,  or  Ce  ~  ax,  (2a) 

according  as  the  function  is  increasing  or  decreasing  in  magnitude. 

Money  lent  at  compound  interest  increases  in  this  way,  and 
hence  the  above  property  has  been  happily  styled  by  Lord  Kelvin 
"the  compound  interest  law"  (Encyc.  Brit.,  art.  "Elasticity," 
1877).  A  great  many  natural  phenomena  possess  this  property. 
The  following  will  repay  study : — 

Illustration  1. — Compound  interest.  If  £100  is  lent  out  at 
5  °/0  per  annum,  at  the  end  of  the  first  year  £105  remains.  If 
this  be  the  principal  for  a  second  year,  the  interest  during  that 
time  will  be  charged  not  only  on  the  original  £100,  but  also  on  the 


fc  20.  THE  DIFFERENTIAL  CALCULUS.  57 

additional  £5.  To  put  this  in  more  general  terms,  let  £p0  be  lent 
at  r  °/0  per  annum,  at  the  end  of  the  first  year  the  interest  amounts 

to  iTu^o>  an^  if  ^i  De  *ne  principal  for  the  second  year,  we  have  at 
the  end  of  the  first  year 

Pi  =  Poi1  +  iro) ; 

and  at  the  end  of  the  second  year, 

V2  =  Pii1  +  m)  =  p0(l  +  iTo)2. 
If  this  be  continued  year  after  year,  the  interest  charged  on  the 
increasing  capital  becomes  greater  and  greater  until  at  the  end  of 
t  years,  assuming  that  the  interest  is  added  to  the  capital  every 
year, 

P=Po(l  +  {J   -...         (3) 
Example. — Find  the  amount  (interest  +  principal)  of  £500  for  10  years 
at  5  °/o  compound  interest.     The  interest  is  added  to  the  principal  annually. 
From  (3),  log^  -  log  500  +  10  log  1-05  ;  .'.  p  =  £814  8s.  (nearly). 

Instead  of  adding  the  interest  to  the  capital  every  twelve 
months,  we  could  do  this  monthly,  weekly,  daily,  hourly,  and  so 
on.  If  Nature  were  our  banker  she  would  not  add  the  interest 
to  the  principal  every  year,  rather  would  the  interest  be  added  to 
the  capital  continuously  from  moment  to  moment.  Natura  non 
facit  saltus.  Let  us  imagine  that  this  has  been  done  in  order  that 
we  may  compare  this  process  with  natural  phenomena,  and  approxi- 
mate as  closely  as  we  can  to  what  actually  occurs  in  Nature.  As 
a  first  approximation,  suppose  the  interest  to  be  added  to  the 
principal  every  month.  It  can  be  shown  in  the  same  way  that  the 
principal  at  the  end  of  twelve  months,  is 

P  =  J>o(l  +  i^oo)12  ...         (4) 

If  we  next  assume  that  during  the  whole  year  the  interest  is  added 
to  the  principal  every  moment,  say  n  per  year,  we  may  replace  12 
by  n,  in  (4),  and 

*-*{v+mh)\     •     •     •     w 

For  convenience  in  subsequent  calculation,  let  us  put 
r         1  ur 

so  that  71   = 


100?i      u'  100' 

From  (5)  and  formula  (11),  page  28, 


P  =  PoU 


m 


58  HIGHER  MATHEMATICS.  §  20. 

But  (1  +  l/u)u  has  been  shown  in  (3),  page  52,  to  be  equivalent  to 
e  when  u  is  infinitely  great ;  hence,  writing  ^  =  a, 

v  =  Poea ; 

which  represents  the  amount  of  active  principal  bearing  interest  at 
the  end  of  one  year  on  the  assumption  that  the  interest  is  added  to 
the  principal  from  moment  to  moment.  At  the  end  of  t  years 
therefore,  from  (3), 

p  =  p0eat  ;  or,  p  =  _po0i°o<.    ...         (6) 

Example. — Compare  the  amount  of  £500  for  10  years  at  5°/0  compound 
interest  when  the'interest  is  added  annually  by  the  banker,  with  the  amount 
which  would  accrue  if  the  interest  were  added  each  instant  it  became  due. 
In  the  first  case,  use  (3),  and  in  the  latter  (6).  For  the  first  casej3  =  £814  8s.; 
for  the  second  p= £824  7s. 

Illustration  2. — Newton's  law  of  cooling.  Let  a  body  have  a 
uniform  temperature  0V  higher  than  its  surroundings,  it  is  required 
to  find  the  rate  at  which  the  body  cools.  Let  00  denote  the  tem- 
perature of  the  medium  surrounding  the  body.  In  consequence  of 
the  exchange  of  heat,  the  temperature  of  the  body  gradually  falls 
from  01  to  00.  Let  t  denote  the  time  required  by  the  body  to  fall 
from  #j  to  0.  The  temperature  of  the  body  is  then  0  -  00  above 
that  of  its  surroundings.  The  most  probable  supposition  that  we 
can  now  make  is  that  the  rate  at  which  the  body  loses  heat 
(-  dQ)  is  proportional  to  the  difference  between  its  temperature 
and  that  of  its  surroundings.     Hence 

where  k  is'  a  coefficient  depending  on  the  nature  of  the  substance. 
From  the  definition  of  specific  heat,  if  s  denotes  the  specific 
heat  of  unit  mass  of  substance. 

Q  =  s(0  -  0Q), ;  or  dQ  =  sdO. 
Substitute  this  in  the  former  expression.      Since  k/s  =  constant  = 
a  (say)  and  00  =  0°  C,  we  obtain 

ri-*  •  •  •  •  W 
or,  in  words,  the  velocity  of  cooling  of  a  body  is  proportional  to 
the  difference  between  its  temperature  and  that  of  its  surroundings. 
This  is  generally  styled  Newton's  law  of  cooling,  but  it  does  not 
quite  express  Newton's  idea  (Phil.  Trans.,  22,  827,  1701). 

Since  the  rate  of  diminution  of  0  is  proportional  to  6  itself,  we 


§  20. 


THE  DIFFEKENTIAL  CALCULUS. 


59 


guess  at  once  that  we  are  dealing  with  the  compound  interest  law, 
and  from  a  comparison  with  (1)  and  (2a)  above,  we  get 

0  =  be-at,        ....         (8) 
or  log  b  -  log  6  =  at.         .         .         .         (9) 

If  d1  represents  the  temperature  at  the  time  tv  and  62  the 
temperature  at  the  time  t2,  we  have 


log  b  -  log  01  =  atv  and  log  b  -  log  02 
By  subtraction,  since  a  is  constant,  we  get 

1 


at. 


a  = 


h  ~  h 


%? 


(10) 


The  validity  of  the  original  "  simplifying  assumption  "  as  to  the 
rate  at  which  heat  is  lost  by  the  body  must  be  tested  by  comparing 
the  result  expressed  in  equation  (10)  with  the  results  of  experiment. 
If  the  logical  consequence  of  the  assumption  agrees  with  facts, 
there  is  every  reason  to  suppose  that  the  working  hypothesis  is 
true.  For  the  purpose  of  comparison  we  may  use  A.  Winkelmann's 
data,  published  in  Wied.  Ann.,  44,  177,  429,  1891,  for  the  rate  of 
cooling  of  a  body  from  a  temperature  of  19"9°  C.  to  0°  C. 

If  0  denote  the  temperature  of  the  body  after  the  interval  of 
time  tl  -  t2  and  02  =  19'9,  $x  =  6,  remembering  that  in  practical 
work  Briggsian  logarithms  are  used,  we  obtain,  from  (10),  the 
expression 

1 


log 


#2 

10-0 


constant,  say  k. 


Winkelmann's  data  for  0  and  t1 
shown  in  the  following  table  : — 


t.2  are  to  be  arranged  as 


6. 

h  -  h. 

k  (calculated). 

18-9 

345 

0006490 

16-9 

10-85 

0-006540 

14-9 

19-30 

0-006511 

12-9 

28-80 

0-006537 

10-9 

40-10 

0-006519 

8-9 

53-75 

0  006502 

6-9 

70-95 

0-006483 

Hence,  h  is  constant  within  the  limits  of  certain  small  irregular 
variations  due  to  experimental  error.  Thus  the  truth  of  the  sup- 
position is  established  within  the  limits  of  the  errors  incidental  to 
Winkelmann's  method  of  measurement. 


60 


HIGHER  MATHEMATICS. 


§20. 


This  is  a  typical  example  of  the  way  in  which  the  logical  de- 
ductions of  an  hypothesis  are  tested.  There  are  other  methods. 
For  instance,  Dulong  and  Petit  (Ann.  Ghim.  Phys.,  [2],  7,  225,  337, 
1817)  have  made  the  series  of  exact  measurements  shown  in  the 
first  and  second  columns  of  the  following  table : — 


e,  excess  of 
temp,  of 

body  above 
that  of 
medium. 

V,  velocity  of  cooling  =  defdt. 

Observed. 

Calculated  by  the  formula  of  : 

Newton. 

Dulong  and 
Petit. 

Stefan. 

220° 
200° 
180° 
160° 
140° 
120° 
100° 

8-81 
7-40 
6-10 
4-89 
3-88 
3-02 
2-30 

682 
6-20 
5-58 
4-96 
4-34 
3-72 
3-10 

8-97 
7-41 
6-06 
4-91 
3-92 
3-08 
2-35 

8-95 
7-44 
6-11 
4-95 
3-94 
8-05 
2-30 

If  we  knew  the  numerical  value  of  the  constant  a  in  formula 
(7),  this  expression  could  be  employed  to  calculate  the  value  of  dO/dt 
for  any  given  value  of  6.  To  evaluate  a,  substitute  the  observed 
values  of  V  and  0  in  (7)  and  take  the  mean  of  the  different  results 
so  obtained.  Thus,  a  =  0*031.  The  third  column  shows  the 
velocities  of  cooling  calculated  on  the  assumption  that  Newton's 
law  is  true.  The  agreement  between  the  experimental  and  theo- 
retical results  is  very  poor.  Hence  it  is  necessary  to  seek  a  second 
approximation  to  the  true  law.  With  this  object,  Dulong  and  Petit 
have  proposed 

V=b(c°-l),  .         .         .         .         (11) 

as  a  second  approximation.  Here  b  =  2-037,  c  =  1*0077.  Column 
4  shows  the  velocity  of  cooling  calculated  from  Dulong  and  Petit'? 
law.  The  agreement  between  theory  and  fact  is  now  very  close. 
This  formula,  however,  has  no  theoretical  basis.  It  is  the  result 
of  a  guess.     Stefan's  guess  is  that 

V  =  a{(273  +  Of  -  (273)4},  .         .        (12) 

where  a  =  10  ~u  x  16- 72.  The  calculated  results  in  the  fifth  column 
are  quite  as  good  as  those  attending  the  use  of  Dulong  and  Petit 's 
formula.  Galitzine  has  pointed  out  that  Stefan's  formula  can  be 
established  on  theoretical  grounds. 

It  is  a  very  common  thing  to  find  different  formulae  agree,  so 


§  20.  THE  DIFFERENTIAL  CALCULUS.  Gl 

far  as  we  can  test  them,  equally  well  with  facts.  The  reader  must, 
therefore,  guard  against  implicit  faith  in  this  criterion — the  agree- 
ment between  observed  and  calculated  results — as  an  infallible 
experimentum  crucis. 

Lord  Kelvin  once  assumed  that  there  was  a  complete  transfor- 
mation of  thermal  into  electrical  energy  in  the  chemical  action  of  a 
galvanic  element.  Measurements  made  by  Joule  and  himself  with  a 
Daniell  element  gave  results  in  harmony  with  theory.  The  agree- 
ment was  afterwards  shown  to  be  illusory.  Success  in  explaining 
facts  is  not  necessarily  proof  of  the  validity  of  an  hypothesis,  for, 
as  Leibnitz  puts  it,  "  le  vrai  peut  etre  tire  du  faux,"  in  other  words, 
it  is  possible  to  infer  the  truth  from  false  premises. 

A  little  consideration  will  show  that  it  is  quite  legitimate  to 
deduce  the  numerical  values  of  the  above  constants  from  the 
experiments  themselves.  For  example,  we  might  have  taken  the 
mean  of  the  values  of  k  in  Winkelmann's  table  above,  and  applied 
the  test  by  comparing  the  calculated  with  the  observed  values  of 
either  t2  -  tv  or  of  0. 

Examples. — (1)  To  again  quote  from  Winkelmann's  paper,  if,  when  the 
temperature  of  the  surrounding  medium  is  99*74°,  the  body  cools  so  that  when 

0=119-97°,     117-97°,     115-97°,     113-97°,     111-97°,     109-97°; 

t  =       0,  12-6  26-7  42-9  61-2  83-1. 

Do  you  think  that  Newton's  law  is  confirmed  by  these  measurements  ? 
Hint.  Instead  of  assuming  that  0O  =  0,  it  will  be  found  necessary  to  retain 
0O  in  the  above  discussion.  Do  this  and  show  that  the  above  results  must  be 
tested  by  means  of  the  formula 

1  a     a 

i T  '  loSio^ a    =  constant. 

t2    —    &l  V\    —    Uq 

(2)  What  will  be  the  temperature  of  a  bowl  of  coffee  in  an  hour's  time  if 
the  temperature  ten  minutes  ago  was  80°,  and  is  now  70°  above  the  tempera- 
ture of  the  room  ?  Assume  Newton's  law  of  cooling.  Ansr.  31-2°  above  the 
surrounding  temperature.  Hint.  From  (8),  70  =  80.<?_10a  ;  .•.  a  =  0-0134; 
and  again,  x  =  80 .  e  -  °'0134  x  70.  We  cannot  apply  the  amended  laws— Dulong 
and  Petit's,  and  Stefan's — until  we  have  taken  up  more  advanced  work.  See 
(14)  and  (15),  page  372. 

Illustkation  3. — The  variation  of  atmospheric  pressure  with 
altitude  above  sea-level  can  be  shown  to  follow  the  compound 
interest  law.  Let  p0  be  the  pressure  in  centimetres  of  mercury  at 
the  so-called  datum  line,  or  sea-level,  p  the  pressure  at  a  height  h 
above  this  level.  Let  p0  be  the  density  of  air  at  sea-level  (Hg  =  1). 
Now  the  pressure  at  the  sea-level  is  produced  by  the  weight  of 


as-™  •     •     •     •     (13) 


62  HIGHER  MATHEMATICS.  §  20. 

the  superincumbent  air,  that  is,  by  the  weight  of  a  column  of  air 
of  a  height  h  and  constant  density  p0.  This  weight  is  equal  to  hpo. 
If  the  downward  pressure  of  the  air  were  constant,  the  barometric 
pressure  would  be  lowered  p0  centimetres  for  every  centimetre  rise 
above  sea-level.  But  by  Boyle's  law  the  decrease  in  the  density 
of  air  is  proportional  to  the  pressure,  and  if  p  denote  the  density 
of  air  at  a  height  dh  above  sea-level,  the  pressure  dp  is  given  by 
the  expression 

dp  —  -  pdh. 
If  we  consider  the  air  arranged  in  very  thin  strata,  we  may  regard 
the  density  of  the  air  in  each  stratum  as  constant.     By  Boyle's  law 

pp0  =  p0p ;  or,  p  =  p0plpQ. 
Substituting  this  value  of  p  in  the  above  formula,  we  get 

dp pop 

'  Po 

The  negative  sign  indicates  that  the  pressure  decreases  vertically 
upwards.  This  equation  is  the  compound  interest  law  in  another 
guise.  The  variation  in  the  pressure,  as  we  ascend  or  descend,  is 
proportional  to  the  pressure  itself.  Since  p0/po  is  constant,  we 
have  on  applying  the  compound  interest  law  to  (13), 

-£*• 

p  =  constant  X  6      p<i 
We  can  readily  find  the  value  of  the  constant  by  noting  that  at 
sea-level  h  =  0 ;  e°  =  1;  p  =  constant  x  e°  =  p0.     Substituting 
these  values  in  the  last  equation,  we  obtain 

-  -*  .         (14) 

p  =  pQe     po  ^  v     t 

a  relation  known  as  Halley's  law.     Continued  p.  260,  Ex.  (2). 

Illustration  4t.-^-The  absorption  of  actinic  energy  from  light 
passing  through  an  absorbing  medium.  The  intensity,  I,  of  a  beam 
of  light  is  changed  by  an  amount  dl  after  it  has  passed  through 
a  layer  of  absorbing  medium  dl  thick  in  such  a  way  that 

dl  =  -  aldl, 
where  a  is  a  constant  depending  on  the  nature  of  the  absorbing 
medium  and  on  the  wave  length  of  light.  The  rate  of  variation 
in  the  intensity  of  the  light  is  therefore  proportional  to  the  in- 
tensity of  the  light  itself,  in  other  words,  the  compound  interest 
law  again  appears.  Hence 
dl 


dl 


I ;    or  I  =c  constant  X  e 


§  20.  THE  DIFFERENTIAL  CALCULUS.  63 

If  I0  denote  the  intensity  of  the  incident  light,  then  when 

,  I  =  0,  I  =  I0  =  constant. 

Hence  the  intensity  of  the  light  after  it  has  passed  through  a 
medium  of  thickness  Z,  is 

/-!,»—      ....        (15) 

Examples. — (1)  A  1*006  cm.  layer  of  an  aqueous  solution  of  copper 
chloride  (2*113  gram  molecules  per  litre)  absorbed  18*13  °/0  of  light  in  the 
region  A  =  551  to  554  of  the  spectrum.  What  °/0  would  be  absorbed  by  a 
layer  of  the  same  solution  7*64  cm.  thick?  Ansr.  78*13  °/0.  Hint.  Find  a  in 
(15)  from  the  first  set  of  observations ;  I0  =  100,  I  =  81*87  ;  .*.  o  =  0*1989.  See 
T.  Ewan's  paper  ••  On  the  Absorption  Spectra  of  some  Copper  Salts  in  Aqueous 
Solution"  (Phil.  Mag.,  [5],  33,  317,  1892).     Use  Table  IV.,  page  616. 

(2)  A  pane  of  glass  absorbs  2  °/0  of  the  light  incident  upon  it.  How  much 
light  will  get  through  a  dozen  panes  of  the  same  glass  ?  Ansr.  78*66  °/0. 
Hint.  I0  =  100  ;  I  =  98 ;  a  =  0*02.     Use  Table  IV.,  page  616. 

Illustration  5. — Wilhelmy's  law  for  the  velocity  of  chemical 
reactions.  Wilhelmy  as  early  as  1850  published  the  law  of  mass 
action  in  a  form  which  will  be  recognised  as  still  another  example 
of  the  ubiquitous  law  of  compound  interest.  "  The  amount  of 
chemical  change  in  a  given  time  is  dir%jtly  proportional  to  the 
quantity  of  reacting  substance  present  in  "ftae  system." 

If  x  denote  the  quantity  of  changing  substance,  and  dx  the 
amount  of  substance  which  disappears  in  the  time  dt,  the  law  of 
mass  action  assumes  the  dress 

dx 

~dt  =  "  ' 
where  h  is  a  constant  depending  on  the  nature  of  the  reacting 
substance.  It  has  been  called  the  coefficient  of  the  velocity  of  the 
reaction,  its  meaning  can  be  easily  obtained  by  applying  the 
methods  of  §  10.  This  equation  is  probably  the  simplest  we  have 
yet  studied.  It  follows  directly,  since  the  rate  of  increase  of  x  is 
proportional  to  x,  that 

x  =  be  "  *', 
where  b  is  a  constant  whose  numerical  value  can  be  determined  if 
we  know  the  value  of  x  when  t  =  0.     The  negative  sign  indicates 
that  the  velocity  of  the  action  diminishes  as  time  goes  on. 

Examples. — (1)  If  a  volume  v  of  mercury  be  heated  to  any  temperature  0, 
the  change  of  volume  dv  corresponding  to  a  small  increment  of  temperature 
d8,  is  found  to  be  proportional  to  v,  hence  dv  =  avde.  Prove  Bosscha's  for- 
mula, v  =  ea&}  for  the  volume  of  mercury  at  any  temperature  6.     Ansr.  v  —  be^% 


64  HIGHER  MATHEMATICS.  §  21. 

where  a,  b  are  constants.     If  we  start  with  unit  volume  of  mercury  at  0°,  6  =  1 
and  we  have  the  required  result. 

(2)  According  to  Nordenskjold's  soluhility  law,  in  the  absence  of  super- 
saturation,  for  a  small  change  in  the  temperature,  dd,  there  is  a  change  in  the 
solubility  of  a  salt,  ds,  proportional  to  the  amount  of  salt  s  contained  in 
the  solution  at  the  temperature  0,  or  ds  =  asdO  where  a  is  a  constant.  Show 
that  the  equation  connecting  the  amount  of  salt  dissolved  by  the  solution 
with  the  temperature  is  s  =  s0ea8,  where  s0  is  the  solubility  of  the  salt  at  0°. 

(3)  If  any  dielectric  (condenser)  be  subject  to  a  difference  of  potential,  the 
density  p  of  the  charge  constantly  diminishes  according  to  the  relation  p  =  be~  at, 
where  b  is  an  empirical  constant ;  and  a  is  a  constant  equal  to  the  product  Air 
into  the  coefficient  of  conductivity,  c,  of  the  dielectric,  and  the  time,  t,  divided 
by  the  specific  inductive  capacity,  p,  i.e.,  a  =  ^irctjfx..  Hence  show  that  the 
gradual  discharge  of  a  condenser  follows  the  compound  interest  law.  Ansr. 
Show  dp/dt  =  -  ap. 

(4)  One  form  of  Dalton's  empirical  law  for  the  pressure  of  saturated 
vapour,  p,  between  certain  limits  of  temperature,  0,  is,  p  =  ae&.  Show  that 
this  is  an  example  of  the  compound  interest  law. 

(5)  The  relation  between  the  velocity  7  of  a  certain  chemical  reaction 
and  temperature,  0°,  is  log  V  =  a  +  bd,  where  a  and  b  are  constants.  Show 
that  we  are  dealing  with  the  compound  interest  law.  What  is  the  logical 
consequence  of  this  law  with  reference  to  reactions  which  (like  hydrogen  and 
oxygen)  take  place  at  high  temperatures  (say  500°),  but,  so  far  as  we  can 
tell,  not  at  ordinary  temperatures  ? 

(6)  The  rate  of  change  of  a  radioactive  element  is  represented  by 
dNjdt  =  -  rN  where  N  denotes  the  number  of  atoms  present  at  the  time  t, 
and  r  is  a  constant.  Show  that  the  law  of  radioactive  change  follows  the 
"  compound  interest  law  ". 

§  21.  Successive  Differentiation. 

The  differential  coefficient  derived  from  any  function  of  a 
variable  may  be  either  another  function  of  the  variable,  or  a  con- 
stant. The  new  function  may  be  differentiated  again  in  order  to 
obtain  the  second  differential  coeScient.  We  can  obtain  the  third 
and  higher  derivatives  in  the  same  way.     Thus,  if  y  =  x3, 

The  first  derivative  is,       -f-  =  Sx2 : 
ax 

The  second  derivative  is,  -^  =  &x '» 

The  third  derivative  is,     ^-4  =  6; 
dxz 


dx± 
It  will  be  observed  that  each  differentiation  reduces  the  index 


The  fourth  derivative  is,  -=-^  =  0. 
dx* 


§  21.  THE  DIFFERENTIAL  CALCULUS.  65 

of  the  power  by  unity.  If  the  index  n  is  a  positive  integer  the 
number  of  derivatives  is  finite. 

d2  d* 

In  the  symbols  ^(y),  T-g(y) ...  ,  the  superscripts  simply  de- 
note that  the  differentiation  has  been  repeated  2,  3 . .  .  times.     In 
differential  notation  we  may  write  these  results 
d2y  =  6x .  dx2 ;  d3y  =  6dx3 \  . .  . 
The  symbol  dx2,  dx* . . .,  meaning  dx  , ,dx,  dx  .dx  .dx...,  must 
not  be  confused  with  dx2  =  d(x2)  =  2x  .dx;  dx2  =  d(x)2  =  Sx2 .  dx  . . . 
The  successive  differential  coefficients  sometimes  repeat  them- 
selves ;  for  instance,  on  differentiating 

y  =  sin  x 
we  obtain  successively 

dy  d2y  .  dzy  dAy 

■3?  =  cos  x  :  -y4  =  -  sin  a; ;  -^  =  -  cos  x  ;  -^  =  sin  x  : . . . 

a#  da;2  da;3  da;4 

The  fourth  derivative  is  thus  a  repetition  of  the  original  function, 
the  process  of  differentiation  may  thus  be  continued  without  end, 
every  fourth  derivative  resembling  the  original  function.  The 
simplest  case  of  such  a  repetition  is 

y  =  &*> 
which  furnishes 

a    ^  *■  «w»  -.    r  d^      ' '  •■• 

The  differential  coefficients  are  all  equal  to  the  original  function 
and  to  each  other. 

Examples. — (1)  If  y  —  log  x ;  show  that  d4yjdx*  =  -  6/x4. 

(2)  If  y  =  x» ;  show  that  dly\dx*  =  n(n  -  l){n  -  2)(w  -  3)a«-4. 

(3)  If  y  =  x  -  2  ;  show  that  d^/dz3  =  -  24a?  -  5. 

(4)  If  2/  =  log  (a:  +  1) ;  show  that  d2y/dx2  =  -  (x  +  1)  -  2. 

(5)  Show  that  every  fourth  derivative  in  the  successive  differentiation  of 
y  =  cos  x  repeats  itself. 

Just  as  the  first  derivative  of  x  with  respect  to  t  measures  a 
velocity,  the  second  differential  coefficient  of  x  with  respect  to  t 
measures  an  acceleration  (page  17).     For  instance,  if  a  material * 


1 A  material  point  is  a  fiction  much  used  in  applied  mathematics  for  purposes 
of  calculation,  just  as  the  atom  is  in  chemistry.  An  atom  may  contain  an  infinite 
number  of  "  material  points  "  or  particles. 

E 


66 


HIGHER  MATHEMATICS. 


21. 


point,  P,  move  in  a  straight  line  AB  (Pig.  8)  so  that  its  distance, 
s,  from  a  fixed  point  0  is  given  by  the  equation  s  =  a  sin  t,  where 
a  represents  the  distance  OA  or  OB,  show  that  the  acceleration 


o 
Fig.  8. 

due  to  the  force  acting  on  the  particle  is  proportional  to  its  distance 
from  the  fixed  point.     The  velocity,  V,  is  evidently 

lit  =  a  cos  t '        •       •       •       w 

and  the  acceleration,  F,  is 


i^  = 


dF      d2s 


=  -  a  sin  t  =  -  s, 


dt~dt2~  "  °'      '         *         (2) 

the  negative  sign  showing  that  the  force  is  attractive,  tending  to 
lessen  the  distance  of  the  moving  point  from  0.  To  obtain  som6 
idea  of  this  motion  find  a  set  of  corresponding  values  of  F,  s  and  V 
from  Table  XIV.,  page  609,  and  (1)  and  (2)  above.     The  result  is 


If  t    = 

0 

^r 

IT 

|7T 

2tt... 

V   = 

a 

0 

-a 

0 

a. . . 

s     = 

0 

a 

0 

a 

0... 

F    = 

0 

-a 

0 

-a 

0... 

Pis  at 

0 

B 

0 

A 

0  ... 

A  careful  study  of  these  facts  will  convince  the  reader  that  the 
point  is  oscillating  regularly  in  a  straight  line,  alternately  right  and 
left  of  the  point  0.  In  this  sense,  the  equation  d2s/dt2  =  -  s 
describes  the  motion  of  the  particle.  It  is  called  an  equation  of 
motion.  An  equation  like  ds/dt  =  a  cos  t,  or  d2s/dt2  =  -  s,  con- 
taining differentials  or  differential  coefficients  is  called  a  differ- 
ential equation. 

Examples. — (1)  If  a  body  falls  from  a  vertical  height  according  to  the 
law  s  =  %gt2,  where  g  represents  the  acceleration  due  to  the  earth's  gravity, 
show  that  g  is  equal  to  the  second  differential  coefficient  of  s  with  respect  to  t. 

(2)  If  the  distance  traversed  by  a  moving  point  in  the  time  t  be  denoted 
by  the  equation  s  =  at2  +  bt  +  c  (where  a,  b  and  c  are  arbitrary  constants), 
show  that  the  acceleration  is  constant. 

(3)  Experiments  show  that  the  velocity  acquired  by  a  body  in  falling  from 
a  height  s  is  given  by  the  expression  V2  =  2g(s  ' x  -  s0  ~  l)r2,  where  g  denotes 
the  acceleration  of  gravitation  at  the  earth's  surface,  and  r  the  radius  of  the 


§  21.  THE  DIFFERENTIAL  CALCULUS.  67 

earth.  Show  that  the  acceleration  of  a  body  at  different  distances  from  the 
earth's  centre  is  inversely  as  the  square  of  its  distance  (Newton's  law).  Hint. 
Differentiate  the  equation  as  it  stands  ;  divide  by  dt  and  cancel  out  the  v  on 
one  side  of  the  equation  with  ds/dt  on  the  other.  Hence,  d^s/df*  =  -  gr2/s2 
remains.  Now  show  that  if  a  body  falls  freely  from  an  infinite  distance  the 
maximum  velocity  with  which  it  can  reach  the  earth  is  less  than  seven  miles 
per  second,  neglecting  the  resistance  of  the  air.  In  the  original  equation,  s0 
is  cd,  and  s  =  r  =  3,962  miles ;  g  =  32£  feet  =  0*00609  miles.  .-.  Ansr/=  6-95 
miles. 

(4)  Show  that  the  motion  of  a  point  at  a  distance  s  =  a  cos  qt  from  a 
certain  fixed  point  is  given  by  the  equation  d2s/dt2  =  -  q2s. 

(5)  Show  that  the  first  and  second  derivatives  of  De  la  Roche's  vapour 
pressure  formula,  p  =  ab9l{m  +  n6\  where  a,  b,  m,  and  n  are  constants,  are 

dp  _    mlogb  nh-^k .  &p  _  mlog  b{m  log  b  -  2n{m  +  nd)}  ,^j^. 
de  ~  (m  +  nef  »  de*  ~  (m  +  ney 

Fortunately,  in  applying  the  calculus  to  practical  work,  only  the 
first  and  second  derivatives  are  often  wanted,  the  third  and  fourth 
but  seldom.  The  calculation  of  the  higher  differential  coefficients 
may  be  a  laborious  process.  Leibnitz's  theorem,  named  after 
its  discoverer,  helps  to  shorten  the  operation.  It  also  furnishes 
us  with  the  general  or  nih.  derivative  of  the  function  which  is  useful 
in  discussions  upon  the  theory  of  the  subject.  We  shall  here 
regard  it  as  an  exercise  upon  successive  differentiation.  The  direct 
object  of  Leibnitz's  theorem  is  to  find  the  nth  differential  coefficient 
of  the  product  of  two  functions  of  x  in  terms  of  the  differential  co- 
efficients of  each  function. 

On  page  40,  the  differential  coefficient  of  the  product  of  two 
variables  was  shown  to  be 

dy  _  d{uv)  _     du        dv 
dx         dx  dx  "      dx* 

where  u  and  v  are  functions  of  x.  By  successive  differentiation 
and  analogy  with  the  binomial  theorem  (1),  page  36,  it  may  be 
shown  that 

dn(uv)        dnu        dv    dn~H  dnv 

The  reader  must  himself  prove  the  formula,  as  an  exercise,  by 
comparing  the  values  of  d2(uv)/dx2 ;  d3(uv)/dx3 ;  . . .,  with  the  de- 
velopments of  (x  +  h)2 ;  (x  +  h)z ;  . . .,  of  page  36. 

Examples.— (1)  If  y  =  x4 .  eax,  find  the  value  of  dzyjdxz.  Substitute  x4 
and  eax  respectively  for  v  and  u  in  (1).    Thus, 

v  =  x4 ;  .-.  dvfdx  =  4a;3 ;  d2v/dx2  =  12a;2 ;  d?v\dxz  =  24a; ; 
u  =  e«* ;  .-.  du/dx  =  ae** ;  dHt/dx2  =  a?eax ;  d5u/dx3  =  aPe**. 
E* 


68  HIGHER  MATHEMATICS.  §  22. 

From  (1) 

d?y  _    d?u        dv  d?u      n(n  -  1)    d?v   du        n{n  -  1)  (n  -  2)    d3v , 

dtf~vda?  +  ndx'dx,i+        2!       'dx*'dx  +  U~       .31  'dx3' 

/  „             dv          d?v      d3v\ 
=  e^v  +  Ba^  +  Sas-2+Wi); (2) 

=  e^ia^x4  +  lZatx3  +  d6ax2  +  24a;). 
(2)  If  y  =  log  x,  show  that  d6y(dx6  =  -  5!/oj8. 

If  we  pretend,  for  the  time  being,  that  the  symbols  of  operation 
■>-,  (  j~)  i  (^_)'m  (2)»  represent  the  magnitudes  of  an  operation, 
in  an  algebraic  sense,  we  can  write 

*SP  -  •-(• +  as)%  -  *"<• +  ^     '     (3) 

instead  of  (2),  and  substituting  D  for  -j-.     The  expression  (a  +  D)3 

is  supposed  to  be  developed  by  the  binomial  theorem,  page  36,  and 
dv/dx,  d2v/dx2, . . .,  substituted  in  place  of  Dv,  D2v,. . .,  in  the  re- 
sult. Equation  (3)  would  also  hold  good  if  the  index  3  were  re- 
placed by  any  integer,  say  n.  This  result  is  known  as  the  symbolic 
form  of  Leibnitz's  theorem. 

§  22.  Partial  Differentiation, 

Up  to  the  present  time  we  have  been  principally  occupied  with 
functions  of  one  independent  variable  x,  such  that 

u=f(x); 
but  functions  of  two,  three  or  more  variables  may  occur,  say 

u  =  f(x,  y,  z,.. .), 
where  the  variables  x,  y,  z, . . .  aie  independent  of  each  other.    Such 

functions  are  common.  As 
illustrations,  it  might  be  pointed 
out  that  the  area  of  a  triangle 
depends  on  its  base  and  altitude ; 
the  volume  of  a  rectangular  box 
depends  on  its  three  dimensions ; 
and  the  volume  of  a  gas  depends 
on  the  temperature  and  pressure. 


Fig.  9. 


I.  Differentials. 

To  find  the  differential  of  a  function  of  two  independent  vari- 
ables.    This  can  be   best  done  in  the  following  manner,  partly 


§  22.  THE  DIFFERENTIAL  CALCULUS.  69 

graphic  and  partly  analytical.  In  Fig.  9,  the  area  u  of  the  rect- 
angle ABGD,  with  the  sides  x,  y,  is  given  by  the  function 

u  =  xy. 
Since  x  and  y  are  independent  of  each  other,  the  one  may  be  sup- 
posed to  vary,  while  the  other  remains  unchanged.  The  function, 
therefore,  ought  to  furnish  two  differential  coefficients,  the  one  re- 
sulting from  a  variation  in  x,  and  the  other  from  a  variation  in  y. 
First,  let  the  side  x  vary  while  y  remains  unchanged.  The 
area  is  then  a  function  of  x  alone,     y  remains  constant. 

.-.  (du)y  =  ydx,  ....  (1) 
where  (du)y  represents  the  area  of  the  rectangle  BB'CG".  The 
subscript  denoting  that  y  is  constant. 

Second,  in  the  same  way,  suppose  the  length  of  the  side  y 
changes,  while  x  remains  constant,  then 

(du)x  =  xdy,  ....  (2) 
where  (du)x  represents  the  area  of  the  rectangle  DD'CC'.  Instead 
of  using  the  differential  form  of  these  variables,  we  may  write  the 
differential  coefficients 

/du\  /du\  7)u  7)u 

Kte)ry' and  w." X]  or  55 = y  • and  ^  -  *• 

in  C.  G.  J.  Jacobi's  notation,  where  ^—  is  the  symbol  of  differ- 
entiation when  all  the  variables,  other  than  x,  are  constant.  Sub- 
stituting these  values  of  x  and  y  in  (1)  and  (2),  we  obtain 

W*  =  ^cdx  ;  W*  =  ^dy' 
Lastly,  let  us  allow  x  and  y  to  vary  simultaneously,  the  total 
increment  in  the  area  of  the  rectangle  is  evidently  represented  by 
the  figure  D'EB'BGD. 

incr.  u  =  BB'CG"  +  DD'CC'  +  CC'C'E 
=  ydx  +  xdy  +  dx  .  dy. 
Neglecting  infinitely  small  magnitudes  of  the  second  order,  we  get 
du  =  ydx  +  xdy ;  .         .         .         .         (3) 

or  du  =  ^dx  +  ^dy,       ...         (4) 

which  is  also  written  in  the  form 

du = (M)?x + ©?y-  •   •   •   (a) 


70  HIGHER  MATHEMATICS.  §  22. 

In  equations  (3)  and  (4),  du  is  called  the  total  differential  of  the 
function  ;  ^-dx  the  partial  differential  of  u  with  respect  to  x  when 

y  is  constant ;  and  ^-dy  the  partial  differential  of  u  with  respect 

to  y  when  x  is  constant.  Hence  the  rule :  The  total  differential  of 
two  (or  more)  independent  variables  is  equal  to  the  sum  of  their 
partial  differentials. 

The  physical  meaning  of  this  rule  is  that  the  total  force  acting 
on  a  body  at  any  instant  is  the  sum  of  every  separate  action. 
When  several  forces  act  upon  a  material  particle,  each  force  pro- 
duces its  own  motion  independently  of  all  the  others.  The  actual 
velocity  of  the  particle  is  called  the  resultant  velocity,  and  the 
several  effects  produced  by  the  different  forces  are  called  the  com- 
ponent velocities.  There  is  here  involved  an  important  principle — 
the  principle  of  the  mutual  independence  of  different  reactions ; 
or  the  principle  of  the  coexistence  of  different  reactions — which  lies 
at  the  base  of  physical  and  chemical  dynamics.  The  principle 
might  be  enunciated  in  the  following  manner : — 

When  a  number  of  changes  are  simultaneously  taking  place  in 
any  system,  each  one  proceeds  as  if  it  were  independent  of  the  others ; 
the  total  change  is  the  sum  of  all  the  independent  changes.  Other- 
wise expressed,  the  total  differential  is  equal  to  the  sum  of  the 
partial  differentials  representing  each  change.  The  mathematical 
process  thus  corresponds  with  the  actual  physical  change. 

To  take  a  simple  illustration,  a  man  can  swim  at  the  rate  of 
two  miles  an  hour,  and  a  river  is  flowing  at  the  rate  of  one  mile  an 
hour.  If  the  man  swims  down-stream,  the  river  will  carry  him 
one  mile  in  one  hour,  and  his  swimming  will  carry  him  two  miles 
in  the  same  time.  Hence  the  man's  actual  rate  of  progress  down- 
stream will  be  three  miles  an  hour.  If  the  man  had  started  to 
swim  up-stream  against  the  current,  his  actual  rate  of  progress 
would  be  the  difference  between  the  velocity  of  the  stream  and  his 
rate  of  swimming.  In  short,  the  man  would  travel  at  the  rate  of 
one  mile  an  hour  against  the  current. 

This  means  that  the  total  change  in  u,  when  x  and  y  vary,  is 
made  up  of  two  parts :  (i)  the  change  which  would  occur  in  u  if 
x  alone  varied,  and  (ii)  the  change  which  would  occur  in  u  if  y 
alone  varied. 

Total  variation  =  variation  due  to  x  alone  +  variation  due  to  y  alone. 


§  22.  THE  DIFFERENTIAL  CALCULUS.  71 

If  the  meaning  of  the  different  terms  in 

,        ~du  ,        ~du  j 
du  =  Tr-dx  +  r-w 
7)x  ty 

is  carefully  noted,  it  will  be  found  that  the  equation  is  really  ex- 
pressed in  differential  notation,  not  differential  coefficients.  The 
partial  derivative  "bufdx  represents  the  rate  of  change  in  the  magni- 
tude of  u  when  x  is  increased  by  an  amount  ?>x,  y  being  constant ; 
similarly  'bufby  stands  for  the  rate  of  change  in  the  magnitude  of  u 
when  y  is  increased  by  an  amount  ~dy,  x  being  constant.  The  rate 
of  change  ~du/~dx  multiplied  by  dx,  furnishes  the  amount  of  change 
in  the  magnitude  of  u  when  x  increases  by  an  amount  dx,  y  being 
constant ;  and  similarly  (du/'dy)  dy  is  the  magnitude  of  the  change 
u  when  y  increases  an  amount  dy,  x  being  maintained  constant. 

Examples.— (1)  If  u  =  x*  +  xhj  +  if 

|^=  Sx2  +  2xy ;  |^  =  x2  +  Sy2  ;  .'.  du  =  (3a;2  +  2xy)dx  +  (x2  +  Sy2)dy. 

(2)  If  u  =  x  log  y  ;  du  =  logy.  dx+  x.dyjy. 

(3)  If  u  =  cos  x .  sin  y  +  sin  x .  cos  y  ; 

du  =  (dx  +  dy)(cos  x  cos  y  -  sin  x  sin  ?/)  =  (dx  +  d//){cos(aj  +  y)}. 

(4)  If  u  =  a? ;  du  =  i/a*  -  1d«  +  a*  log  ardi/. 

(5)  The  differentiation  of  a  function  of  three  independent  variables  may 
be  left  as  an  exercise  to  the  reader.  Neglecting  quantities  of  a  higher  order* 
if  u  be  the  volume  of  a  rectangular  parallelopiped l  having  the  three  dimen- 
sions x,  y,  z,  independently  variable,  then  u  =  xyz,  and 

^g^+l^l^  ....  (6) 
or  an  infinitely  small  increment  in  the  volume  of  the  solid  is  the  sum  of  the 
infinitely  small  increments  resulting  when  each  variable  changes  indepen- 
dently of  the  others.     Show  that 

du  =  yzdx  +  xzdy  +  xydz (7) 

(6)  If  the  relation  between  the  pressure  p,  and  volume  v,  and  tempera- 
ture d  of  a  gas  is  given  by  the  gas  law  pv  =  RT,  show  that  the  total  change 
in  pressure  for  a  simultaneous  change  of  volume  and  temperature  is 

(!).--"-  -*•  (#).=f =!■-  -  «r  >♦** 

This  expression  is  only  true  when  the  changes  dT  and  dv  are  made  in- 
finitesimal.    The  observed  and  calculated  values  of  dp,  arranged  side  by  side 

1  Mis-spelt  "  parallelopiped  "  by  false  analogy  with  "  parallelogram  ".  I  follow 
the  will  of  custom — quern  penes  arbitrium  est  etjus  et  norma  loquendi.  Etymologically 
the  word  should  be  spelt  "  parallelepiped  ".  It  only  adds  new  interest  to  learn  that 
the  word  is  derived  from  "  irapaWrjKeiriireSov  used  by  Plutarch  and  others  "  ;  and 
makes  one  lament  the  decline  of  classics. 


72 


HIGHEE  MATHEMATICS. 


§22. 


in  the  following  table  (from  J.  Perry's  The  Steam  Engine,  London,  564, 1904), 
show  that  even  when  dv  and  dT  are  relatively  large,  the  observed  values  agree 
pretty  well  with  the  calculated  results,  but  the  error  becomes  less  and  less  as 
dT  and  dv  are  made  smaller  and  smaller  : — 


T 

dT 
by  difference. 

V 

dv 
by  difference. 

V 

Obs. 

dp 

Calc. 

Obs. 

500 
501 

500-1 
500-01 

10 

o-i 
o-oi 

14-4 
14-5 
14-41 
14-40 

o-i 

o-oi 

o-ooi 

2000 
1990-2 
1999-2 
1999-9 

-9-8 
-1-0 

-o-i 

-9-9 
-0-90 

-o-io 

(7)  Clairaut's  formula  for  the  attraction  of  gravitation,  g,  at  different 
latitudes,  L,  on  the  earth's  surface,  and  at  different  altitudes,  H,  above  mean 
tide  level,  is 

g  =  980-6056  -  2-5028  cos  2Z,  -  0-000003H,  dynes. 
Discuss  the  changes  in  the  force  of  gravitation  and  in  the  weight  of  a  sub- 
stance with  change  of  locality.     Note,  "  weight "  is  nothing  more  than  a 
measure  of  the  force  of  gravitation. 


II.  Differential  Coefficients. 

To  find  the  differential  coefficient  of  a  function  of  two  indepen- 
dent variables.  If  the  variables  x  and  y  are  both  functions  of  t 
(say),  we  may  pass  directly  from  differentials  to  differential  co- 
efficients by  dividing  through  with  dt,  thus 

du  _  "du     dx      "du     dy 

dt  ~  ~dx     dt      ~dy  '  dt1 
which  may  also  be  written 

du  _  /du\  dx       /du\  dy  .ft. 

~di  =  \dx)yTt  +  \dy)x~dt'  •  •  •  W 
In  words,  the  total  variation  of  a  function  of  x  and  y  is  equal  to 
the  partial  derivative  of  the  function  when  y  is  constant  multiplied 
by  the  rate  of  variation  of  x,  added  to  the  partial  derivative  of  the 
function  when  x  is  constant  multiplied  by  the  rate  of  variation  of  y. 
If  the  function  remains  constant  while  its  variables  change,  the 
total  rate  of  change  of  the  function  is  zero, 

~du     dx      ~du     dy  _  _  ... 

^  •  St  +  ^  *  It  "  U      '       *       '       (yj 

Examples  of  this  will  be  given  very  shortly. 

When  there  is  likely  to  be  any  doubt  as  to  what  variables  have 
been  assumed  constant,  a  subscript  is  appended  to  the  lower  corner 


§  22.  THE  DIFFERENTIAL  CALCULUS.  73 

on  the  right  of  the  bracket.  The  subscripts  can  only  be  omitted 
when  there  is  no  possibility  of  confusing  the  variables  which  have 
been  assumed  constant.  For  example,  the  expression  ~dCvfiT  may 
have  one  of  three  meanings. 


(dC,\      (dC,\      (dC\ 
\dTjv'  \dTjp'  \dTJt 


Perry  suggests 1  the  use  of  the  alternative  symbols 

^±>     ^±>     15r 
l.T'  \T]  1>+T 

I  have  just  explained  the  meanings  of  the  partial  derivatives  of 
u  with  respect  to  x  and  y.  Let  me  again  emphasize  the  distinction 
between  the  partial  differential  coefficient  'du/'dx,  and  the  differential 
coefficient  du/dx.  In  'du/'dx,  y  is  treated  as  a  constant ;  in  du/dx, 
y  is  treated  as  a  function  of  x.  The  partial  derivative  denotes  the 
rate  of  change  of  u  per  unit  change  in  the  value  of  x  when  the 
other  variable  or  variables  remain  constant ;  du/dx  represents  the 
total  rate  of  change  of  u  when  all  the  variables  change  simultan- 
eously. 

Example. — If  y  and  u  are  functions  of  x  such  that 

y  =  sin  x ;  u  =  x  sin  x,  .  •  .  .  (10) 
we  can  write  the  last  expression  in  several  ways.  The  rate  of  change  of  u 
with  respect  to  x  (y  constant)  and  to  y  (x  constant)  will  depend  upon  the  way 
y  is  compounded  with  x.  The  total  rate  of  change  of  u  with  respect  to  x  will 
be  the  same  in  all  cases.     For  example,  we  get,  from  equations  (10), 

u  =  xy;  .-.  du  =  y .  dx  +  x  .  dy ;  u  =  x  sin  x ;  .-.  du  =  (x  cos  x  +  sin  x)dx ; 
u  =  sin  ~ x  y  .  sin  x ;  .*.  du  =  sin  ~xy  .  cos  x  .  dx  +  sin  x .  (1  -  y2)  ~  *  dy. 
The  partial  derivatives  are  all  different,  but  du/dx,  in  every  case,  reduces  to 
sin  x  +  x  cos  x. 

Many  illustrations  of  functions  with  properties  similar  to  those 
required  in  order  to  satisfy  the  conditions  of  equation  (8)  may 
occur  to  the  reader.  The  following  is  typical:' — When  rhombic 
crystals  are  heated  they  may  have  different  coefficients  of  ex- 
pansion in  different  directions.  A  cubical  portion  of  one  of  these 
crystals  at  one  temperature  is  not  necessarily  cubical  at  another. 
Suppose  a  rectangular  parallelopiped  is  cut  from  such  a  crystal, 
with  faces  parallel  to  the  three  axes  of  dilation.     The  volume  of 

the  crystal  is 

v  =  xyz, 


1  J.  Perry,  Nature,  66,  53,  271,  520,  1902  ;  T.  Muir,  same  references. 


74  HIGHER  MATHEMATICS.  §  22. 

where  x,  y,  z  are  the  lengths  of  the  different  sides.     Hence 

~dv  ~dv  ~dv 

—-  =  yz;  ^-   =  xz :  r—  =  xy. 

Substitute  in  (6)  and  divide  by  dO,  where  dO  represents  a  slight 
rise  of  temperature,  then 
dv  _     dx  dy  dz  1     dv  _  1    dx      1    dy       1    dz 

dO  ~  yZTt>  +  XZdO  +  xyd0 ;  or'  v  '  W  =  x  '  dO  +  y  '  TO  +  ~z  '  W 
where  the  three  terms  on  the  right  side  respectively  denote  the 
coefficients  of  linear  expansion,  A,  of  the  substance  along  the  three 
directions,  x,  y,  or  z.     The  term  on  the  left  is  the  coefficient  of 
cubical  expansion,  a.     For  isotropic  bodies,  a  =  3 A,  since 
1    dx  _  1    dy  __  1    dz 
x"dO~  y'dO~  z'dO' 
Examples. — (1)  Loschmidt  and  Obermeyer's  formula  for  the  coefficient 
of  diffusion  of  a  gas  at  T°  (absolute),  assuming  k0  and  TQ  are  constant,  is 

k  _  h  (l\nJL 
\TJ  760' 

where  k0  is  the  coefficient  of  diffusion  at  0°  C.  and  p  is  the  pressure  of  the  gas. 
Required  the  variation  in  the  coefficient  of  diffusion  of  the  gas  corresponding 
with  small  changes  of  temperature  and  pressure.     Put 

k  7)k  7)k 

a  =  7602^  ^j^=apnTn-^dT;  ^dp  =  aT-dp. 

.  .  dk-^1  +  ^dp.  ..dk-  ^^ 

(2)  Biot  and  Arago's  formula  for  the  index  of  refraction,  /*,  of  a  gas  or 

vapour  at  6°  and  pressure  p  is 

_  i       ^o  ~  1      P 
M  ~  1  +  a6  '  760' 

where  /t0  is  the  index  of  refraction  at  0°,  o  the  coefficient  of  expansion  of  the 
gas  with  temperature.  What  is  the  effect  of  small  variations  of  temperature 
and  pressure  on  the  index  of  refraction?     Ansr.  To  cause  it  to  vary  by  an 

,  j        fto~1f    dP  Pade    \ 

amount  dh  =  -^"Af^  "  (1  +  <*)*)' 

(3)  If  y  =  f(x  +  at),  show  that  dxjdt  =  a.     Hint.  Find  dy/dx,  and  dyldt; 
divide  the  one  by  the  other. 

(4)  If  u  =  xy,  where  x  and  y  are  functions  of  t,  show  that  (8)  reduces  to 

our  old  formula  (9),  page  41, 

du        dy        dx  /<<v 

(5)  If  x  is  a  function  of  t  such  that  x  =  t,  show  that  on  differentiation 
with  respect  to  t,  u  =  xy  becomes 

dJL  _  ^H  .   ^   fy  /i  o\ 

dt  ~  dx      dy'dt* \1Z> 


since  dtjdt  is  self-evidently  unity. 


§  23.  THE  DIFFERENTIAL  CALCULUS.  75 

(6)  If  x  and  y  are  functions  of  t,  show  that  on  differentiation  of  u  =  xy 
with  respect  to  tt 

du_ydudydu_'dudt 

dt~15y'  dt'  dy~  dt'dy *** 

A  result  obtained  in  a  different  way  on  page  44. 

§  23.  Euler's  Theorem  on  Homogeneous  Functions. 

One  object  of  Euler's  theorem  is  to  eliminate  certain  arbitrary- 
conditions  from  a  given  relation  between  the  variables  and  to  build 
up  a  new  relation  free  from  the  restrictions  due  to  the  presence  of 
arbitrary  functions.  I  shall  however  revert  to  this  subject  later 
on.  Euler's  theorem  also  helps  us  to  shorten  the  labour  involved 
in  making  certain  computations.  According  to  Euler's  theorem : 
In  any  homogeneous  function,  the  sum  of  the  products  of  each 
variable  with  the  partial  differential  coefficients  of  the  original 
function  with  respect  to  that  variable  is  equal  to  the  product  of 
the  original  function  with  its  degree.  In  other  words,  if  u  is  a 
homogeneous  function l  of  the  nth  degree,  Euler's  theorem  states 
that  if 

u  =  %aayyt      .        .        .     '    .        (1) 
when  a  +  /?  =  n,  then 2 

^u        ~bu 
Xte  +  yiy  =  nu-         •         •         •         (2) 

The  proof  is  instructive.     By  differentiation  of  the  homogeneous 
function, 

u  =  axay&  +  bx\y$\  +  . . .  =  %axa/y^% 

when  a  +  /?  =  aT  +  /?!  =  . . .  =  n,  we  obtain 

7)u  <)w 

5^  -  %aax-Y ;  and  ^  =  la/3x°yfi-\ 

Hence,  finally,  by  multiplying  the  first  with  x,  and  the  second 
with  y}  and  adding  the  two  results,  we  obtain 
7)U         ~du 

x^x  +  yty  =  ^a(a  +  ®xayfi  =  n^axayp  ■  nu- 

The  theorem  may  be  extended  to  include  any  number  of  variables 


i  An  homogeneous  function  is  one  in  which  all  the  terms  containing  the  variables 
have  the  same  degree.  Examples  :  x2  +  bxy  +  z2 ;  x*  +  xyz2  +  xsy  +  x2z2  are  homo- 
geneous functions  of  the  second  and  fourth  degrees  respectively. 

2  The  sign  "2"  is  to  be  read  "the  sum  of  all  terms  of  the  same  type  as  . . .," 
or  here  "  the  sum  of  all  terms  containing  x,  y  and  constants  ".  The  symbol  "  n  "  is 
sometimes  used  in  the  same  way  for  "the  product  of  all  terms  of  the  type  ". 


76  HIGHER  MATHEMATICS.  §  24. 

so  that  if 

*--*$$ -:v)     ...     (3) 

we  may  write  down  at  once, 

•&+*%*••'•?*?  ■    •    •    w 

and  we  have  got  rid  of  the  conditions  imposed  upon  u  in  virtue  of 
the  arbitrary  function  /(. .  .). 

7n?y  r*\n  *^t7V 

Examples. — (1)  If  u  =  x2y  +  xy2  +  Zxyz,  then  x^-  +  y^~  +  z-~-  =  3w. 

Prove  this  result  by  actual  differentiation.     It  of  course  follows  directly  from 

Euler's  theorem,  since  the  equation  is  homogeneous  and  of  the  third  degree. 

/ov  T*         xP  +  xhj  +  y3      'du       du  M 

(2)  If  %  =  a,a  +  gy  +  y8  ?  ^  +  2/^  =  ^»  smce  tne  equation  is  of  the  first 

degree  and  homogeneous. 

(x\  'du        *du 

-  ),  show  that  x-^r  +  y^~  =  0.     Here  n  in  (3)  is  zero.     Prove 

the  result  by  actual  differentiation. 

§  2$.    Successive  Partial  Differentiation. 

We  can  get  the  higher  partial  derivatives  by  combining  the 
operations  of  successive  and  partial  differentiation.     Thus  when 

u  =  x2  +  y2  +  x2y*, 
the  first  derivatives  of  u  with  respect  to  x,  when  y  is  constant,  and 
to  y,  when  x  is  constant  are  respectively 

g-ar  +  a^;  §-2y  +  3*Y;    .       .       (1) 

repeating  the  differentiation, 

S^  =  2(l  +  2/8);^=2(l  +  3A),    .        .        (2) 

If  we  had  differentiated  "bufbx  with  respect  to  y,  and  "hufby  with 
respect  to  x,  we  should  have  obtained  two  identical  results,  viz. : — 

Wx  =  6y2x'*nd^~y  =  6fx'      *      *      (3) 

The  higher  partial  derivatives  are  independent  of  the  order  of 
differentiation.      By  differentiation   of  ~du/~dx  with   respect  to  y, 

assuming    x    to    be    constant,    we    get  -r — ,   which   is   written 
;  on  the  other  hand,   by  the  differentiation  of  <—  with  re- 


7>yl>x'  u"  °"°  v""~  "-T-   ujr   ""  umciDUWawuu  ^  ty 


§  25.  THE  DIFFERENTIAL  CALCULUS.  77 

spect  to  x,  assuming  y  to  be  constant,  we  obtain  ^  ■<  .      That  is 

to  say 

Vu         Wu  (i) 

~by!)x  ~  ~dxby 
This  was  only  proved  in  (3)  for  a  special  case.  As  soon  as  the 
reader  has  got  familiar  with  the  idea  of  differentiation,  he  will  no 
doubt  be  able  to  deduce  the  general  proof  for  himself,  although  it 
is  given  in  the  regular  text- books.  The  result  stated  in  (4)  is  of 
great  importance. 

Example. — If  y  =  e**  +  &  + *  is  to  satisfy  the  equation 

6how  that  a2  =  A0*  +  Bfi,  where  a,  £,  y,  are  constants.     Hint.  First  find  the 
three  derivatives  and  substitute  in  the  second  equation  ;  reduce. 

§  25.    Complete  or  Exact  Differentials. 

To  find  the  condition  that  u  may  be  a  function  of  x  and  y  in 
the  equation 

du  =  Mdx  +  Ndy,  ...         (5) 

where  M  and  N  are  functions  of  x  and  y.     We  have  just  seen  that 
if  u  is  a  function  of  x  and  y 

(6) 


du    =      _^    +    _dyf 

•      .     •         .'  • 

that  is  to  say,  by  comparing  (5)  and  (6) 

TIT            ^U 

Differentiating  the 

first  with 

respect  to  y, 

and  the  secoi 

respect  to  x,  we  have, 

from  (4) 

?)M 

_7)N 

~dy  ' 

~dx' 

(7) 

In  text-books  on  differential  equations  this  condition  is  shown 
to  be  necessary  and  sufficient  in  order  that  certain  equations  may 
be  solved,  or  "  integrated  "  as  it  is  called.  Equation  (7)  is  called 
Euler's  criterion  of  integrability.  An  equation  that  satisfies 
this  condition  is  said  to  be  a  complete  or  an  exact  differential. 

Example. — Show  that  ydx  -  xdy  =  0,  is  not  exact,  and  that  ydx  +  xdy  =  0 
is  a  complete  differential.  Hint.  dM/dy  =  dy/dy  \  and  dN/dx  =  -  dx/dx ; 
hence,  in  the  first  case,  dM/dy  is  not  equal  to  dN/dx,  and  therefore  the  equa- 
tion is  not  exact ;  etc. 


78  HIGHER  MATHEMATICS.  §  26. 

§  26.   Integrating  Factors. 

The  equation 

Mdx  +  Ndy  =  0  .         .        .        .         (8) 

can  always  be  made  exact  by  multiplying  through  with  some  func- 
tion of  x  and  y,  called  an  integrating  factor.  (M  and  N  are  sup- 
posed to  be  functions  of  x  and  y.) 

Since  M  and  N  are  functions  of  x  and  y,  (8)  may  be  written 

dy  _     M 

dx f-T-  .TT     -       *       *       '       (9) 

or  the  variation  of  y  with  respect  to  x  is  as  -  M  is  to  N ;  that  is 
to  say,  x  is  some  function  of  y,  say 

f(x,  y)  =  a, 
then  from  (5),  page  69, 

~^^~^  +  _^_^  =  a  '  '  (10) 
By  a  transformation  of  (10),  and  a  comparison  of  the  result  with  (9), 
we  find  that 


(ii) 


dy  _  da;  If 

d#  . . .       of(x,  2/)  ~       27 

Hence 

where  p  is  either  a  function  of  #  and  y,  or  else  a  constant.  Multi- 
plying the  original  equation  by  the  integrating  factor  /x,  and 
substituting  the  values  of  fiM,  fxN  obtained  in  (12),  we  obtain 

WMldx  +  MM>4*  -  0, 
ox  oy 

which  fulfils  the  condition  of  exactness.  The  function  f(x,  y)  is  to 
be  derived  in  any  particular  case  from  the  given  relation  between 
x  and  y. 

Example. — Show  that  the  equation  ydx  -  xdy  =  0  becomes  exact  when 
multiplied  by  the  integrating  factor  1/y2. 

*dM=  _  1  \  3^=_  1 

dy  .       y2'  dx  ~     y* 
Hence  'dMf'dy  =  dN/dx,  the  condition  required  by  (7).     In  the  same  way  show 
that  \\xy  and  ljx2  are  also  integrating  factors. 

Integrating  factors  are  very  much  used  in  solving  certain  forms 
of  differential  equations  (q.v.),  and  in  certain  important  equations 
which  arise  in  thermodynamics. 


§  27.  THE  DIFFEKENTIAL  CALCULUS.  79 

§  27.    Illustrations  from  Thermodynamics. 

As  a  first  approximation  we  may  assume  that  the  change  of 
state  of  every  homogeneous  liquid,  or  gaseous  substance,  is  com- 
pletely defined  by  some  law  connecting  the  pressure,  p,  volume,  v, 
and  temperature,  T.  This  law,  called  the  characteristic  equation, 
or  the  equation  of  state  of  the  substance,  has  the  form 

f{p.v,T)-0 (1) 

Any  change,  therefore,  is  completely  determined  when  any  two  of 
these  three  variables  are  known.     Thus,  we  may  have 

p  =  Mv,  T);v-  f2(p,  T) ;  or,  T  =  f3(p,  v).  .  (2) 
Confining  our  attention  to  the  first,  we  obtain,  by  partial  differen- 
tiation, 

dp  -  8$  * + (&),dT>    •  •   « 

The  first  partial  derivative  on  the  right  represents  the  coefficient 
of  elasticity  of  the  gas,  the  second  is  nothing  but  the  so-called 
coefficient  of  increase  of  pressure  with  temperature  at  constant 
volume.  If  the  change  takes  place  at  constant  pressure,  dp  =  0, 
and  (3)  may  be  written  in  the  forms 

/dp\  L{^1\ 

(dv\  \dT)v .        fdp\  v  \dTJp 


ap\  \dTj9        i/av\ 

\dvJT  v \dp)T 


The  subscript  is  added  to  show  which  factor  has  been  supposed 
constant  during  the  differentiation.  Note  the  change  of  ~ov[oT  to 
dv/dT  at  constant  pressure.  The  first  of  equations  (4)  states  that 
the  change  in  the  volume  of  a  gas  when  heated  is  equal  to  the  ratio 
of  the  increase  of  pressure  with  temperature  at  constant  volume, 
and  the  change  in  the  elasticity  of  the  gas  ;  the  second  tells  us 
that  the  ratio  of  the  coefficients  of  thermal  expansion  and  of  com- 
pressibility is  equal  to  the  change  in  the  pressure  of  the  gas  per 
unit  rise  of  temperature  at  constant  volume. 

Examples. — (1)  Show  that  a  pressure  of  60  atmospheres  is  required  to 

keep  unit  volume  of  mercury  at  constant  volume  when  heated  1°  0.     Co- 

l/dv\ 
efficient    of    expansion    of    Hg   =  0*00018  =  Z\Tm)    '»   °*    compressibility 

1  /dv\ 


=  0  000003  =-— (-T-)  .       M.   Planck,   Vorlesungen  iiber  Thermodynamik. 

Leipzig,  8,  1897. 

(2)  J.  Thomsen's  formula  for  the  amount  of  heat  Q  disengaged  when  one 


80  HIGHER  MATHEMATICS.  §  27. 

molecule  of  sulphuric  acid,  H2S04,  is  mixed  with  n  molecules  of  water,  H20, 
is  g  =  17860  n/(l-798  +  n)  cals.  Put  a  =  17860  and  b  =  1-798,  for  the  sake  of 
brevity.  If  x  of  H2S04  be  mixed  with  y  of  H20,  the  quantity  of  heat  dis- 
engaged by  the  mixture  is  x  times  as  great  as  when  one  molecule  of  H2S04 
unites  with  yjx  molecules  of  water.  Since  y/x  =  n  in  Thomsen's  formula 
Q  =  x  x  ay/(bx  +  y)  cals.  If  dx  of  acid  is  now  mixed  with  x  of  H2S04  and  y 
of  H20,  show  that  the  amount  of  heat  liberated  is 

^dx = whyfx ;  or>  j^hzfx  cals- 

In  the  same  way  the  amount  of  heat  liberated  when  dy  of  water  is  added  to  a 
similar  mixture  is 

Let  Q,  T,  p,  v,  represent  any  four  variable  magnitudes  what- 
ever.    By  partial  differentiation 

Equate  together  the  second  and  last  members  of  (5),  and  substitute 
the  value  of  dp  from  (3),  in  the  result.     Thus, 

Put  dv  =  0,  and  divide  by  dT, 

(§Hi).(H),  .  .  •  .  <v» 

Again,  by  partial  differentiation 

dT- (§).*  +  ©* <8> 

Substitute  this  value  of  dT  in  the  last  two  members  of  (5), 
Put  dp  =  0,  and  write  the  result 

_  (iHIHm > 

By  proceeding  in  this  way,  the  reader  can  deduce  a  great 
number  of  relations  between  Q,  T,  p,  v,  quite  apart  from  any 
physical  meaning  the  letters  might  possess.  If  Q  denotes  the 
quantity  of  heat  added  to  a  substance  during  any  small  changes 
of  state,  and  p,  v,  T,  the  pressure,  volume  and  absolute  tempera- 
ture of  the  substance,  the  above  formulae  are  then  identical  with 
corresponding  formulae  in  thermodynamics.  Here,  however,  the 
relations  have  been  deduced  without  any  reference  to  the  theory 
of  heat.      Under  these  circumstances,  {dQftT)vdT  represents  the 


§  27.  THE  DIFFERENTIAL  CALCULUS.  81 

quantity  of  heat  required  for  a  small  rise  of  temperature  at  con- 
stant volume:  (bQfiT),  is  nothing  but  the  specific  heat  of  the 
substance  at  constant  volume,  usually  written  0,;  similarly, 
(dQfiT)p  is  the  specific  heat  of  constant  pressure,  written  Cp;  and 
(pQfiv)T  and  (dQ/~i)p)T  refer  to  the  two  latent  heats. 

These  results  may  be  applied  to  any  substance  for  which  the 
relation  pv  =  BT  holds  good.     In  this  case, 

Examples. — (1)   A  little  ingenuity,  and  the  reader  should  be  able  to 
deduce  the  so-called  Reech's  Theorem : 


m 

-c.-J7W?> (11) 

\dv)T 


employed  by  Clement  and  Desormes  for  evaluating  y.  See  any  text-book  on 
physics  for  experimental  details.  Hint.  Find  dp  for  v  and  Q  ;  and  for  v  and 
T  as  in  (3) ;  use  (7)  and  (10). 

(2)  By  the  definition  of  adiabatic  and  isothermal  elasticities  (page  113), 
E(j)  =  -  v(dp/dv)^  ;  and  ET  =  -  v(dp[dv)T,  respectively. 
The  subscripts  <p  and  Vindicating,  in  the  former  case,  that  there  has  been 
neither  gain  nor  loss  of  heat,  in  other  words  that  Q  has  remained  constant, 
and  in  the  latter  case,  that  the  temperature  remained  constant  during  the 
process  'dp/dv.  Hence  show  from  the  first  and  last  members  of  (5),  when  Q 
is  constant, 

fdQ\ 

V^A  fdQ\ ' 

\dpj* 

From  (7),  (10)  and  (4),  we  get  the  important  result 

E4>_  \-dv)„  \dT)p\^v)f\dTjv      \'dTjp      Cv 

VdphYdojT  {dTjXdvJr  \dTjv 

According  to  the  second  law  of  thermodynamics,  for  reversible 
changes  "the  expression  dQ/T  is  a  perfect  differential".  It  is 
usually  written  d<f>,  where  <f>  is  called  the  entropy  of  the  substance. 
From  the  first  two  members  of  (5),  therefore, 

§^*-MMldT+W^-  •  •   (18» 

is  a  perfect  differential.     From  (7),  page  77,  therefore, 

dfl   -dQ\        dfl   dQ\  fdCv\       fdL\       L  (u) 

where  C„  has  been  written  for  (dQfiT),,  L  for  (lQfdv)r 


82  HIGHER  MATHEMATICS.  §  27. 

According  to  the  first  law  of  thermodynamics,  when  a  quantity 
of  heat  dQ  is  added  to  a  substance,  part  of  the  heat  energy  dU  is 
spent  in  the  doing  of  internal  work  among  the  molecules  of  the 
substance  and  part  is  expended  in  the  mechanical  work  of  expansion, 
p .  dv  against  atmospheric  pressure.     To  put  this  symbolically, 

dQ  =  dU  +  pdv;  or  dU  =  dQ  -  pdv.  .  .  .  (15) 
Now  d  U  is  a  perfect  differential.  This  means  that  however  much 
energy  U,  the  substance  absorbs,  all  will  be  given  back  again  when 
the  substance  returns  to  its  original  state.  In  other  words,  U  is  a 
function  of  the  state  of  the  substance  (see  page  385).  This  state 
is  determined,  (2)  above,  when  any  two  of  the  three  variables 
p}  v,  T,  are  known. 

For  the  first  two  members  of  (5),  and  the  last  of  equations  (15), 
therefore, 

dU  =  Cv.dT  +  L.dv  -  pdv  =  Cv.dT  +  {L  -  p)dv,         .      (16) 
is  a  complete  differential.     In  consequence,  as  before, 

(m-mi-m  ■  ■  ■  ■  ™ 

From  (14)  and  (17), 

\w)v=T\Wv)t' {18) 

a  "  law  "  which  has  formed  the  starting  point  of  some  of  the  finest 
deductions  in  physical  chemistry. 

Examples. — (1)  Establish  Mayer's  formula,  for  a  perfect  gas. 

Cp-  CV  =  R, (19) 

Hints:  (i.)  Since pv  =  RT,  @p/3Z%  =  Rfv  ;  .'.  (dQ/dv)T  =  RT/v  =  p,  by  (18). 
(ii.)  Evaluate  dv  as  in  (3),  and  substitute  the  result  in  the  second  and  third 
members  of  (5).  (iii.)  Equate  dp  to  zero.  Find  2>v/dT  from  the  gas  equation, 
use  (18),  etc.     Thus, 

(2)  Establish  the  so-called  "Four  thermodynamic  relations"  between 
P,  v,  T,  (f>,  when  any  two  are  taken  as  independent  variables. 

(dr\     _fdp_\  .  fd$\  _/dp\  ,  fdr\  _fdv\  .  fd±\      ;/i\ 

Ydvjtt,-     \d<t>)v'  \dvJT~\dTjv'  \^pJ4>~\^ct>JP,  \dpjr~     \dT); 

It  is  possible  that  in  some  future  edition  of  this  work  a  great 
deal  of  the  matter  in  the  next  chapter  will  be  deleted,  since 
"graphs  and  their  properties  "  appears  in  the  curriculum  of  most 
schools.  However,  it  is  at  present  so  convenient  for  reference  that 
I  have  decided  to  let  it  remain. 


CHAPTER  II. 

COORDINATE  OR  ANALYTICAL  GEOMETRY. 

"  Order  and  regularity  are  more  readily  and  clearly  recognised  when 
exhibited  to  the  eye  in  a  picture  than  they  are  when  presented 
to  the  mind  in  any  other  manner." — Dr.  Whewell. 

§  28.    Cartesian  Coordinates. 

The  physical  properties  of  a  substance  may,  in  general,  be  con- 
cisely represented  by  a  geometrical  figure.  Such  a  figure  furnishes 
an  elegant  method  for  studying  certain  natural  changes,  because 
the  whole  history  of  the  process  is  thus  brought  vividly  before  the 
mind.  At  the  same  time  the  numerical  relations  between  a  series 
of  tabulated  numbers  can  be  exhibited  in  the  form  of  a  picture  and 
their  true  meaning  seen  at  a  glance. 

Let  xOx'  and  yOy'  (Fig.  10)  be  two  straight  lines  at  right  angles 
to  each  other,  and  intersecting  at  the  point  0,  so  as  to  divide  the 
plane  of  this  paper  into  four  quadrants  I,  II,  III  and  IV.  Let 
Pl  be  any  point  in  the  first  quadrant  yOx ;  draw  PiMY  parallel  to 
Oy  and  PYN  parallel  to  Ox.  Then,  if  the  lengths  0Ml  and  P^ 
are  known,  the  position  of  the  point  P  with  respect  to  these  lines 
follows  directly  from  the  properties  of  the  rectangle  NP-^Mft 
(Euclid,  i.,  34).  For  example,  if  OMY  denotes  three  units,  PlM1 
four  units,  the  position  of  the  point  P1  is  found  by  marking  off 
three  units  along  Ox  to  the  right  and  four  units  along  Oy  vertically 
upwards.  Then  by  drawing  NP1  parallel  to  Ox,  and  P^M\  parallel 
to  Oy,  the  position  of  the  given  point  is  at  Pv  since, 

P^  =  ON  =  4  units ;  NP1  =  0M1  =  3  units. 

x'Ox,  yOy'  are  called  coordinate  axes  or  "  frames  of  reference  " 
(Love).  If  the  angle  yOx  is  a  right  angle  the  axes  are  said  to  be 
rectangular.     Conditions  may  arise  when  it  is  more  convenient 

88  f* 


84 


HIGHER  MATHEMATICS. 


§28. 


to  make  yOx  an  oblique  angle,  the  axes  are  then  said  to  be  oblique. 
xOx'  is  called  the  abscissa  or  x-axis,  yOy'  the  ordinate  or  y-axis. 

The  point  0  is  called  the  origin ;  OM1  the  abscissa  of  the  point 
P,  and  P1M1  the  ordinate  of  the  same  point.  In  referring  the  posi- 
tion of  a  point  to  a  pair  of  coordinate  axes,  the  abscissa  is  always 
mentioned  first,  P{  is  spoken  of  as  the  point  whose  coordinates  are 
3  and  4 ;  it  is  written  "the  point  P:(3,  4)".  In  memory  of  its 
inventor,  Rene  Descartes,  this  system  of  notation  is  sometimes 
styled  the  system  of  Cartesian  coordinates. 

The  usual  conventions  of  trigonometry  are  made  with  respect 
to  the  algebraic  sign  of  a  point  in  any  of  the  four  quadrants.  Any 
abscissa  measured  from  the  origin  to  the  right  is  positive,  to  the 


1 

f 

n 

M 

P, 

I 

p* 

=x* 

0 

JC 

M2 

M3 

M* 

Mi 

P4 

Pa 

m 

&-, 

TV 

Fig.  10. — Cartesian  Coordinates — Two  Dimensions. 


left,  negative  ;  ordinates  measured  vertically  upward  are  positive, 
and  in  the  opposite  direction,  negative.  For  example,  if  a  and 
b  be  any  assigned  number  of  units  corresponding  respectively  to 
the  abscissa  and  ordinate  of  some  given  point,  then  the  Car- 
tesian coordinates  of  the  point  Px  are  represented  as  P1  (a,  b),  of 
P2  as  P2(  -  a,  b),  of  P3  as  P3(  -  a,  -  b)  and  of  P4  as  P4(a,  -  b). 
Points  falling  in  quadrants  other  than  the  first  are  not  often  met 
with  in  practical  work. 

Thus,  any  point  in  a  plane  represents  two  things,  (1)  its  hori- 


§29. 


COORDINATE  OR  ANALYTICAL  GEOMETRY. 


85 


zontal  distance  along  some  standard  line  of  reference — the  #-axis, 
and  (2)  its  vertical  distance  along  some  other  standard  line  of  refer- 
ence— the  2/-axis. 

When  the  position  of  a  point  is  determined  by  two  variable  mag- 
nitudes (the  coordinates),  the  point  is  said  to  be  two  dimensional. 

We  are  always  making  use  of  coordinate  geometry  in  a  rough 
way.  Thus,  a  book  in  a  library  is  located  by  its  shelf  and  number ; 
and  the  position  of  a  town  in  a  map  is  fixed  by  its  latitude  and 
longitude.  See  H.  S.  H.  Shaw's  "  Report  on  the  Development  of 
Graphic  Methods  in  Mechanical  Science,"  B.  A.  Beports,  373, 
1892,  for  a  large  number  of  examples. 

§  29.    Graphical  Representation. 

Consider  any  straight  or  curved  line  OP  situate,  with  refer- 
ence to  a  pair  of  rectangular  co-ordinate  axes,  as  shown  in  Fig.  11. 
Take  any  abscissae    OMv  OM^  OM3, .  . .  OM,   and   through   Mv 


Fig.  11. 

M2...M  draw  the  ordinates  MlPv  M2P2 . . .  MP  parallel  to  the 
2/-axis.  The  ordinates  all  have  a  definite  value  dependent  on  the 
slope  of  the  line 1  and  on  the  value  of  the  abscissas.  If  x  be  any 
abscissa  and  y  any  ordinate,  x  and  y  are  connected  by  some 
definite  law  called  the  equation  of  the  curve. 

It  is  required  to  find  the  equation  of  the  curve  OP.     In  the 
triangle  OPM 

MP  =  OM  tan  MOP, 


3  Any  straight  or  curved  line  when  referred  to  its  coordinate  axes,  is  called  a 
curve  ". 


80  HIGHER  MATHEMATICS.  §  30. 

or  y  =  x  tan  a,     .         .         .         .         (1) 

where  a  denotes  the  positive  angle  MOP.    But  if  OM  =  MP, 

MP 
tan  MOP  =  g±  =  1  =  tan  45°. 

The  equation  of  the  line  OP  is,  therefore, 

y  =  x;  .         .        .         .         (2) 

and  the  line  is  inclined  at  an  angle  of  45°  to  the  #-axis. 

It  follows  directly  that  both  the  abscissa  and  ordinate  of  a  point 
situate  at  the  origin  are  zero.  A  point  on  the  #-axis  has  a  zero 
ordinate  ;  a  point  on  the  ?/-axis  has  a  zero  abscissa.  Any  line 
parallel  to  the  #-axis  has  an  equation 

y  =  b;  .         .         .         .         (3) 

any  line  parallel  to  the  ^-axis  has  an  equation 

x  =  a,  .         .        .         .         (4) 

where  a  and  b  denote  the  distances  between  the  two  lines  and  their 

respective  axes. 

It  is  necessary  to  warn  the  reader  not  to  fall  into  the  bad  habit 

of  writing  the  line  OM  indifferently  «  OM"  and  "  MO  "  so  that  he 

will  have  nothing  to  unlearn  later  on.     Lines  measured  from  left 

to  right,  and  from  below  upwards  are  positive ;  negative,  if  measured 

in  the  reverse  directions.      Again,  angles  measured  in  the  opposite 

direction  to  the  motion  of  the  hands  of  a  watch,  when  the  watch  is 

facing  the  reader,  are  positive,  and  negative  if  measured  in  the 

opposite  direction.      Many  difficulties  in  connection  with  optical 

problems,  for  instance,  will  disappear  if  the  reader  pays  careful 

attention  to  this.     In  the  diagram,  the  angle  MOP  will  be  positive, 

POM  negative.      The  line  MP  is  positive,  PM  negative.     Hence, 

since 

4-  MP  -  PM 

tan  MOP  =  ^i^£  =  +  ;  tan  POM  =  — ^  -  -. 
+  OM  +  OM 

§  30.  Practical  Illustrations  of  Graphical  Representation. 

Suppose,  in  an  investigation  on  the  relation  between  the  pres- 
sure, p,  and  the  weight,  w,  of  a  gas  dissolved  by  unit  volume  of  a 
solution,  we  obtained  the  following  successive  pairs  of  observations, 
p  =  i,     2,     4,     8  . . .  =  x. 
to -i,     1,     2,     4...=  y. 


P(8*j 


/ 


§  30.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  87 

By  setting  off  on  millimetre,  coordinate  or  squared  paper 
(Fig.  12)  points  P^l,  i),  P2(2,  1) 
. . . ,  and  drawing  a  line  to  pass 
through  all  these  points,  we  are 
said  to  plot  the  curYe.  This  has 
been  done  in  Fig.  12.  The  only 
difference  between  the  lines  OP 
of  Figs.  11  and  12  is  in  their 
slope  towards  the  two  axes. 

From  equation  (1)  we  can  put  FlG-  12"^£lullt^  °f  GaSGS 

w  =  p  tan  a,  or  tan  a  =  \, 

that  is  to  say,  an  angle  whose  tangent  is  \.     This  can  be  found  by 

reference  to  a  table  of  natural  tangents.     It  is  26°  33'  (approx.). 

Putting  tan  a  =  m,  we  may  write 

w  =  mp, (5) 

where  m  is  a  constant  depending  on  the  nature  of  the  gas  and 
liquid  used  in  the  experiment.  Equation  (5)  is  the  mathematical 
expression  for  the  solubility  of  a  gas  obeying  Henry's  law,  viz.  : 
"  At  constant  temperature,  the  weight  of  a  gas  dissolved  by  unit 
volume  of  a  liquid  is  proportional  to  the  pressure ".  The  curve 
OP  is  a  graphical  representation  of  Henry's  law. 

To  take  one  more  illustration.  The  solubility  of  potassium 
chloride,  X,  in  100  parts  of  water  at  temperatures,  0,  between  0° 
and  100°  is  approximately  as  follows : 

0  =  0°,         20°,         40°,         60°,        80°,         100°  =  x, 

X  =  28-5,     39-7,       49-8,       59-2,       69*5,       79-5  =  y. 

By  plotting  these  numbers,  as  in  the  preceding  example,  we  obtain 

a  curve  QP  (Fig.  13)  which,  instead  of  passing  through  the  origin 

at  O,  cuts  the  ?/-axis  at  the  point  Q  such  that 

OQ  =  28-5  units  =  b  (say). 

If  OP'  be  drawn  from  the  point  O  parallel  to  QP,  then  the  equation 
for  this  line  is  obviously,  from  (5), 

X  =  m6 ; 
but  since  the  line  under  consideration  cuts  the  ?/-axis  at  Q, 

X  =  mO  +  b,      .         .         .         .         (6) 

where  b  =  OQ.  In  these  equations,  b,  X  and  0  are  known,  the 
value  of  m  is  therefore  obtained  by  a  simple  transposition  of  (6), 


88 


HIGHEE  MATHEMATICS. 


§30. 


m  = 


0 


=  tan  27°  43'  -  0-5254. 


Substituting  in  (6)  the  numerical  values  of  m  and  b{=  28*5), l  we 
can  find  the  approximate  solubility  of  potassium  chloride  at  any 
temperature  {$)  between  0°  and  100°  from  the  relation 
X  =  O51280  +  28-5. 
The  curve  QP  in  Fig.  13  is  a  graphical  representation  of  the 


0  20         iO         eo         eo       200 

Fig.  13.— Solubility  Curve  for  KC1  in  water. 

variation  in  the   solubility  of  KC1  in  water  at  different  tempera- 
tures. 

Knowing  the  equation  of  the  curve,  or  even  the  form  of  the 
curve  alone,  the  probable  solubility  of  KC1  for  any  unobserved 
temperature  can  be  deduced,  for  if  the  solubility  had  been  de- 
termined every  10°  (say)  instead  of  every  20°,  the  corresponding 
ordinates  could  still  be  connected  in  an  unbroken  line.  The  same 
relation  holds  however  short  the  temperature  interval.  From  this 
point  of  view  the  solubility  curve  may  be  regarded  as  the  path  of 
a  point  moving  according  to  some  fixed  law.  This  law  is  defined 
by  the  equation  of  the  curve,  since  the  coordinates  of  every  point 
on  the  curve  satisfy  the  equation.  The  path  described  by  such  a 
point  is  called  the  picture,  locus  or  graph  of  the  equation. 

Examples. — (1)  Let  the  reader  procure  some  "  squared  "  paper  and  plot : 
y  =  lx  -  2  ;  2y +  Sx  =  12. 

(2)  The  following  experimental  results  have  been  obtained  : — 
When  x  =      0,  1,  10,  20,  30,... 

y  =  -  3,  1-56,  11-40,  25-80,  40-20, . . . 


1  Determined  by  a  method  to  be  described  later. 


§  31.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  89 

(a)  Plot  the  curve,  (b)  Show  (i)  that  the  slope  of  the  curve  to  the  sc-axis 
is  nearly  1*44  =  tan  o  =  tan  55°,  (ii)  that  the  equation  to  the  curve  is 
y  =  l-44aj  -  3.  (c)  Measure  off  5  and  15  units  along  the  ic-axis,  and  show 
that  the  distance  of  these  points  from  the  curve,  measured  vertically  above 
the  a>axis,  represents  the  corresponding  ordinates.  (d)  Compare  the  values 
of  y  so  obtained  with  those  deduced  by  substituting  x  =  5  and  x  =  15  in  the 
above  equation.  Note  the  laborious  and  roundabout  nature  of  process  (c)  when 
contrasted  with  (d).  The  graphic  process,  called  graphic  interpolation  (q.v.), 
is  seldom  resorted  to  when  the  equation  connecting  the  two  variables  is 
available,  but  of  this  anon. 

(3)  Get  some  solubility  determinations  from  any  chemical  text-book  and 
plot  the  values  of  the  composition  of  the  solution  (C,  ordinate)  at  different 
temperatures  (0°,  abscissa),  e.g.,  Loewel's  numbers  for  sodium  sulphate  are 

C  =  5-0,        19-4,       550,       46-7,        44-4,        43-1,        42-2; 

6°  =  0°,  20°,         34°,  50°,  70°,  90°,         103*5°. 

What  does  the  peculiar  bend  at  34°  mean  ? 

In  this  and  analogous  cases,  a  question  of  this  nature  has  to  be  decided : 
What  is  the  best  way  to  represent  the  composition  of  a  solution?  Several 
methods  are  available.  The  right  choice  depends  entirely  on  the  judgment, 
or  rather  on  the  finesse,  of  the  investigator.  Most  chemists  (like  Loewel 
above)  follow  Gay  Lussac,  and  represent  the  composition  of  the  solution  as 
"parts  of  substance  which  would  dissolve  in  100  parts  of  the  solvent". 
Etard  found  it  more  convenient  to  express  his  results  as  "  parts  of  substance 
dissolved  in  100  parts  of  saturated  solution  ".  The  right  choice,  at  this  day, 
seems  to  be  to  express  the  results  in  molecular  proportions.  This  allows  the 
solubility  constant  to  be  easily  compared  with  the  other  physical  constants. 
In  this  way,  Gay  Lussac's  method  becomes  "  the  ratio  of  the  number  of 
molecules  of  dissolved  substance  to  the  number,  say  100,  molecules  of 
solvent  "  ;  Etard's  "  the  ratio  of  the  number  of  molecules  of  dissolved  sub- 
stance to  any  number,  say  100,  molecules  of  solution  ". 

(4)  Plot  logex  =  y,  and  show  that  logarithms  of  negative  numbers  are 
impossible.  Hint.  Put  x  =  0,  e~2,  e-1,  1,  e,  e2,  oo  ,  etc.,  and  find  correspond- 
ing values  of  y. 

So  many  good  booklets  have  recently  been  published  upon 
"  Graphical  Algebra  "  as  to  render  it  unnecessary  to  speak  at  greater 
length  upon  the  subject  here. 

§  31.    Properties  of  Straight  Lines. 

If  equations  (1)  and  (6)  be  expressed  in  general  terms,  using 
x  and  y  for  the  variables,  m  and  b  for  the  constants,  we  can 
deduce  the  following  properties  for  straight  lines  referred  to  a  pair 
of  coordinate  axes. 

I.  A  straight  line  passing  through  the  origin  of  a  pair  of 
rectangular  coordinate  axes,  is  represented  by  the  equation 

y  =  mx,  .        .         .         .         (7) 


90  HIGHEK  MATHEMATICS.  §  3L 

where  m  =  tan  a  =  y/x,  a  constant  representing  the  slope  of  the 
curve.     The  equation  is  obtained  from  (5)  above. 

II.  A  straight  line  which  cuts  one  of  the  rectangular  coordinate 
axes  at  a  distance  b  from  the  origin,  is  represented  by  the  equation 

y  =  mx  +  b  .         .         .         (8) 

where  m  and  b  are  any  constants  whatever.  For  every  value  of 
m  there  is  an  angle  such  that  tan  a  =  ra.  The  position  of  the  line 
is  therefore  determined  by  a  point  and  a  direction.  Equation  (8) 
follows  immediately  from  (G). 

III.  A  straight  line  is  always  represented  by  an  equation  of  the 
first  degree, 

Ax  +  By  +  G  =  0 ;        .        .        .         (9) 

and  conversely,  any  equation  of  the  first  degree  between  two  variables 
represents  a  straight  line.1 

This  conclusion  is  drawn  from  the  fact  that  any  equation 
containing  only  the  first  powers  of  x  and  y,  represents  a  straight 
line.  By  substituting  m  =  -  A/B  and  b  =  -  G/B  in  (8),  and 
reducing  the  equation  to  its  simplest  form,  we  get  the  general 
equation  of  the  first  degree  between  two  variables  :  Ax  +  By  +  0  =  0. 
This  represents  a  straight  line  inclined  to  the  positive  direction  of 
the  ic-axis  at  an  angle  whose  tangent  is  -  A/B,  and  cutting  the 
y-Sbxis  at  a  point  -  G/B  below  the  origin. 

IV.  A  straight  line  which  cuts  each  coordinate  axis  at  the  re- 
spective distances  a  and  b  from  the  origin,  is  represented  by  the 
equation 

ifl-l     •      •      •      •     do) 

Consider  the  straight  line  AB  (Fig.  14)  which  intercepts  the 
x-  and  2/-axes  at  the  points  A  and  B  respectively.  Let  OA  =  a 
OB  =  b.     From  the  equation  (9)  if 

y  =  0,  x  =  a  ;  Aa  +  G  =  0,  a  =  -  C/A. 
Similarly  if     x  =  0,  y  =  b;  Bb  +  G  =  0,  b  =  -  G/B. 
Substituting  these  values  of  a  and  b  in  (9),  i.e.,  in 
A        B        -  x      y      ., 

-  -7&  ~  -qV  =  J-  J  and  we  get  -  +   y  =  1. 

1The  reader  met  with  the  idea  conveyed  by  a  "general  equation,"  on  page 
26.  By  assigning  suitable  values  to  the  constants  A,  B,  O,  he  will  be  able  to  deduce 
every  possible  equation  of  the  first  degree  between  the  two  variables  x  and  y. 


§  31.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  91 

There  are  several  proofs  of  this  useful  equation.     Formula  (10)  ia 
called  the  intercept  form  of  the  equa-      y 
tion  of  the  straight   line,  equation  (8) 
the  tangent  form. 

V.  The  so-called  normal  or  per- 
pendicular form  of  the  equation  of  a 
straight  line  is 

p  =  X  COS  a  +  y  COS  a,     .      (11) 

where  p  denotes  the  perpendicular  dis- 
tance of  the  line  BA  (Fig.  14)  from  the 
origin  0,  and  a  represents  the  angle 
which  this  line  makes  with  the  rc-axis. 

Draw  OQ  perpendicular  to  AB  (Fig.  14).  Take  any  point  P{x,  y)  and 
drop  a  perpendicular  PR  on  to  the  z-axis,  draw  RD  parallel  to  AB  cutting  OQ 
in  D.  Drop  PC  perpendicular  on  to  RD,  then  PRC  =  a  =  QOA.  Then, 
OQ  =  OD  +  PC  OD  =  x  cos  o ;  PC  =  y  sin  a.     Hence  follows  (11). 

Many  equations  can  be  readily  transformed  into  the  intercept 
form  and  their  geometrical  interpretation  seen  at  a  glance.  For 
instance,  the  equation 

X  +  y  =  2  becomes  \x  +  \y  =  1, 

which  represents  a  straight  line   cutting  each  axis  at  the  same 
distance  from  the  origin. 

One  way  of  stating  Charles'  law  is  that  "  the  volume  of  a 
given  mass  of  gas,  kept  at  a  constant  pressure,  varies  directly  as 
the  temperature  ".  If,  under  these  conditions,  the  temperature  be 
raised  6°,  the  volume  increases  the  -zfaOrd  part  of  what  it  was  at 
the  original  temperature.1     Let  the  original  volume,  v0,  at  0°  C, 

1  Many  students,  and  even  some  of  the  text-books,  appear  to  have  hazy  notions  on 
this  question.  According  to  u  Guy  Lussac's  law  "  the  increase  in  the  volume  of  a  gas 
at  any  temperature  for  a  rise  of  temperature  of  1°,  is  a  constant  fraction  of  its  initial 
volume  at  0°  C. ;  "  J.  Dalton's  law  "  {Manchester  Memoirs,  5,  595,  1802),  on  the  other 
hand,  supposes  the  increase  in  the  volume  of  a  gas  at  any  temperature  for  a  rise  of  1°, 
is  a  constant  fraction  of  its  volume  at  that  temperature  (the  "  Compound  Interest 
Law,"  in  fact).  The  former  appears  to  approximate  closer  to  the  truth  than  the  latter. 
(See  page  285.)  J.  B.  Gay  Lussac  {Annates  de  Chimie,  43, 137;  1802)  says  that  Charles 
had  noticed  this  same  property  of  gases  fifteen  years  earlier  and  hence  it  is  sometimes 
called  Charles'  law,  or  the  law  of  Charles  and  Gay  Lussac.  After  inspecting  Charles' 
apparatus,  Gay  Lussac  expressed  the  opinion  that  it  was  not  delicate  enough  to  es* 
tablish  the  truth  of  the  law  in  question.  But  then  J.  Priestley  in  his  Experiments  and 
Observations  on  Different  Kinds  of  Air  (2,  448,  1790)  says  that  "  from  a  very  coarse 
experiment  which  I  made  very  early  I  concluded  that  fixed  and  common  air  expanded 


92 


HIGHER  MATHEMATICS. 


§31. 


be  unity  ;  the  final  volume  v,  then  at  0° 


2T3( 


This  equation  resembles  the  intercept  form  of  the  equation  of  a 
straight  line  (10)  where  a  =  -  273  and  6  =  1.  The  intercepts  a  and 
b  may  be  found  by  putting  x  and  y,  or  rather  their  equivalents, 


-273°C 


0  and  v,  successively  equal  to  zero.  If  0  =  0,  v  =  1 ;  if  v  =  0, 
0  =  -  273,  the  well-known  absolute  zero  (Fig.  15). 

It  is  impossible  to  imagine  a  substance  occupying  no  space. 
But  this  absurdity  in  the  logical  consequence  of  Charles' 
law  when  0  =  -  273°.  Where  is  the  fallacy  ?  The  answer  is  that 
Charles'  law  includes  a  "  simplifying  assumption ".  The  total 
volume  occupied  by  the  gas  really  consists  of  two  parts :  (i)  the 
volume  actually  occupied  by  the  molecules  of  the  substance ;  and 
(ii)  the  space  in  which  the  molecules  are  moving.  Although  we 
generally  make  v  represent  the  total  volume,  in  reality,  v  only  refers 
to  the  space  in  which  the  molecules  are  moving,  and  in  that  case 
the  conclusion  that  v  =  0,  when  0  =  -  273°  involves  no  absurdity. 

No  gas  has  been  investigated  at  temperatures  within  four  degrees 
of  -  273°.     However  trustworthy  the  results  of  an  interpolation 


Fig.  16. 


may  be,  when  we  attempt  to  pass  beyond  the  region  of  measure- 


alike  with  the  same  degree  of  heat  ".  The  cognomen  "Priestley's  law  "  would  settle 
all  confusion  between  the  three  designations  "Dalton's,"  "Gay  "Lussac's "  and 
"  Charles'  "  of  one  law. 


§  32.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  93 

ment,  the  extrapolation,  as  it  is  called,  becomes  more  or  less 
hazardous.  Extrapolation  can  only  be  trusted  when  in  close  prox- 
imity to  the  point  last  measured.  Attempts  to  find  the  probable 
temperature  of  the  sun  by  extrapolation  have  given  numbers 
varying  between  the  1,398°  of  Vicaire  and  the  9,000,000°  of 
Waterston  !  We  cannot  always  tell  whether  or  not  new  forces 
come  into  action  when  we  get  outside  the  range  of  observation. 
In  the  case  of  Charles'  law,  we  do  know  that  the  gases  change 
their  physical  state  at  low  temperatures,  and  the  law  does  not 
apply  under  the  new  conditions. 

VI.  To  find  the  angle  at  the  point  of  intersection  of  two  curves 
whose  equations  are  given.     Let  the  equations  be 

y  =  mx  +  b ;  y'  =  mx'  +  b'. 
Let  <£  be  the  angle  required  (see  Fig.  16),  m  =  tana,  m'  —  tana'. 
From  Euclid,  i.,  32,  a  -  a  =  <£,  .-.  tan  (a'  -  a)  =  tan  <j>.    By  formula, 
page  612, 

tana'  -  tana         m'  -  m 
n  ^  =  1  +  tan  a  .  tan  a  ~~  1  +  mm'  ^     ' 

Examples. — (1)  Find  the  angle  at  the  point  of  intersection  of  the  two 
lines  x  +  y  =  1,  and  y  =  x  +  2.     m  =  1,  m'  =  -  1 ;  tan  <p  =  -  oo  =  -  90°. 

(2)  Find  the  angle  between  the  lines  3y  -  x  =  0,  and  2x  +  y  =  1.  Ansr. 
Tan  (81°  52')  =  7. 

VII.  To  find  the  distance  between  two  points  in  terms  of  their 
coordinates.    In  Fig.  17,  let  P^y^  and  Q(x2y2)  be  the  given  points. 

Draw  QM'  parallel  to  NM.     OM=xv  MP  =  yY\  ON= x2,  NQ  =  y2; 
MP  =  MP  -  MM  =  MP  ~NQ  =  yl-y2\ 
QM  =  NM  =OM  -  ON  =  x1  -  xv 
Since  QMP  is  a  right-angled  triangle 

(QPf  =  (QM')*  +  {PMf. 

.-.  QP  =  J(x,  -  x2f  +  (Vl  -  y2)\    .        .        (13) 

Examples. — (1)  Show  that  the  distance  between  the  points  ( -  2, 1)  and 
( -  6,  -  2)  is  5  units. 

(2)  Show  that  the  distance  from  (10,  -  18)  to  the  point  (3,  6)  is  the  same 
as  to  the  point  (-  5,  2).     Ansr.  25  units  in  each  case. 

§  32.    Curves  Satisfying  Conditions. 

The  reader  should  work  through  the  following  examples  so  as 
to  familiarize  himself  with  the  conceptions  of  coordinate  geometry. 
Many  of  the  properties  here  developed  for  the  straight  line  can 
easily  be  extended  to  curved  lines. 


94  HIGHER  MATHEMATICS.  §  32. 

I.  The  condition  that  a  curve  may  pass  through  a  given  point. 
This  evidently  requires  that  the  coordinates  of  the  point  should  satisfy 
the  equation  of  the  line.     Let  the  equation  be  in  the  tangent  form 

y  =  mx  +  b. 
If  the  line  is  to  pass  through  the  point  (xv  y^)t 

y1  =  mx1  +  by 
and,  by  subtraction, 

(y  -  yi)  =  m(x  -  xx)  .  .  .  (15) 
which  is  an  equation  of  a  straight  line  satisfying  the  required 
conditions. 

Examples. — (1)  The  equation  of  a  line  passing  through  a  point  whose 
abscissa  is  5  and  ordinate  3  is  y  -  mx  =  3  -  5m. 

(2)  Find  the  equation  of  a  line  which  will  pass  through  the  point  (4,  -  4) 
and  whose  tangent  is  2.     Ansr.   y  -  2x  +  12  =  0. 

II.  The  condition  that  a  curve  may  pass  through  two  given 
points.  Continuing  the  preceding  discussion,  if  the  line  is  to  pass 
through  (x2,  y2),  substitute  x2,  y2,  in  (14) 

(2/2  -  Vi)  =  ™(*2  -  *i) ;  ••.'»■-  l4r|-; 

Substituting  this  value  of  m  in  (14),  we  get  the  equation, 

y  ~  yi  =  x  ~  xit        ,         .         .         (15) 

2/2   ~~  Vl         X2   ~  Xl 

for  a  straight  line  passing  through  two  given  points  (xv  yj  and 

(*s»  2/2)- 

Examples. — (1)  Show  that  the  equation  of  the  straight  line  passing 
through  the  points  Pa(2,  3)  and  P2(4,5)  is  x  -  y  +  1  =  0.  Hint.  Substitute 
x1  =  2,x2  =  4:)y1  =  3,  y2  =  5,  in  (15). 

(2)  Find  the  equation  of  the  line  which  passes  through  the  points 
Px(4,  -  2),  and  P2(0,  -  7).     Ansr.    5x  -  ±y  =  28. 

III.  The  coordinates  of  the  point  of  intersection  of  two  given 
lines.     Let  the  given  equations  be 

y  =  mx  +  b;  and  y  =  m'x  +  b' . 

Now  each  equation  is  satisfied  by  an  infinite  number  of  pairs  of 
values  of  x  and  y.  These  pairs  of  values  are  generally  different 
in  the  two  equations,  but  there  can  be  one,  and  only  one  pair  of 
values  of  x  and  y  that  satisfy  the  two  equations,  that  is,  the 
coordinates  of  the  point  of  intersection.  The  coordinates  at  this 
point  must  satisfy  the  two  equations,  and  this  is  true  of  no  other 
point.     The  roots  of  these  two  equations,  obtained  by  a  simple 


§  32.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  95 

algebraic  operation,  are  the  coordinates  of  the  point  required.     The 
point  whose  coordinates  are 

V  -  b  b'm  -  bm  f^ax 

x  =  -, ;  y  =  —        .         .         (lo) 

m  -  m     u         m  -  m 

satisfies  the  two  equations. 

Examples. — (1)  Find  the  coordinates  of  the  point  of  intersection  of  the 
two  lines  x  +  y  =  1,  and  y  =  x  +  2..  Ansr.  x  =  -  £,  y  =  § .  Hint,  m  =  -  1, 
ml  =  1§  b  =  1,  V  m  2,  etc. 

(2)  The  coordinates  of  the  point  of  intersection  of  the  curves  Sy  -  x  =  1, 
and  2x  +  y  =  3  are  x  =  f ,  y  =  f . 

(3)  Show  that  the  two  curves  y2  =  4a:,  and  x1  =  4=2/  meet  at  the  point 
x  =  4,  y  =  4. 

IV.  The  condition  that  three  given  lines  may  meet  at  a  point. 
The  roots  of  the  equations  of  two  of  the  lines  are  the  coordinates  of 
their  point  of  intersection,  and  in  order  that  this  point  may  be  on  a 
third  line  the  roots  of  the  equations  of  two  of  the  lines  must  satisfy 
the  equation  of  the  third. 

Examples. — (1)  If  three  lines  are  represented  by  the  equations  5x  +  3y=7t 
Sx  -  4ty  =  10,  and  x  +  2y  =  0,  show  that  they  will  all  intersect  at  a  point 
whose  coordinates  are  x  =  2  and  y  =  -  1.  Solving  the  last  two  equations, 
we  get  x  =  2  and  y  =  -  1,  but  these  values  of  x  and  y  satisfy  the  first  equation, 
hence  these  three  lines  meet  at  the  point  (2,  -  1). 

(2)  Show  that  the  lines  3z  +  5y  +  7=0;  x  +  2y  +  2=0;  and  ix-By- 10  =  0, 
do  not  pass  through  one  point.  Hint.  From  the  first  and  second,  y  =  1, 
x  =  -  4.     These  values  do  not  satisfy  the  last  equation. 

V.  The  condition  that  two  straight  lines  may  be  parallel  to  one 
another.  Since  the  lines  are  to  be  parallel  they  must  make  equal 
angles  with  the  a;-axis,  i.e.,  angle  a  =  angle  a,  or  tan  a  =  tan  a, 

.'.m  =  m\  .  .  .  .  (17) 
that  is  to  "say,  the  coefficient  of  x  in  the  two  equations  must  be 
equal. 

Examples. — (1)  Show  that  the  lines  y  =  Sx  +  9,  and  2y  =  6x  +  7  are 
parallel.  Hint.  Show  that  on  dividing  the  last  equation  by  2,  the  coefficient 
of  x  in  each  equation  is  the  same. 

(2)  Find  the  equation  of  the  straight  line  passing  through  (2,  -  1)  parallel 
ioSx  +  y  =  2.     Ansr.  y  +  3a;  =  5.    Hint.  Use  (17)  and  (14).  y  +  1  =  -  3  (x  -  2). 

VI.  The  condition  that  two  lines  may  be  perpendicular  to  one 
another.     If  the  angle  between  the  lines  is  <£  =  90°,  see  (12), 

a!  -  a  =  90°, 

1 

.•.  tan  a  =  tan  (90  +  a)  =  -  cot  a  =  -  7—— » 


96  HIGHER  MATHEMATICS.  §  33. 

.-.  m  =  -  -,  .         .         .         .         (18) 
m 

or,  the  slope  of  the  one  line  to  the  £-axis   must  be  equal  and 

opposite  in  sign  to  the  reciprocal  of  the  slope  of  the  other. 

Examples. — (1)  Find  the  equation  of  the  line  which  passes  through  the 
point  (3,  2),  and  is  perpendicular  to  the  line  y  =  2x  +  5.  Ansr.  x  +  2y  =  7. 
Hint.  Use  (18)  and  (14). 

(2)  Find  the  equation  of  the  line  which  passes  through  the  point  (2,  -  4) 
and  is  perpendicular  to  the  line  Sy  +  2x  -  1  =  0.     Ansr.  2y  -  3x  +  14  =  0. 

§  33.   Changing  the  Coordinate  Axes. 

In  plotting  the  graph  of  any  function,  the  axes  of  reference 
should  be  so  chosen  that  the  resulting  curve  is  represented  in  the 
most  convenient  position.     In  many  problems  it  is  necessary  to 

pass  from  one  system  of  coordinate  axes 
to  another.  In  order  to  do  this  the 
equation  of  the  given  line  referred  to 
the  new  axes  must  be  deduced  from  the 
corresponding  equation  referred  to  the 
old  set  of  axes. 

I.    To  pass  from  any  system  of  co- 
ordinate axes  to  another  set  parallel  to 
Fig.  18.— Transformation  of   the  former  but  having  a  different  origin. 
Axes-  Let  Ox,  Oy  (Fig.  18)  be  the  original  axes, 

and  EO^,  HO-^y,  the  new  axes  parallel  to  Ox  and  Oy.  Let  MMXP 
be  the  ordinate  of  any  point  P  parallel  to  the  axes  Oy  and  0^. 
Let  h,  k  be  the  coordinates  of  the  new  origin  Ox  referred  to  the  old 
axes.  Let  (x,  y)  be  the  coordinates  of  P  referred  to  the  old  axes 
Ox,  Oy,  and  {xYy^)  its  coordinates  referred  to  the  new  axes.  Then 
OH  =  h,  HO,  =  k, 

x  =  OM  =  OH  +  HM  =  OH  +  0^1,  =  h  +  xY ; 
y  =  MP  =  MM1  +  MXP  =  H01  +  MXP  =  k  +  yv 
That  is  to  say,  we  must  substitute 

x  =  h  +  xY ;  and  y  =  k  +  yv  .  .  (19) 
in  order  to  refer  a  curve  to  a  new  set  of  rectangular  axes.  The 
new  coordinates  of  the  point  P  being 

xY  =  x  -  h ;  and  y1  =  y  -  k.         .         .         (20) 

Example. — Given  the  point  (2,  3)  and  the  equation  2x  +  By  =  6,  find  the 
coordinates  of  the  former,  and  the  equation  of  the  latter  when  referred  to  a 
set  of  new  axes  parallel  to  the  original  axes  and  passing  through  the  point 


y 

y 

,p 

K 

.27 

o.: 

|m, 

yt 

h 

JT 

0 

t 

\ 

M 

§34. 


COORDINATE  OR  ANALYTICAL  GEOMETRY. 


97 


Fig.  19. — Transformation  of  Axes. 


3,  2).  Ansr.  a^  =  3-3  =  2  -  3=  -  1;  y1=y  -2  =  1.  The  position  of  the 
point  on  the  new  axes  is  ( -  1,  1).  The  new  equation  will  be  2(3  +  xj  + 
3(3  +  y1)  =  0;  .-.  2xx  +  3Vl  +  12  =  0. 

II.  To  pass  from  one  set  of  axes  to  another  having  the  same 
origin  but  different  directions. 
Let  the  two  straight  lines  xY0 
and  yxO,  passing  through  0  (Fig. 
19),  be  taken  as  the  new  system 
of  coordinates.  Let  the  coordin- 
ates of  the  point  P  (x,  y)  when 
referred  to  the  new  axes  be  xv 
yv  Draw  MP  perpendicular  to 
the  old  #-axes,  and  MYP  perpen- 
dicular to  the  new  axes,  so  that 
the  angle  MPM1  =  BOMl  =  a, 

OM  =  x}  OM^  =  xv  MP  =  y,  MYP  =  yv 
Draw  BMX  perpendicular  and  QM1  parallel  to  the  #-axis.     Then 
x=  OM=  OB  -  MR  =  OB  -  QMV 
.*.  x  =  OMY  cos  a  -  MXP  sin  a  ; 

.-.  x  =  x1  cos  a  -  yY  sin  a. .  .         .       (21) 

Similarly  y  =  MP  =  MQ  +  QP  =  BMY  +  QP ; 

.\  y  =  0M1  sin  a  +  MXP  cos  a, 

.*.  y  =  x1  sin  a  +  2/j  cos  a.    .         .         .       (22) 
Equations  (21)  and  (22)  enable  us  to  refer  the  coordinates  of  a 
point  P.  from  one  set  of  axes  to  another.     Solving  equations  (21) 
and  (22)  simultaneously, 

xx  —  x  cos  a  +  y  sin  a  ;  yx  =  y  cos  a  -  a;  sin  a.  .  (23) 
Example. — Find  what  the  equation  xf  -  yx2  =  a2  becomes  when  the 
axes  are  turned  through  -  45°,  the  origin  remaining  the  same.  Here 
sin  (-  45°)  =-\/J;  cos  (-  45°)  =  \/ J".  From  (23),  a^  =  *J$x  -  \% ; 
Vl  =  sl%x+  sj\y.  Hence,  xx  -  yx=  -  J2x;  x^+ij^J^x  ; .-.  «12-2/]2=  -2xy, 
.*.  from  the  original  equation,  2xy  =  -  a2  ;  or,  xy  =  constant. 

In  order  to  pass  from  one  set  of  axes  to  another  set  having  a 
different  origin  and  different  directions,  the  two  preceding  transfor- 
mations must  be  made  one  after  another. 


§  3$.    The  Circle  and  its  Equation; 

There  is  a  set  of  important  curves  whose  shape  can  be  obtained 
by  cutting  a  cone  at  different  angles.     Hence  the  name  conic  sec- 


98 


HIGHER  MATHEMATICS. 


34. 


tions.     They  include  the  parabola,  hyperbola  and  ellipse,  of  which 

the  circle  is  a  special  case.  I 
shall  describe  their  chief  pro- 
perties very  briefly. 

A  circle  is  a  curve  such 
that  all  points  on  the  curve 
are  equi-distant  from  a  given 
point.  This  point  is  called  the 
centre,  the  distance  from  the 
centre  to  the  curve  is  called 
the  radius.  Let  r  (Fig.  20)  be 
the  radius  of  the  circle  whose 
centre  is  the  origin  of  the  rect- 
angular coordinate  axes  xOx' 
and    yOy' .      Take    any  point 

P(x,  y)  on  the  circle.     Let  PM  be  the  ordinate  of  P.     From  the 

definition  of  a  circle   OP  is  constant  and  equal  to  r.     Then  by 

Euclid,  i.,  47, 


Fig.  20.— The  Circle. 


(OMf  +  (MPf  =  (OP)2,  or  x2  +  y2  =  r2 


(1) 


which  is  said  to  be  the  equation  of  the  circle. 

In  connection  with  this  equation  it  must  be  remembered  that 
the  abscissae  and  ordinates  of  some  points  have  negative  values, 
but,  since  the  square  of  a  negative  quantity  is  always  positive,  the 
rule  still  holds  good.  Equation  (1)  therefore  expresses  the  geo- 
metrical fact  that  all  points  on  the  circumference  are  at  an  equal 
distance  from  the  centre. 

Examples. — (1)  Required  the  locus  of  a  point  moving  in  a  path  according 
to  the  equations  y  =  a  cos  t,  x  =  a  sin  t,  where  t  denotes  any  given  interval  of 
time.     Square  each  equation  and  add, 


+  xa 


a2(cos2*  +  sin2£). 


The  expression  in  brackets  is  unity  (19),  page  611,  and  hence  for  all  values  of  t 

y2  +  x*  m  a2, 
i.e.,  the  point  moves  on  the  perimeter  of  a  circle  of  radius  a. 

(2)  To  find  the  equation  of  a  circle  whose  centre,  referred  to  a  pair  of 
rectangular  axes,  has  the  coordinates  h  and  k.     From  (19),  previous  paragraph, 

(x  -  hf  +  (y  -  fe)2  =  r2,  .  .  .  .  (2) 
where  P(sc,  y)  is  any  point  on  the  circumference.  Note  the  product  xy  is 
absent.  The  coefficients  of  re2  and  y2  are  equal  in  magnitude  and  sign. 
These  conditions  are  fulfilled  by  every  equation  to  a  circle.     Such  ia 

3a2  +  3#2  +  7<e  -  12  =  0. 

(3)  The  general  equation  of  a  circle  is 


x2  +  y%  +  ax  +  by  +  c  =»  0. 


(3) 


§35. 


COORDINATE  OR  ANALYTICAL  GEOMETRY. 


99 


Plot  (3)  on  squared  paper.  Try  the  effect  of  omitting  ax  and  of  by  separately 
and  together.  This  is  a  sure  way  of  getting  at  the  meaning  of  the  general 
equation. 

(4)  A  point  moves  on  a  circle  x2  +  y*  =  25.  Compare  the  rates  of  change 
of  x  and  y  when  x  =  3.  If  x  =  3,  obviously  y  =  ±  4.  By  differentiation 
dy/dt :  dxjdt  =  -  xjy  =  +  f .  The  function  decreases  when  x  and  y  have  the 
same  sign,  i.e.,  in  the  first  and  third  quadrants,  and  increases  in  the  second 
and  fourth  quadrants ;  y  therefore  decreases  or  increases  three-quarters  as 
fast  as  x  according  to  the  quadrant. 

§  35.    The  Parabola  and  its  Equation. 

A  parabola  is  a  curve  such  that  any  point  on  the  curve  is  equi- 
distant from  a  given  point  and  a  given  straight  line.  The  given 
point  is  called  the  focus,  the  straight  line 
the  directrix,  the  distance  of  any  point  on 
the  curve  from  the  focus  is  called  the  focal 
radius.  O  (Fig.  21)  is  called  vertex  of  the 
parabola.  AK  is  the  directrix ;  OF,  FP, 
FPX ...  are  focal  radii ;  OF  =  AO  ;  FP  - 
KP ;  FPY  =  KXPY ...  It  can  now  be  proved 
that  the  equation  of  the  parabola  is  given  by 
the  expression 

y2  m  ±ax,       .         .         (1) 
where  a  is  a  constant  equal  to  AO  in  the 
above  diagram.     In  words,  this  equation  tells  us  that  the  abscissse 
of  the  parabola  are  proportional  to  the  square  of  the  ordinates. 

Examples. — (1)  By  a  transformation  of  coordinates  show  that  the  para- 
bola represented  by  equation  (1),  may  be  written  in  the  form 

x  =  a  +  by  +  cy2, '    (2) 

where  a,  b,  c,  are  constants.  Let  x  become  x  +  h;  y  =  y  +  k;  a=j  where  h,  k, 
and  j  are  constants.  Substitute  the  new  values  of  x  and  y  in  (1) ;  multiply 
out.  Collect  the  constants  together  and  equate  to  a,  b  and  c  as  the  case 
might  be. 

(2)  Investigate  the  shape  of  the  parabola.  By  solving  the  equation  of 
the  parabola,  it  follows  that 

V  =  ±  2  ^ lax. 

First.  Every  positive  value  of  x  gives  two  equal  and  opposite  values  of 
y,  that  is  to  say,  there  are  two  points  at  equal  distances  perpendicular  to  the 
ar-axis.  This  being  true  for  all  values  of  x,  the  part  of  the  curve  lying  on  one 
Bide  of  the  #-axis  is  the  mirror  image  of  that  on  the  opposite  side  l ;  in  this 


Fig.  21. 


1  The  student  of  stereo-chemistry  would  say  the  two  sides  were  "  enantiomorphic  ". 


100 


HIGHER  MATHEMATICS. 


§36. 


case  the  sc-axis  is  said  to  be  symmetrical  with  respect  to  the  parabola.  Hence 
any  line  perpendicular  to  the  a>axis  cuts  the  curve  at  two  points  equidistant 
from  the  o>axis. 

Second.  When  x  —  0,  the  y-axis  just  touches 1  the  curve. 

Third.  Since  a  is  positive,  when  x  is  negative  there  is  no  real  value  of  y, 
for  no  real  number  is  known  whose  square  is  negative  ;  in  consequence,  the 
parabola  lies  wholly  on  the  right  side  of  the  y-&xis. 

Fourth.  As  x  increases  without  limit,  y  approaches  infinity,  that  is  to  say, 
the  parabola  recedes  indefinitely  from  the  x  or  symmetrical-axes  on  both 
sides. 

§  36.    The  Ellipse  and  its  Equation. 

An  ellipse  is  a  curve  such  thai  the  sum  of  the  distances  of  any 
point  on  the  curve  from  two  given  points  is  always  the  same.  Let 
P  (Fig.  22)  be  the  given  point  which  moves  on  the  curve  PPX 
so  that  its  distance  from  the  two  fixed  points  Fv  F2,  called  the 


R 

y 

a/ 

b 

'■.a 

ac'l 

7r 

1? 

rL 

Fl 

0 

JC 

J^M 

|f+ 

a 

i 

"1 

rF" 

Fig.  22.— The  Ellipse. 

foci,  has  a  constant  value  say  2a.  The  distance  of  P  from  either 
focus  is  called  the  focal  radius,  or  radius  vector.  0  is  the  so-called 
centre  of  the  ellipse.     The  equation  of  the  ellipse 


£  + 1  - 1 


(1) 


can  now  be  deduced  from  the  above  described  properties  of  the 
curve.  The  line  P^P^  (Fig.  22)  is  called  the  major  axis  ;  P^P 6  the 
minor  axis,  their  respective  lengths  being  2a  and  2b  ;  the  magni- 
tudes a  and  b  are  the  semi-axes  ;  each  of  the  points  Pv  P2,  Pz,  P4, 
is  a  vertex. 

Examples. — (1)  Let  the  point  P(x,  y)  move  on  a  curve  so  that  the  position 


1Some  mathematicians  define  a  "  tangent"  to  be  a  straight  line  which  cuts  the 
curve  in  two  coincident  points.    See  The  School  World,  6,  323,  1904. 


§37.        COORDINATE  OR  ANA&YlTOAL  GE;")M  FIHV.  101 

of  the  point,  at  any  moment,  is  given  by  the  equations,  x  =  a  cos  t  and 
y  =  b  sin  t ;  required  the  path  described  by  the  moving  point.  Square  and 
add  ;  since  cos2*  +  sin2*  is  unity  (page  611),  x2ja2  +  y2jb2  =  1.  The  point 
therefore  moves  on  an  ellipse. 

(2)  Investigate  the  shape  of  the  ellipse.     By  solving  the  equation  of  the 

ellipse  we  get  

/  "ZS  I  »,2 

•         •         (2) 


y  -  ±  b^l  -  ^;  and ^  =  ±a^Jl  -  t 


First.  Since  y2  must  be  positive,  x^a2  "%>  1,  that  is  to  say,  x  cannot  be 
numerically  greater  than  a.  Similarly  it  can  be  shown  that  y  cannot  be 
numerically  greater  than  b. 

Second.  Every  positive  value  of  x  gives  two  equal  and  opposite  values  of 
y,  that  is  to  say,  there  are  two  points  at  equal  distances  perpendicularly  above 
and  below  the  a>axis.  The  ellipse  is  therefore  symmetrical  with  respect  to 
the  <c-axis.  In  the  same  way,  it  can  be  shown  that  the  ellipse  is  symmetrical 
with  respect  to  the  y-axis. 

Third.  If  the  value  of  x  increases  from  the  zero  until  x  =  ±  a,  then  y=0, 
and  these  two  values  of  x  furnish  two  points  on  the  aj-axis.  If  x  now  increases 
until  x  >  a,  there  is  no  real  corresponding  value  of  y2.  Hence  the  ellipse 
lies  in  a  strip  bounded  by  the  limits  x  =  +  a  ;  similarly  it  can  be  shown  that 
the  ellipse  is  bounded  by  the  limits  y  =  +  b. 

Obviously,  if  a  =6,  the  equation  of  the  ellipse  passes  into  that  of  a  circle. 
The  circle  is  thus  a  special  case  of  the  ellipse. 

The  absence  of  first  powers  of  x  and  y  in  the  equation  of  the  ellipse  shows 
that  the  origin  of  the  coordinates  is  at  the  "  centre  "  of  the  ellipse.  A  term  in 
xy  shows  that  the  principal  axes — major  and  minor — are  not  generally  the 
x-  and  z/-axes. 


§  37.    The  Hyperbola  and  its  Equation. 

The  hyperbola  is  a  curve  stich  that  the  difference  of  the  distance 
of  any  point  on  the  curve 
from  two  fixed  points  is  al- 
ways the  same.  Let  the  point 
P  (Fig.  23)  move  so  that  the 
difference  of  its  distances 
from  two  fixed  points  F,  F't 
called  the  foci,  is  equal  to  2a. 
0  is  the  so-called  centre  of 
the  hyperbola  ;  OM  =  x ; 
MP  =  y;OA  =  a;OB  =  b.  FlGL  23-~The  Hyperbola. 

Starting  from  these  definitions  it  can  be  shown  that  the  equation  of 
the  hyperbola  has  the  form 

a2    b* {1) 


+y 

^> 

4 

r 

\\R 

p    ^ 

f/ 

/ 

-X 

V 

+x 

A/    1 

K«a 

\ 

M 

/XR' 

B'          £ 

^ 

V 

-0 

^ 

102  HIGHE.l-i  MATHEMATICS.  §38. 

The  £-axis  is  called  the  transverse  or  real  axes  of  the  hyperbola  ; 
the  ?/-axis  the  conjugate  or  imaginary  axes ;.  the  points  A,  A'  are 
the  vertices  of  the  hyperbolas,  a  is  the  real  semi-axis,  b  the  imaginary 
semi-axis. 

Examples. — (1)  Show  that  the  equation  of  the  hyperbola  whose  origin 
is  at  its  vertex  is  ahj2  =  2ab2x  +  b2x2.  Substitute  x  +  a  for  x  in  the  regular 
equation.     Note  that  y  does  not  change. 

(2)  Investigate  the  shape  of  the  hyperbola.  By  solving  equation  (1)  for 
xt  and  y,  we  get 

y  =  ±  -  \/a2  -  a*,  and  x  =  ±  £  Jy^+W.      ...         (2) 

First.  Since  y2  must  be  positive,  x2  <£  a2,  or  x  cannot  be  numerically  less 
than  a.     No  limit  with  respect  to  y  can  be  inferred  from  equation  (8). 

Second.  For  every  positive  value  of  x,  there  are  two  values  of  y  differing 
only  in  sign.  Hence  these  two  points  are  perpendicular  above  and  below  the 
x-axis,  that  is  to  say,  the  hyperbola  is  symmetrical  with  respect  to  the  x-axis. 
There  are  two  equal  and  opposite  values  of  x  for  all  values  of  y.  The  hyper- 
bola is  thus  symmetrical  with  respect  to  the  y-a,xis. 

Third.  If  the  value  of  x  changes  from  zero  until  x  =  ±  a,  then  y  =  0,  and 
these  two  values  of  x  furnish  two  points  on  the  a:-axis.  If  x  ->  a,  there  are 
two  equal  and  opposite  values  of  y.  Similarly  for  every  value  of  y  there  are 
two  equal  and  opposite  values  of  x.  The  curve  is  thus  symmetrical  with 
respect  to  both  axes,  and  lies  beyond  the  limits  x  =  ±  a. 

Before  describing  the  properties  of  this  interesting  curve  I 
shall  discuss  some   fundamental  properties    of  curves  in  general. 

§  38.  The  Tangent  to  a  Curve. 

We  sometimes  define  a  tangent  to  a  curve  as  a  straight  line 
which  touches  the  curve  at  two  co- 
incident points.1  If,  in  Fig.  24,  P 
and  Q  are  two  points  on  a  curve 
such  that  MP  =  NB  =  y ;  BQ  =  dy  ; 
OM  =x;  MN  =  PB  =  dx;  the  straight 
line  PQ  =  ds.  Otherwise,  the  diagram 
explains  itself.  Now  let  the  line  APQ 
FlGl  24,  revolve  about  the  point  P.     We  have 

already  shown,  on  page  15,  that  the  chord  PQ  becomes  more  and 
more  nearly  equal  to  the  arc  PQ  as  Q  approaches  P ;   when  Q 

1  Note  the  equivocal  use  of  the  word  "  tangent"  in  geometry  and  in  trigonometry. 
In  geometry,  a  "  tangent  is  a  line  between  which  and  the  curve  no  other  straight  line 
can  be  drawn,"  or  "a  line  which  just  touches  but  does  not  cut  the  curve  ".  The 
slope  of  a  curve  at  any  point  can  be  represented  by  a  tangent  to  the  curve  at  that 
point,  and  this  tangent  makes  an  angle  of  tan  o  with  the  cc-axis. 


§  38.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  103 

coincides  with  P,  the  angle  MTP  =  angle  BPQ  =  a ;  dx,  dy  and 
ds  are  the  sides  of  an  infinitesimally  small  triangle  with  an  angle 
at  P  equal  to  a ;  consequently 

|  -  tan  a.         ....         (1) 

This  is  a  most  important  result.  The  differential  coefficient  repre- 
sents the  slope  of  gradient  of  the  curve.  In  other  words,  the  tan- 
gent of  the  angle  made  by  the  slope  of  any  part  of  a  curve  with 
the  a;-axis  is  the  first  differential  coefficient  of  the  ordinate  of  the 
curve  with  respect  to  the  abscissa. 

We  can  also  see  very  readily  that  in  the  infinitely  small  triangle, 
dx  =  ds  .  cos  a ;  dy  =  ds  .  sin  a  ;        .         .         (2) 
and,  since  B  is  a  right  angle, 

(dsf  =  {dyf  +  {dxf.        ...        (3) 

If  we  plot  the  distances,  x,  traversed  by  a  particle  at  different 
intervals  of  time  (abscissae) ;  or  the  amounts  of  substance,  x,  trans- 
formed in  a  chemical  reaction  at  different  intervals  of  time,  t,  we 
get  a  curve  whose  slope  at  any  point  represents  the  velocity  of  the 
process  at  the  corresponding  interval  of  time.  This  we  call  a 
velocity  curve.  If  the  curve  slopes  downwards  from  left  to  right, 
dx/dt  will  be  negative  and  the  velocity  of  the  process  will  be 
diminishing ;  if  the  curve  slopes  upwards  from  left  to  right,  dy/dx 
will  be  positive,  and  the  velocity  will  be  increasing. 

If  we  plot  the  velocity,  V,  of  any  process  at  different  intervals  of 
time,  t,  we  get  a  curve  whose  slope  indicates  the  rate  at  which  the 
velocity  is  changing.  This  we  call  an  acceleration  cur  Ye.  The 
area  bounded  by  an  acceleration  curve  and  the  coordinate  axes 
represents  the  distance  traversed  or  the  amount  of  substance  trans- 
formed in  a  chemical  reaction  as  the  case  might  be. 

Examples. — (1)  At  what  point  in  the  curve  y^  =  Axx  does  the  tangent 
make  an  angle  of  60°  with  the  sc-axis  ?  Here  dy1/dx1  =  2\yx  =  tan  60°  m  >/s. 
Ansr.  yx  =2/>/|";  «,*=*. 

(2)  Find  the  tangent  of  the  angle,  a,  made  by  any  point  P(x,  y)  on  the 
parabolic  curve.  In  other  words,  it  is  required  to  find  a  straight  line  which 
has  the  same  slope  as  the  curve  has  which  passes  through  the  point  P(x,  y). 
Since  y2  =  ±ax  ;  dyfdx  =  2a\y  =  tan  o.  If  the  tangent  of  the  angle  were  to 
have  any  particular  value,  this  value  would  have  to  be  substituted  in  place  of 
dy/dx.  For  instance,  let  the  tangent  at  the  point  P(x,  y)  make  an  angle  of 
45°.  Since  tan  45  =  unity,  2a/y  =  tan  a  =  1,  .♦.  y  =  2a,  Substituting  in  the 
original  equation  y2  =  iax,  we  get  x  =  a,  that  is  to  say,  the  required  tangent 


104  HIGHER  MATHEMATICS.  §  38. 

passes  through  the  extremity  of  the  ordinate  perpendicular  on  the  focus. 
If  the  tangent  had  to  be  parallel  to  the  x-axis,  tan  0  being  zero,  dyjdx  is 
equated  to  zero ;  while  if  the  tangent  had  to  be  perpendicular  to  the  x-axis, 
since  tan  90°  =  oo,  dyjdx  =  oo. 

(3)  Required  the  direction  of  motion  at  any  moment  of  a  point  moving 
according  to  the  equation,  y  =  a  cos  2ir(x  +  e).  The  tangent,  at  any  time  t, 
has  the  slope,  -  %ra  sin  2ir(x  f  e). 

(4)  E.  Mallard  and  H.  le  Ghatelier  represent  the  relation  between  the 
molecular  specific  heat,  s,  of  carbon  dioxide  and  temperature,  0,  by  the 
expression  s  =  6'3  +  0*005640  -  0*000001, O802.  Plot  the  (9,ds/de) -curve  from 
0  =  0°  to  6  =  2,000  (abscissae).  Possibly  a  few  trials  will  have  to  be  made 
before  the  "  scale  "  of  each  coordinate  will  be  properly  proportioned  to  give 
the  most  satisfactory  graph.  The  student  must  learn  to  do  this  sort  of  thing 
for  himself.  What  is  the  difference  in  meaning  between  this  curve  and  the 
(s,6)  -curve? 

(5)  Show  that  dxjdy  is  the  cotangent  of  the  angle  whose  tangent  is  dyjdx. 

Let  TP  (Fig.  25)  be  a  tangent  to  the  curve  at  the  point 
P(xv  y^.  Let  OM  =  xv  MP  =  yv  Let  y  =  mx  +  b,  be  the  equation 
of  the  tangent  line  TPT',  and  yY  =  /(#i)  the  equation  of  the  curve, 
BOP.  From  (14),  page  94,  we  know  that  a  straight  line  can  only 
pass  through  the  point  P(xv  yj,  when 

y  -y1  =  m(x  -xY)  .         .         .         (4) 

where  m  is  the  tangent  of  the  angle  which  the  line  y  =  mx  makes 
with  the  ic-axis  ;  and  x  and  y  are  the  coordinates  of  any  point  taken 
at  random  on  the  tangent  line.  But  we  have  just  seen  that  this 
angle  is  equal  to  the  first  differential  coefficient  of  the*  ordinate  of 
the  curve  ;  hence  by  substitution 

9-*-i[fe-**>'  ■     ■     •     (5) 

which  is  the  required  equation  of  the  tangent  to  a  curve  at  a 
point  whose  coordinates  are  xv  yv 

Examples. — (1)  Find  the  equation  of  the  tangent  at  the  point  (4,  2)  in 
the  curve  yx2  =  4^.  Here,  dy1/dx1  =  1 ;  xl  =  4,  yx  =  2.  Hence,  from  (4), 
y  =  x  -  2  is  the  required  equation. 

(2)  Required  the  equation  of  the  tangent  to  a  parabola.     Since 
yt*  =  ±axv  dy1jdx1  =  2ajyx. 
Substituting  in  (5)  and  rearranging  terms, 

(V  ~  VdVi  -  VVi  ~  2/i2  =2a(*  ~  «i)- 
Substituting  for  y^,  we  get 

yyx  =  2a{x  +  a^) (6) 

as  the  equation  for  the  tangent  line  of  a  parabola.  If  x  =  0,  tan  a  =  oo,  and 
the  tangent  is  perpendicular  to  the  «-axis  and  touches  the  y-axis.     To  get  the 


8  38.        COOKDINATE  OR  ANALYTICAL  GEOMETRY.  105 

point  of  intersection  of  the  tangent  with  the  sc-axis  put  y  =  0,  then  x  =  -  xv 
The  vertex  of  the  parabola  therefore  bisects  the  aj-axis  between  the  point  of 
intersection  of  the  tangent  and  of  the  ordinate  of  the  point  of  tangency. 

(3)  Find  the  equation  of  the  tangent  to  the  ellipse, 

a?  I"  b2  -■*•     •>  dXi  -      a2yi> 

substituting  this  value  of  dy1jdxl  in  (5),  multiply  the  result  by  yx\  divide 
through  by  b2 ;  rearrange  terms  and  combine  the  result  with  the  equation  of 
the  ellipse,  (1),  page  100.     The  result  is  the  tangent  to  any  point  on  the  ellipse, 

■a+m..i (7) 

a2  +  62       *•  \n 

where  xlt  yx  are  coordinates  of  any  point  on  the  curve  and  x,  y  the  coordi- 
nates of  the  tangent. 

(4)  Find  the  equation  of  the  tangent  at  any  point  P{xv  yx)  on  the  hyper- 
bolic ourve.     Differentiate  the  equation  of  the  hyperbola 

d>      62      L-    '  •  dx'a*  Vl'  "  V    Vx~  a?  Vl(X      Xl)' 

Multiply  this  equation  by  yx ;  divide  by  62 ;  rearrange  the  terms  and  combine 
the  result  with  the  second  of  the  above  equations.  We  thus  find  that  the 
tangent  to  any  point  on  the  hyperbola  has  the  equation 

HF    "fe^"1 (8) 

At  the  point  of  intersection  of  the  tangent  to  the  hyperbola  with  the  cc-axis, 
y  =  0  and  the  corresponding  value  of  x  is 

xxx  =  a2 ;  or,  x  =  a2/^,  ....        (9) 

the  same  as  for  the  ellipse. 

From  (9)  if  xx  is  infinitely  great,  x  =  0,  and  the  tangent  then 
passes  through  the  origin.  The  limiting  position  of  the  tangent 
to  the  point  on  the  hyperbola  at  an  infinite  distance  away  is 
interesting.  Such  a  tangent  is  called  an  asymptote.  To  find 
the  angle  which  the  asymptote  makes  with  the  ic-axis  we  must 
determine  the  limiting  value  of 

b2      a2         ' 
when  xx  is  made  infinitely  great.     Multiply  both  sides  by  ft/x^2, 
and 

•  g£    £  .'■* 

'''x2~a2     X* 

If  x1  be  made  infinitely  great  the  desired  ratio  is 

Lt        Vi2-b2     •    Lt         y*-b 
1      x,2     a2  x,     a 


106 


HIGHER  MATHEMATICS. 


§39. 


Differentiate  the  equation  of  the  hyperbola,    and  introduce  this 
value  for  xjyv  and  we  get 


dy  ,     t       a     b*     b 

-f-  =  tan  a   say    =  T  .  -^=  - 

dx  b     a2    a 


(10) 


If  we  now  construct  the  rectangle  BSS'B'  (Fig.  23,  page  101) 
with  sides  parallel  to  the  axes  and  cut  off  OA  =  OA'  =  a,  OB  =  OB  =  b, 
the  diagonal  in  the  first  quadrant  and  the  asymptote,  having  the 
same  relation  to  the  two  axes,  are  identical.  Since  the  x-  and 
2/-axes  are  symmetrical,  it  follows  that  these  conditions  hold  for 
every  quadrant.  Hence,  B'OS,  and  BOS'  are  the  asymptotes  of 
the  hyperbola. 


§  39.    A  Study  of  Curves. 

A  normal  line  is  a  perpendicular  to  the  tangent  at  a  given 

point  on  the  curve,  drawn  to  the 
a;-axis.  Let  NP  be  normal  to  the 
curve  (Fig.  25)  at  the  point  P 
ixv  V\l'  ^et  V  —  mx  +  ^>  ke  the 
equation  of  the  normal  line ;  yx  — 
/(o^),  the  equation  of  the  curve. 
The  condition  that  any  line  may 
be  perpendicular  to  the  tangent 
line  TP,  is  that  m' «  -  1/m,  (17), 
page  96.     From  (5) 


??r 


or,  the  equation  of  the  normal  line  is 

dx,,  . 

y  -  v\  -  -  tut(x  -  xi)  »  or>  - 


dyx 


dxY 


y  -  vi 

X   —   X-,' 


(i) 


Examples. — (1)  Find  the  equation  of  the  normal  at  the  point  (4,  3)  in  the 
curve  xfyj*  —  a.  Here  dxljdyl  =  -  Sx1/2yv  Hence,  by  substitution  in  (1), 
y  =  28  -  5. 

(2)  Show  that  y  =  2(x  -  6)  is  the  equation  of  the  normal  to  the  curve 
tyi  +  x\  =  0.  at  the  point  (4,  -  4). 

(3)  The  tangent  to  the  ellipse  cuts  the  sc-axis  at  a  point  where  y  =  0 ; 
from  (7),  page  105, 

.\xxl=a?;  or,  x  =  a2/^ (2) 

In  Fig.  26  let  PT  be  a  tangent  to  the  ellipse,  NP  the  normal.     From  (2), 
FXT  =  x  +  c  =  tf\xx  +  c ;  FT  =x  -  c  =  a?\xx  -  c, 


§39. 


COORDINATE  OR  ANALYTICAL  GEOMETRY. 


107 


since  FxO  =  OF  =  c;  OT  =  x;  OM  =  xv 

FXT  _  a2  +  ex, 
'  '•  FT  ~  a2  -  cx1     '     ■    ' 
Since  FP  =  r,  FXP  m 'r,;  OF  =  FxO  =  c;  OM  =  xx ;  MP  =  yx, 

H  =  2/12  +  (e  -  *i)2 ;  n2  =  2/i2  +  (c  +  a?!)2. 
.-.  r2  -  rj2  =  (r  +  r2)  (r  -  r^  =  -  40^ ; 

But  by  the  definition  of  an  ellipse,  pages  100  and  101, 

r  +  rx  =  2a;  .*.  r  -  rx  =  -  2cxx/a;  .\  r  =  a  -  exja ;  rx  =  a  +  exja. 


(3) 


FXP 
FP 


a1  +  ex, 


(4) 


From  (3)  and  (4),  therefore, 


FXT :  FT  =  FXP  :  FP. 


Fig.  26.— The  Foci  of  the  Ellipse. 

By  Euclid,  vi.,  A:  "If,  in  any  triangle,  the  segments  of  the  base  produced 
have  to  one  another  the  same  ratio  as  the  remaining  sides  of  the  triangle, 
the  straight  line  drawn  from  the  vertex  to  the  point  of  section  bisects  the 
external  angle".  Hence  in  the  triangle  FPFX,  the  tangent  bisects  the  ex- 
ternal angle  FPB,  and  the  normal  bisects  the  angle  FPFX. 

The  preceding  example  shows  that  the  normal  at  any  point  on 
the  ellipse  bisects  the  angle  enclosed  by  the  focal  radii ;  and  the 
tangent  at  any  point  on  the  ellipse  bisects  the  exterior  angle  formed 
by  the  focal  radii.  This  property  accounts  for  the  fact  that  if  FYP 
be  a  ray  of  light  emitted  by  some  source  Fv  the  tangent  at  P 
represents  the  reflecting  surface  at  that  point,  and  the  normal  to  the 
tangent  is  therefore  normal  to  the  surface  of  incidence.  From  a 
well-known  optical  law,  "the  angles  of  incidence  and  reflection 
are  equal,"  and  since  FXPN  is  equal  to  NPF  when  PF  is  the  re- 
flected ray,  all  rays  emitted  from  one  focus  of  the  ellipse  are 
reflected  and  concentrated  at  the  other  focus.  This  phenomenon 
occurs  with  light,  heat,  sound  and  electro-magnetic  waves. 


108  HIGHER  MATHEMATICS.  §  39. 

To  find  the  length  of  the  tangent  and  of  the  normal.  The  length 
of  the  tangent  can  be  readily  found  by  substituting  the  values 
MP  and  TM  in  the  equation  for  the  hypotenuse  of  a  right-angled 
triangle  TPM  (Euclid,  i.,  47);  and  in  the  same  way  the  length 
of  the  normal  is  obtained  from  the  known  values  of  MN  and  PM 
already  deduced. 

The  subnormal  of  any  curve  is  that  part  of  the  #-axis  lying 
between  the  point  of  intersection  of  the  normal  and  the  ordinate 
drawn  from  the  same  point  on  the  curve.  Let  MN  be  the  sub- 
normal of  the  curve  shown  in  Fig.  25,  then 

MN  =  x  -  xv 
and  the  length  of  MN  is,  from  (1), 

l3a;1',  "*'  dxx  ™      y1 

when  the  normal  is  drawn  from  the  point  P(xv  y^). 

The  subtangent  of  any  curve  rs  that  part  of  the  #-axis  lying 
between  the  points  of  intersection  of  the  tangent  and  the  ordinate 
drawn  from  the  given  point.  Let  TM  (Fig.  25)  be  the  subtangent, 
then 

x1  -  x  =  TM. 

Putting  y  =  0  in  equation  (1),  the  corresponding  value  for  the 
length,  TM,  of  the  subtangent  is 

dx-.  dx,      X-,  —  x  ... 

*-■'**%' *%-*•  \    •     (6) 

Examples. — (1)  Find  the  length  of  the  subtangent  and  subnormal  lines 
in  the  parabola,  y-f  =  4Laxv  Since  yidyjdx-y  =  2a,  the  subtangent  is  2x1 ;  the 
subnormal,  2a.     Hence  the  vertex  of  the  parabola  bisects  the  subtangent. 

(2)  Show  that  the  subtangent  of  the  curve  pv  =  constant,  is  equal  to  -  v. 

(3)  Let  P(x,  y)  be  a  point  on  the  parabolic  curve  (Fig.  27)  referred  to  the 
coordinate  axes  Ox,  Oy;  PT  a  tangent  at  the  point  P,  and  let  KA  be  the 
directrix.  Let  F  be  the  focus  of  the  parabola  y2  =  4ax.  Join  PF.  Draw  KP 
parallel  to  Ox.  Join  KT.  Then  KPFT  is  a  rhombus  (Euclid,  i.,  34) ,  for  it 
has  been  shown  that  the  vertex  of  the  parabola  O  bisects  the  subtangent,  Ex. 
(1)  above.    Hence,  TO  =  OM;  and,  by  definition,  AO  =  OF ' ; 

.;•.  TA  =  FM;  and  KP  =  TF\ 
consequently,  the  sides  KT  and  PF  are  parallel,  and  by  definition  of  the  para- 
bola, KP  =  PFt  .'.  the  two  triangles  KPT  and  PTF  are  equal  in  all  respects, 
and  (Euclid,  i.,  5)  the  angle  KPT  =  angle  TPF,  that  is  to  say,  the  tangent 
to  the  parabola  at  any  given  point  bisects  the  angle  made  by  the  focal  radius 
and  the  perpendicular  dropped  on  to  the  directrix  from  the  given  point. 

In  Fig.  27,  the  angle  TPF  =  angle  TPK  =  opposite  angle  RPT  (Euclid, 


§  40.        COOEDINATE  OR  ANALYTICAL  GEOMETRY. 


109 


i.,  15).     But,  by  construction,  the  angles  TPN  and  NPT'  are  right  angles 
take  away  the  equal  angles  TPF  and  BPT 
and  the  angle  FPN  is  equal  to  the  angle 
NPR. 


The  normal  at  any  point  on  the 
parabola  bisects  the  angle  enclosed  by 
the  focal  radius  and  a  line  drawn 
through  the  given  point,  parallel  to  the 
x-axis.  This  property  is  of  great  im- 
portance in  physics.  All  light  rays 
falling  parallel  to  the  principal  (or  x-) 
axis  on  to  a  parabolic  mirror  are  reflect- 
ed at  the  focus  F,  and  conversely  all 
light  rays  proceeding  from  the  focus 
are  reflected  parallel  to  the  #-axis. 


Fig.  27.— TheFocus  of  the 
Parabola. 


Hence  the  employment  of 
parabolic  mirrors  for  illumination  and  other  purposes.  In  some 
of  Marconi's  recent  experiments  on  wireless  telegraphy,  electrical 
radiations  were  directed  by  means  of  parabolic  reflectors.  Hertz, 
in  his  classical  researches  on  the  identity  of  light  and  electro- 
magnetic waves,  employed  large  parabolic  mirrors,  in  the  focus  of 
which  a  "generator,"  or  "receiver"  of  the  electrical  oscillations 
was  placed.  See  D.  E.  Jones'  translation  of  H.  Hertz's  Electric 
Waves,  London,  172,  1893. 


§  $0.    The  Rectangular  or  Equilateral  Hyperbola. 

If  we  put  a  =  6  in  the  standard  equation  to  the  hyperbola,  the 
result  is  an  hyperbola  (Fig.  28)  for  which 

x>  -  y*  =  a2,  .  (1) 
and  since  tan  a  =  1  =  tan 
45°,  each  asymptote  makes 
an  angle  of  45°  with  the  x- 
or  2/-axes.  In  other  words, 
the  asymptotes  bisect  the 
coordinate  axes.  This  special 
form  of  the  hyperbola  is  called 
an  equilateral  or  rectangular 
hyperbola.  It  follows  directly 
that  the  asymptotes  are  a't 
right  angles  to  each  other. 
The  asymptotes  may,   therefore,  serve  as   a  pair  of  rectangular 


Fig.  28. — The  Kectangular  Hyperbola. 


110  HIGHER  MATHEMATICS.  §  41. 

coordinate  axes.     This  is  a  valuable  property  of  the  rectangular 
hyperbola. 

The  equation  of  a  rectangular  hyperbola  referred  to  its  asymp- 
totes as  coordinate  axes,  is  best  obtained  by  passing  from  one  set  of 
coordinates  to  another  inclined  at  an  angle  of  -  45°  to  the  old  set, 
but  having  the  same  origin,  as  indicated  on  page  96.  In  this  way 
it  is  found  that  the  equation  of  the  rectangular  hyperbola  is 

xy  =  a}  .         .         .         .         (2) 

where  a  is  a  constant. 

It  is  easy  to  see  that  as  y  becomes  smaller,  x  increases  in  magni- 
tude. When  y  •=  0,  x  =  oo,  the  rr-axis  touches  the  hyperbola  an 
infinite  distance  away.     A  similar  thing  might  be  said  of  the  y -axis. 

§  $1.  Illustrations  of  Hyperbolic  Curves. 

I.  The  graphical  representation  of  the  gas  equation,  pv  =  BO, 
furnishes  a  rectangular  hyperbola  when  6  is  fixed  or  constant. 
The  law  as  set  forth  in  the  above  equation  shows  that  the  volume 
of  a  gas,  v,  varies  inversely  as  the  pressure,  p,  and  directly  as  the 
temperature,  0.  For  any  assigned  value  of  0,  we  can  obtain  a 
series  of  values  of  p  and  v.  For  the  sake  of  simplicity,  let  the 
constant  B  ■-  1.     Then  if 

0=1 

»*■*..        M         ^         M         M  M         etc. 

The  "curves"  of  constant  temperature  obtained  by  plotting 
these  numbers  are  called  isothermals.  Each  isothermal  (i.e., 
curve  at  constant  temperature)  is  a  rectangular  hyperbola  obtained 
from  the  equation  pv  =  BO  =  constant,  similar  to  (2),  above. 

A  series  of  isothermal  curves,  obtained  by  putting  0  successively 
equal  to  0V  02,  0S  .  . .  and  plotting  the  corresponding  values  of 
p  and  v,  is  shown  in  Fig.  29. 

We  could  have  obtained  a  series  of  curves  from  the  variables 
p  and  0,  or  v  and  0,  according  as  we  assume  v  or  p  to  be  constant. 
If  v  be  constant,  the  resulting  curves  are  called  isometric  lines, 
or  isochores ;  if  p  be  constant  the  curves  are  isopiestic  lines, 
or  isobars. 

II.  Exposure  formula  for  a  thermometer  stem.  When  a  ther- 
mometer stem  is  not  exposed  to  the  same  temperature  as  the 


p=    0-1, 

0-5, 

10, 

5-0, 

10-0, 

v  =  10-0, 

2-0, 

i-o, 

0-2, 

o-i, 

p-    0-1, 

0-5, 

1-0, 

5-0, 

10-0, 

v  =    5-0, 

i-o, 

0-5, 

o-i, 

0-05; 

§  41.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  Ill 

bulb,  the  mercury  in  the  exposed  stem  is  cooled,  and  a  small 
correction  must  be  made  for  the  consequent  contraction  of  the 
mercury  exposed  in  the  stem.  If  x  denotes  the  difference  between 
the  temperature  registered  by  the  thermometer  and  the  tempera- 

p 


Fig.  29. — Isothermal  ^w-curves. 

ture  of  the  exposed  stem,  y  the  number  of  thermometer  divisions 
exposed  to  the  cooler  atmosphere,  then  the  correction  can  be 
obtained  by  the  so-called  exposure  formula  of  a  thermometer, 
namely, 

0  =  0-00016^, 

which  has  the  same  form  as  equation  (2),  page  110.  By  assuming 
a  series  of  suitable  values  for  0  (say  0*1 ... )  and  plotting  the 
results  for  pairs  of  values  of  x  and  y,  curves  are  obtained  for  use 
in  the  laboratory.  These  curves  allow  the  required  correction  to 
be  seen  at  a  glance. 

III.  Dissociation  curves.  Gaseous  molecules  under  certain 
conditions  dissociate  into  similar  parts.  Nitrogen  peroxide,  for 
instance,  dissociates  into  simpler  molecules,  thus : 

N204-2N02. 
Iodine  at  a  high  temperature  does  the  same  thing,  I2  becoming  21. 
In  solution  a  similar  series  of  phenomena  occur,  KC1  becoming 
K  +  01.  and  so  on.     Let  x  denote  the  number  of  molecules  of  an 


112 


HIGHER  MATHEMATICS. 


§41. 


acid  or  salt  which  dissociate  into  two  parts  called  ions  ;  (1  -  x) 
the  number  of  molecules  of  the  acid,  or  salt  resisting  ionization  ; 
c  the  quantity  of  substance  contained  in  unit  volume,  that  is  the 


75 

* 

50 

25 

?r\ 

5  IO  15  2C 

Fig.  30. — Dissociation  Isotherm. 


ss 


30 


concentration  of  the  solution.     Nernst  has  shown  that  at  constant 
temperature 

Km 


CXl 


where  E  is  the  so-called  dissociation  constant  whose  meaning  is 
obtained  by  putting  x  =  0'5.  In  this  case  K  =  Jc,  that  is  to  say, 
K  is  equal  to  half  the  quantity  of  acid  or  salt  in  solution  when 
half  of  the  acid  or  salt  is  dissociated. 

Putting  Z=lwe  can  obtain  a  series  of  corresponding  values 
of  c  and  x.     For  example,  if 

x  =  -16,        0-25,        0-5,        0-75,        0'94     . . . ; 
then       g=   32,  12,  2,        0-44,        0-07     ... 

It  thus  appears  that  when  the  concentration  is  very  great,  the 
amount  of  dissociation  is  very  small,  and  vice  versd,  when  the  con- 
centration is  small  the  amount  of  dissociation  is  very  great.  Com- 
plete dissociation  can  perhaps  never  be  obtained.  The  graphic 
curve  (Fig.  30),  called,  by  Nernst,  the  dissociation  isotherm,  is 
asymptotic  towards  the  two  axes,  but  when  drawn  on  a  small  scale 
the  curve  appears  to  cut  the  ordinate  axis. 

IV.  The  volume  elasticity  of  a  substance  is  defined  as  the  ratio 
of  any  small  increase  of  pressure  to  the  diminution  of  volume  per 
unit  volume  of  substance.      If  the  temperature  is  kept  constant 


§41. 


COORDINATE  OR  ANALYTICAL  GEOMETRY. 


113 


during  the  change,  we  have  isothermal  elasticity,  while  if  the 
change  takes  place  without  gain  or  loss  of  heat,  adiabatic  elas- 
ticity. If  unit  volume  of 
gas,  v,  changes  by  an  P 
amount  dv  for  an  increase 
of  pressure  dp,  the  elastic- 
ity, E,  is 


E  =  - 


=  -  v 


dp 
dv 


(i) 


A  similar  equation  can  be 
obtained  by  differentiating 
Boyle's  law, pv  =  constant, 
for  an  isothermal  change 
of  state.  The  result  is 
that 

dp 


V 


dv 


(2) 


M  T 

Fig.  31.— pv-curves. 


an  equation  identical  with  that  deduced  for  the  definition  of  volume 
elasticity.  The  equation  pv  =  constant  is  that  of  a  rectangular 
hyperbola  referred  to  its  asymptotes  as  axes. 

Let  P(p,  v)  (Fig.  31)  be  a  point  on  the  curve  pv  =  constant. 
In  constructing  the  diagram  the  triangles  ENP  and  PMT  were 
made  equal  and  similar  (Euclid,  i.,  26).  See  Ex.  (2)  page  108,  and 
note  that  EN  is  the  vertical  subtangent  equivalent  to  -.  p. 

EN  =  -  NP  tan  a  =  -  v  tan  EPN  =  -  v-£, 

that  is  to  say,  the  isothermal  elasticity  of  a  gas  in  any  assigned 
condition,  is  numerically  equal  to  the  vertical  subtangent  of  the 
curve  corresponding  to  the  substance  in  the  given  state. 

But  since  in  the  rectangular  hyperbola  EN  =  PM,  the  iso- 
thermal elasticity  of  a  gas  is  equal  to  the  pressure  (2).  The 
adiabatic  elasticity  of  a  gas  may  be  obtained  by  a  similar  method 
to  that  used  for  equation  (1).  If  the  gas  be  subject  to  an  adia- 
batic change  of  pressure  and  volume  it  is  known  that 

pVy  =  constant  =  G.  .  ...  (3) 

Taking  logarithms,  (3)  furnishes  log  #>  +  y  log  v  =  log  C.  By 
differentiation  and  rearrangement  of  terms,  we  get 


114  HIGHER  MATHEMATICS.  §  42. 

in  other  words  the  adiabatic  elasticity1  of  a  gas  is  y  times  the 
pressure.     A  similar  construction  for  the  adiabatic  curve  furnishes 

EN :  PM  =  KP  :  PT  =  y  :  1, 
that  is  to  say,  the  tangent  to  an  adiabatic  curve  is  divided  at  the 
point  of  contact  in  the  ratio  y  :  1. 

Examples.— (1)  Assuming  the  Newton-Laplace  formula  that  the  square 
of  the  velocity  of  propagation,  V,  of  a  compression  wave  (e.g.,  of  sound)  in  a 
gas  varies  directly  as  the  adiabatic  elasticity  of  the  gas,  E^,  and  inversely  as 
the  density,  p,  or  T2  oc  E^jp ;  show  that  72  oc  yRT.  Hints  :  Since  the  com- 
pression wave  travels  so  rapidly,  the  changes  of  pressure  and  volume  may  be 
supposed  to  take  place  without  gain  or  loss  of  heat.  Therefore,  instead  of 
using  Boyle's  law,  pv  =  constant,  we  must  employ  pvi  =  constant.  Hence 
deduce  yp  =  v .  dp/dv  =  E$.  Note  that  the  volume  varies  inversely  as  the 
density  of  the  gas.     Hence,  if 

T2  oc  E^/p  oc  E$v  oc  ypv  oc  yRT.   ....       (5) 

(2)  R.  Mayer's  equation,  page  82  and  (5)  can  be  employed  to  determine 
the  two  specific  heats  of  any  gas  in  which  the  velocity  of  sound  is  known. 
Let  a  be  a  constant  to  be  evaluated  from  the  known  values  of  R,  T,  "P, 

.-.  Cv  =  i2/(l  -  a),  and  Cp  -  aCv (6) 

Boynton  has  employed  van  der  Waals'  equation  in  place  of  Boyle's.  Per- 
haps the  reader  can  do  this  for  himself.  It  will  simplify  matters  to  neglect 
terms  containing  magnitudes  of  a  high  order  (see  W.  P.  Boynton,  Physical 
Review,  12,  353, 1901). 

§  $2.   Polar  Coordinates. 

Instead  of  representing  the  position  of  a  point  in  a  plane  in 
terms  of  its  horizontal  and  vertical  distances  along  two  standard 
lines  of  reference,  it  is  sometimes  more  convenient  to  define  the 
position  of  the  point  by  a  length  and  a  direction.  For  example,  in 
Fig.  32  let  the  point  0  be  fixed,  and  Ox  a  straight  line  through  0. 
Then,  the  position  of  any  other  point  P  will 
be  completely  defined  if  (1)  the  length  OP 
and  (2)  the  angle  OP  makes  with  0x}  are 
known.  These  are  called  the  polar  coordin- 
ates of  P,  the  first  is  called  the  radius 
vector,  the  latter  the  vectorial  angle.  The 
Fig.  32.— Polar  Co-  radius  vector  is  generally  represented  by  the 
ordinates.  symbol  r,  the  vectorial  angle  by  0,  and  P  is 

called  the  point  P(r,  0),  0  is  called  the  pole  and  Ox  the  initial  line. 
As  in  trigonometry,  the  vectorial  angle  is  measured  by  supposing 

i  From  other  considerations,  Eq  is  usually  written  E$. 


§  42.        COORDINATE  OK.  ANALYTICAL  GEOMETRY.  115 


the  angle  0  has  been  swept  out  by  a  revolving  line  moving  from 
a  position  coincident  with  Ox  to  OP.  It  is  positive  if  the  direction 
of  revolution  is  contra  wise  to  the  motion  of  the  hands  of  a  clock. 

To  ohange  from  polar  to  rectangular  coordinates  and  vice  versd. 
In  Fig.  33,  let  (r,  0)  be  the  polar  coordinates  of  the  point  P(x,  y). 
Let  the  angle  x'OP  -  0. 


I.  To  pass  from  Cartesian  to  polar  coordinates. 

.  MP      y  OM      x, 

sm  6  =  ^p  =  -  ;  cos  6  =  -gp  =  -  > 


.-.  y  =  rsinfl  and  x  =  rcos0, . 
which  expresses  x,  and  y,  in  terms  of  r  and  0. 


(i) 


Examples. — (1)  Transform  the  equation  x2  -  y2  =  3  from  rectangular  to 
polar  coordinates,  pole  at  origin.  Ansr.  r2cos  20  =  3.  Hint.  Cos20  -  sin20  = 
cos  29. 


K 

y 

/    le\ 

0 

a     f      r 

I 

Fig.  33. 


Fig.  34. 


3.     Hint. 


(2)  Show  that  x2  +  y2  =  9  represents  the  same   line  as 
r2(sin20  +  cos20)  =  9 ;  and  sin20  +  cos20  =  1. 

(3)  A  point  P  moves  along  a  curve  in  such  a  way  that  the  ratio  of  its 
distance  from  a  given  point  F,  and  from  a  given  straight  line  OK  (Fig.  34)  is 
a  constant  quantity,  say  e.  Find  the  path  of  the  point.  Hint.  In  Fig.  34 
FP  =  eKP.  Let  KP  =  OM  =  x ;  MP  =  y.  Required  the  equation  connect- 
ing a;  and  y.  (FP)2  =  (MP)2  +  (FMf  =  y2  +  (x  -  a)2,  where  OF  is  put  =  a. 
If  e  is  unity,  the  curve  is  a  parabola ;  if  e  <  1  the  curve  is  an  ellipse ;  if 
e  >  1  the  curve  is  an  hyperbola,  e  is  called  the  eccentricity  of  the  curve. 
In  polar  coordinates  KP  =  OF  +  FM  =  a  +  r  cos  0, 

t  ae 

•*'  e  =  a  +  rcoae'  °r  *  =    1  -ecos0'  ® 

whether  curve  be  an  hyperbola,  ellipse  or  parabola. 

II.  To  pass  from  polar  to  Cartesian  coordinates.  In  the  same 
figure 


116 


H1GHEK  MATHEMATICS. 


§43. 


tantf  = 


MP      y, 

0M~  x' 
r2  =  (OPf  =  (OM)2  +  (MPf  . 

.-.  0  =  tan"1^;  r  =  ±   V 


;2   +  ^2 


a;^  +  yl 


(3) 


which  expresses  0,  and  r,  in  terfhs  of  x  and  ?/.  The  sign  of  r  is 
ambiguous,  but,  by  taking  any  particular  solution  for  6,  the  pre- 
ceding remarks  will  show  which  sign  is  to  be  taken. 

Just  as  in  Cartesian  coordinates,  the  graph  of  a  polar  equation 
may  be  obtained  by  assigning  convenient  values  to  0  (say  0°,  30°, 
45°,  60°,  90°  . . .)  and  calculating  the  corresponding  value  of  r  from 
the  equation. 

Examples. — (1)  What  are  the  rectangular  coordinates  of  the  points  (2, 60°), 
and  (2,  45°)  respectively?    Ansr.  (1,  \/3),  and  {J2,  \/2). 

(2)  Express  the  equation  r  =  m  cos  0  in  rectangular  coordinates.  Ansr. 
x2  +  y2  =  mx.     Hint.  Cos  6  =  x\r ;  .*.  r2  =  mx,  eto. 

Polar  coordinates  are  particularly  useful  in  astronomical  and 
geodetical  investigations.  In  meteorological  charts  the  relation 
between  the  direction  of  the  wind,  and  the  height  of  the  barometer, 
or  the  temperature,  is  often  plotted  in  polar  coordinates.  The 
treatment  of  problems  involving  direction  in  space,  displacement, 
•velocity,  acceleration,  momentum,  rotation,  and  electric  current 
are  often  simplified  by  the  use  of  vectors.  But  see  O.  Henrici 
and  G.  C.  Turner's  Vectors  and  Botors,  London,  1903,  for  a  simple 
exposition  of  this  subject. 

§  43.    Spiral  Curves. 

The  equations  of  the  spiral  curves  are  considerably  simplified 
by  the  use  of  polar  coordinates.  For  in- 
stance, the  curve  for  the  logarithmic  spiral 
(Fig.  35),  though  somewhat  complex  in 
Cartesian  coordinates,  is  represented  in 
polar  coordinates  by  the  simple  equation 

r  =  cfi,        .         .         (1) 
where  a  has  a  constant  value.     Hence 

log  r  =  6  log  a, 
35)  be  a  series  of  points  on  the  spiral  cor- 


FiG.  35. — Logarithmic 
Spiral. 

Let  0,  0V  G2, . . .  (Fig 

responding  with  the  angles    0V 

rlf  r9, . . .  Hence, 


0, 


and  the  radii  vectores 


§  43.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  117 

log  rx  =  0l  log  a  ;  log  r2  =  02  log  a . . . 
Since  log  a  is  constant,  say  equal  to  k, 

2 

that  is,  the  logarithm  of  the  ratio  of  the  distance  of  any  two  points 
on  the  curve  from  the  pole  is  proportional  to  the  angle  between 
their  radii  vectores.    If  rY  and  r2  lie  on  the  same  straight  line,  then 

0!-  02  =  2tt  =  360°;    and  log^  =  2&7T, 

r2 

7r  being  the  symbol  used,  as  in  trigonometry,  to  denote  180°. 

Similarly,  it  can  be  shown  that  if  r3,  r4 ...  lie  on  the  same 
straight  line,  the  logarithm  of  the  ratio  of  rx  to  r3,  r4 . . .  is  given  by 

4Zc7r,  6&7r This  is  true  for  any  straight  line  passing  through 

0  ;  and  therefore  the  spiral  is  made  up  of  an  infinite  number  of 
turns  which  extend  inwards  and  outwards  without  limit. 

If  the  radii  vectores  OG,  OD,  OE . .  .  OGv  OD1  ...  be  'taken  to 
represent  the  number  of  vibrations  of  a  sounding  body  in  a  given 
time,  the  angles  GOD,  DOE . . .  measure  the  logarithms  of  the 
intervals  between  the  tones  produced  by  these  vibrations.  A  point 
travelling  along  the  curve  will  then  represent  a  tone  continuously 
rising  in  pitch,  and  the  curve,  passing  successively  through  the 
same  line  produced,  represents  the  passage  of  the  tone  through 
successive  octaves  The  geometrical  periodicity  of  the  curve  is 
a  graphical  representation  of  the  periodicity  perceived  by  the  ear 
when  a  tone  continuously  rises  in  pitch. 

This  diagram  may  also  be  used  to  illustrate  the  Newlands- 
Mendeleeff  law  of  octaves,  by  arranging  the  elements  along  the 
curve  in  the  order  of  their  atomic  weights.  E.  Loew  (Zeit.  phys. 
Chem.,  23,  1,  1897)  represents  the  atomic  weight,  W,  as  a  function 
of  the  radius  vector,  r,  and  the  vectorial  angle,  0  :  W  =  f(r,  0),  so 
that  r  =  0  =  JW.  He  thus  obtains  W  =  rO.  This  curve  is  the 
well-known  Archimedes'  spiral.  If  r  is  any  radius  vector,  the 
distances  of  the  points  Pv  P2,  P3, . . .  from  0  are 

r2  =  r  +  7T ;  rA  =  r  +  Sir ;  r6  =  r  +  6w ;  . . . 
r3  =  r  +  2tt  ;  r5  =  r  +  4*r ;  r7  =  r  +  6?r ;  . . . 

Examples. — (1)  Plot  Archimedes'  spiral,  r  =  ad ;  and  show  that  the  re- 
volutions of  the  spiral  are  at  a  distance  of  2air  from  one  another. 

(2)  Plot  the  hyperbolic  spiral,  rd  =  a ;  and  show  that  the  ratio  of  the 
distance  of  any  two  points  from  the  pole  is  inversely  proportional  to  the 
angles  between  their  radii  vectores. 


118 


HIGHEK  MATHEMATICS. 


§44. 


§  $4.  Trilinear  Coordinates  and  Triangular  Diagrams. 

Another  method  of  representing  the  position  of  a  point  in  a 
plane  is  to  refer  it  to  its  perpendicular  distance  from  the  sides  of  a 

triangle  called  the  triangle  of  reference. 
The  perpendicular  distances  of  the 
point  from  the  sides  are  called  tri- 
linear coordinates.  In  the  equi- 
lateral triangle  ABG  (Fig.  36),  let  the 
perpendicular  distance  of  the  vertex  A 
from  the  base  BG  be  denoted  by  100 
units,  and  let  p  be  any  point  within  the 
B  triangle  whose  trilinear  coordinates  are 
Pig.  36.—  Trilinear  Coordinates.  Pa>  Pb>  Pc>  tnen 

pa  +  pb  +  pc  =  100. 
This  property1  has  been  extensively   used   in  the  graphic  repre- 
sentation of  the  composition  of  certain  ternary  alloys,  and  mixtures 
of  salts.     Each  vertex  is  supposed  to  represent  one  constituent  of 

the      mixture.       Any 
0CO3  point   within   the   tri- 

angle corresponds  to 
that  mixture  whose 
percentage  composi- 
tion is  represented  by 
the  trilinear  coordin- 
ates of  that  point. 
Any  point  on  a  side  of 
the  triangle  represents 
a  binary  mixture. 
Fig,  37  shows  the 
melting  points  of  ter- 
nary mixtures  of  iso- 
morphous  carbonates 
Such  a  diagram  is  sometimes 


BaC03 


Fig.  37.— Surface  of  Fusibility, 
of  barium,  strontium  and  calcium. 


SrC03 


called  a  surface  of  fusibility.     A  mixture  melting  at  670°  may 


1  It  is  not  difficult  to  see  this.  Through  p  draw pG  parallel  to  AG  cutting  AB  at 
G  ;  through  G  draw  GK  parallel  to  BG  cutting  AD  at  F,  and  AG  at  K\  produce  the 
line  ap  until  it  meets  GK  at  E ;  draw  GH  perpendicular  to  AG.  Now  show  that 
AF  =  HG  =  pb  ;  that  pE  =  pc  ;  that  1)F  =pc  +  pa  ;  and  that  DA  =•  pa  +  pb  +  pc. 


§  44.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  119 

have  the  composition  represented  by  any  point  on  the  isothermal 
curve  marked  670°,  and  so  on  for  the  other  isothermal  curves. 

In  a  similar  way  the  composition  of  quaternary  mixtures  has 
been  graphically  represented  by  the  perpendicular  distance  of  a 
point  from  the  four  sides  of  a  square. 

Roozeboom,  Bancroft  and  others  have  used  triangular  diagrams 
with  lines  ruled  parallel  to  each  other  as  shown  in  Fig.  38.    Sup- 

B 


A 

Fig.  38. — Concentration-Temperature  diagram. 

pose  we  have  a  mixture  of  three  salts,  A,  B,  C,  such  that  the  three 
vertices  of  the  triangle  ABC  represent  phases x  containing  100  °/0  of 
each  component.  The  composition  of  any  binary  mixture  is  given 
by  a  point  on  the  boundary  lines  of  the  triangle,  while  the  com- 
position of  any  ternary  mixture  is  represented  by  some  point  inside 
the  triangle. 

The  position  of  any  point  inside  the  triangle  is  read  directly 
from  the  coordinates  parallel  to  the  sides  of  the  triangle.  For 
instance,  the  composition  of  a  mixture  represented  by  the  point  0 
is  obtained  by  drawing  lines  from  0  parallel  to  the  three  sides  of 
the  triangle  OP,  OB,  OQ.  Then  start  from  one  corner  as  origin 
and  measure  along  the  two  sides,  AP  fixes  the  amount  of  C,  AQ 

1 A  phase  is  a  mass  of  uniform  concentration.  The  number  of  phases  in  a  system 
is  the  number  of  masses  of  different  concentration  present.  For  example,  at  the  tem- 
perature of  melting  ice  three  phases  may  be  present  in  the  H20-system,  viz.,  solid  ice, 
liquid  water  and  steam  ;  if  a  salt  is  dissolved  in  water  there  is  a  solution  and  a  vapour 
phase,  if  golid  salt  separates  out,  another  phase  appears  in  the  system. 


120  HIGHER  MATHEMATICS.  §  45. 

> 
the  amount  of  B,  and,  by  difference,  CB  determines  the  amount  A. 

For  the  point  chosen,  therefore  A  =  40,  B  =  40,  G  =  20. 

(i)  Suppose  the  substance  A  melts  at  320°,  B  at  300°,  and  G  at  305°,  and 
that  the  point  D  represents  an  eutectic  alloy *  of  A  and  C  melting  at  215°  ; 
Ey  of  an  eutectic  alloy  of  A  and  B  melting  at  207° ;  F,  of  an  eutectic  alloy  of 
B  and  C  melting  at  268°. 

(ii)  Along  the  line  DO,  the  system  A  and  G  has  a  solid  phase ;  along  EO, 
A  and  B  have  a  solid  phase  ;  and  along  FO,  B  and  G  have  a  solid  phase. 

(iii)  At  the  triple  point  O,  the  system  A,  B  and  G  exists  in  the  three-solid, 
solution  and  vapour-phases  at  a  temperature  at  186°  (say). 

(iv)  Any  point  in  the  area  A  DOE  represents  a  system  comprising  solid, 
solution  and  vapour  of  A, — in  the  solution,  the  two  components  B  and  C  are 
dissolved  in  A.  Any  point  in  the  area  CDOF  represents  a  system  comprising 
solid,  solution  and  vapour  of  C,— in  the  solution,  A  and  B  are  dissolved  in  C. 
Any  point  in  the  area  BEOF  represents  a  system  comprising  solid,  solution 
and  vapour  of  B, — in  the  solution,  A  and  G  are  dissolved  in  B. 

Each  apex  of  the  triangle  not  only  represents  100  °/0  of  a  substance,  but 
also  the  temperature  at  which  the  respective  substances  A,  B,  or  C  melt ; 
D,  E,  F  also  represent  temperatures  at  which  the  respective  eutectic  alloys 
melt.  It  follows,  therefore,  that  the  temperature  at  D  is  lower  than  at  either 
A  or  C.  Similarly  the  temperature  at  E  is  lower  than  at  A  or  Bt  and  at  F 
lower  than  at  either  B  or  G.  The  melting  points,  therefore,  rise  as  we  pass 
from  one  of  the- points  D,  E,  F  to  an  apex  on  either  side. 

For  details  the  reader  is  referred  to  W.  D.  Bancroft's  The  Phase  Bute, 
Ithaca,  1897. 

§  45.    Orders  of  Curves. 

The  order  of  a  curve  corresponds  with  the  degree  of  its  equa- 
tion. The  degree  of  any  term  may  be  regarded  as  the  sum  of  the 
exponents  of  the  variables  it  contains  ;  the  degree  of  an  equation 
is  that  of  the  highest  term  in  it.  For  example,  the  equation 
xy  +  x  +  b3y  =  0,  is  of  the  second  degree  if  b  is  constant ;  the 
equation  xz  +  xy  =  0,  is  of  the  third  degree  ;  x2yzs  +  ax  =  0,  is  of 
the  sixth  degree,  and  so  on.  A  line  of  the  first  order  is  repre- 
sented by  the  general  equation  of  the  first  degree 

ax  +  by  +  c  =  0 (1) 

This  equation  is  that  of  a  straight  line  only.  A  line  of  the  second 
order  is  represented  by  the  general  equation  of  the  second  degree 
between  two  variables,  namely, 

ax2  +  bxy  +  oy2  +  fx  +  gy  +  h  =  0.    .         .         (2) 

1  An  eutectic  alloy  is  a  mixture  of  two  substances  in  such  proportions  that  the 
alloy  melts  at  a  lower  temperature  than  a  mixture  of  the  same  two  substances  in  any 
other  proportions. 


§  40.        COORDINATE  OE  ANALYTICAL  GEOMETRY.  121 

This  equation  includes,  as  particular  oases,  every  possible  form  of 
equation  in  which  no  term  contains  x  and  y  as  factors  more  than 
twice.  The  term  bxy  can  be  made  to  disappear  by  changing  the 
direction  of  the  rectangular  axes,  and  the  terms  containing  fx  and 
gy  can  be  made  to  disappear  by  changing  the  origin  of  the  co- 
ordinate axes.  Every  equation  of  the  second  degree  can  be  made 
to  assume  one  of  the  forms 

ax2  +  cy2  -  h,  or,  y2  =  fx.  .  .  .  (3) 
The  first  can  be  made  to  represent  a  circle,1  ellipse,  or  hyperbola ; 
the  second  a  parabola.  Hence  every  equation  of  the  second  degree 
between  two  variables  includes  four  species  of  curves — circle, 
ellipse,  parabola  and  hyperbola. 

It  must  be  here  pointed  out  that  if  two  equations  of  the  first 
degree  with  all  their  terms  collected  on  one  side  be  multiplied 
together  we  obtain  an  equation  of  the  second  degree  which  is 
satisfied  by  any  quantity  which  satisfies  either  of  the  two  original 
equations.  An  equation  of  the  second  degree  may  thus  represent 
two  straight  lines,  as  well  as  one  of  the  above  species  of  curves. 

The  condition  that  the  general  equation  of  the  second  degree 
may  represent  two  straight  lines  is  that 

(bg  -  2c/)2  =  (b2  -  lac)  (g2  -  ±ch).     .         .         (4) 

The  general  equation  of  the  second  degree  will  represent  a 
parabola,  ellipse,  or  hyperbola,  according  as  b2  -  4ac,  is  zero, 
negative,  or  positive. 

Examples. — (1)  Show  that  the  graph  of  the  equation 
2a2  -  lOxy  +  12y2  +  5x  -  16y  -  3  =  0, 
represents  two  straight  lines.     Hint.  a«=2 ;  6=  -10  ;  e=12  ;  /=5  ;  g=  - 16 ; 
h  =  -  3  ;  (bg  -  2c/)2  -  1600 ;  (o2  -  4ac)  (g2  -  ±ch)  =  1600. 

(2)  Show  that  the  graph  of  a?3  -  2xy  +  y2  -  8x  +  16  =  0  represents  a 
parabola.     Hint.  From  (2),  62  -  4ac  =  -  2  x  -  2  -  4  x  1  x  1  =  0. 

(3)  Show,  that  the  graph  of  x2  -  6xy  +  y*  +  2x  +  2y  +  2  =  0  represents 
a  hyperbola.     Here  62  -  4ac  =  -  6  x  -  6  -  4  x  1  x  1  =  32. 

§  46.  Coordinate  Geometry  in  Three  Dimensions. — Geometry 

in  Space. 

Methods  have  been  described  for  representing  changes  in  the 
state  of  a  system  involving  two  variable  magnitudes  by  the  locus 
of  a  point  moving  in  a  plane  according  to  a  fixed  law  defined  by 

1  The  circle  may  be  regarded  as  an  ellipse  with  major  and  minor  axes  equal. 


122 


HIGHER  MATHEMATICS. 


§46. 


the  equation  of  the  curve.  Such  was  the  ^w-diagram  described  on 
page  111.  There,  a  series  of  isothermal  curves  were  obtained,  when 
0  was  made  constant  during  a  set  of  corresponding  changes  of  p 
and  v  in  the  well-known  equation  pv  =  BO. 

When  any  three  magnitudes,  x,  y,  z,  are  made  to  vary  together 
we  can,  by  assigning  arbitrary  values  to  two  of  the  variables,  find 
corresponding  values  for  the  third,  and  refer  the  results  so  obtained 
to  three  fixed  and  intersecting  planes  called  the  coordinate  planes. 
Of  the  resulting  eight  quadrants,  four  of  which  are  shown  in 
Fig.  39,  only  the  first  is  utilized  to  any  great  extent  in  mathe- 
matical physics.  This  mode  of  graphic  representation  is  called 
geometry  in  space,  or  geometry  in  three  dimensions.  The  lines 
formed  by  the  intersection  of  these  planes  are  the  coordinate 
axes.  It  is  necessary  that  the  student  have  a  clear  idea  of  a  few 
properties  of  lines  and  surfaces  in  working  many  physical  problems. 

If  we  get  a  series  of  sets 
of  corresponding  values  of  x, 
y,  z  from  the  equation 

x  +  y  =  z, 
and  refer  them  to  coordinate 
axes  in  three  dimensions,  as 
described  below,  the  result  is  a 
plane  or  surface.  If  one  of 
the  variables  remains  constant, 
the  resulting  figure  is  a  line. 
A  surface  may,  therefore,  be 
considered  to  be  the  locus  of 
Fig.  39. — Cartesian  Coordinates — Three   a  line  moving  in  space. 

I.  To  find  the  point  whose 
coordinates  OA,  OB,  OC  are  given.  The  position  of  the  point  P 
with  reference  to  the  three  coordinate  planes  xOy,  xOz,  yOz  (Fig. 
39)  is  obtained  by  dropping  perpendiculars  PL,  PM,  PN  from 
the  given  point  on  to  the  three  planes.  Complete  the  parallelo- 
piped,  as  shown  in  Fig.  39.  Let  OP  be  a  diagonal.  Then  LP 
=  OA,  PN  =  BO,  MP  =  OC.  Draw  three  planes  through  A,  B, 
C  parallel  respectively  to  the  coordinate  planes ;  the  point  of 
intersection  of  the  three  planes,  namely  P,  will  be  the  required 
point. 

If  the  coordinates  of  P,  parallel  to  Ox,  Oy,  Oz,  are  respectively 
x,  y  and  z,  then  P  is  said  to  be  the  point  x,  y,  z.     A  similar  con- 


§  46.        COORDINATE  OR  ANALYTICAL  GEOMETRY. 


123 


vention  with  regard  to  the  sign  is  used  as  in  analytical  geometry  of 
two  dimensions.  It  is  conventionally  agreed  that  lines  measured 
from  below  upwards  shall  be  positive,  and  lines  measured  from 
above  downwards  negative;  lines  measured  from  left  to  right 
positive,  and  from  right  to  left  negative ;  lines  measured  inwards 
from  the  plane  of  the  paper  are  negative,  lines  measured  towards 
the  reader  are  positive. 

If  a  watch  be  placed  in  the  plane  xy  with  its  face  pointing  up- 
wards, towards  +  z,  the  hands 
of  the  watch  move  in  a  negative 
direction ;  if  the  watch  be  in  the 
xz  plane  with  its  face  pointing 
towards  the  reader,  the  hands 
also  move  in  a  negative  direction. 

II.  To  find  the  distance  of  a 
point  from  the  origin  in  terms  of 
the  rectangular  coordinates  of 
that  point.  In  Fig.  40,  let  Ox, 
Oy,  Oz  be  three  rectangular  axes, 
P(x,  y,  z)  the  given  point  such 
that  MP  =  z,  AM  =  y,  OA  =  x. 
OP  —  r,  say. 

OP2  =  OM2  +  MP2 ;  or,  r2  =  OM2  +  z\ 
but  OM2  =  AM*  +  OA2  =  x2  +  y2. 

.-.  r2  =  x2  +  y2  +  z2         .         .         .         (1) 

In  words,  the  sum  of  the  squares  of  the  three  coordinates  of  a 
point  are  equal  to  the  square  of  the  distance  of  that  point  from  the 
origin. 

Example. — Find  the  distance  of  the  point  (2a,  -  3a,  6a)  from  the  origin. 
Hint,  r  =  \/4a2  +  9a2  +  36a2  =  la. 

Let  the  angle  AOP  =  a ;  BOP  =  /? ;  POG  =  y,  then 
x  =  r  cos  a ;  y  =  r  cos  (3;  z  =  r  cos  y.    . 


B "M 

Fig.  40. 

It  is  required  to  find  the  distance 


(2) 


These  equations  are  true  wherever  the  point  P  may  lie,  and 
therefore  the  signs  of  x,  y,  z  are  always  the  same  as  those  of  cos  a, 
cos  (3,  cos  y  respectively.  Substituting  these  values  in  (1),  and 
dividing  through  by  r2,  we  get  the  following  relation  between  the 
three  angles  which  any  straight  line  makes  with  the  coordinate  axes 


124  HIGHER  MATHEMATICS.  §  46. 

COS2a  +  COS2^S  +  COS2y  ml.-.  .  .  (3) 

The  cosines  of  the  angles  a,  /?,  y  which  the  given  line  makes 
with  the  axes  x,  y,  z  respectively  are  called  the  direction 
cosines,  and  are  often  symbolized  by  the  letters  lt  in,  n.  Thus 
(3)  becomes 

P  +  m2  +  n2  =  1. 
If  we  know  r,  cos  a,  cos  ft,  and  cos  y  we  are  able  to  fix  the 
position  of  the  point.     If  a,  b,  c  are  proportional  to  the  direction 
cosines  of  some  line,  we  can  at  once  find  the  direction  cosines- 
For,  from  page  23,  if 

I:  a  =  m  :  b  =  n  :  c ;  .  •.  I  =  ra\  m=  rb  ;  n  =  re. 
Substitute  in  the  preceding  equation,  and  we  get  at  once 
a  b  c 


I 


Ja2  +  ftl  +  & 


m  m 


n  = 


J  a2  +  b2  +  c2'  J  a2  +  b2  +  c2 


Example. — The  direction  cosines  of  a  line  are  proportional  to  3,  -  4, 
and  2.  Find  their  values.  Ansr.  3  n^,  -  4  J~fo,  2  *Jfa.  Hint,  a =3,  b=  -  4, 
c=2. 


III.  To  find  the  distance  between  two  points  in  terms  of  their 

rectangular  coordinates.  Let  P1(xv 
yv  ^i),P2(x2,  Vz*  z<i)  ^e  tne  given  points, 
it  is  required  to  find  the  distance  P1P2 
in  terms  of  the  coordinates  of  the 
points  Pj  and  P2.  Draw  planes 
through  P1  and  P2  parallel  to  the 
coordinate  planes  so  as  to  form  the 
parallelepiped  ABCDE.  Join  P2E. 
By  the  construction  (Fig.  41),  the 
angle  PYEP2  is  a  right  angle. 
Fio.  41.  Hence 

(PXP2)2  =  (P.E)2  +  (P2E)2  =  (P.E)2  +  (ED)2  +  (P2D)2. 

But  PYE  is  evidently  the  difference  of  the  distance  of  the  foot  of 
the  perpendiculars  from  Px  and  P2  on  the  #-axis,  or  PXE  =  x2  -  xv 


Similarly,  ED  =  y2  -  y1 ;  P2D  m  *s  -  zv     Hence 

r2  -  (x2  -  x,)2  +  (y2  -  y,f+  (z2  -  z,)2. 


W 


Example. — Find  the  distance  between  the  points  (3,  4,  -  2)  and  (4,  -  3, 1). 
Here  x}  =  4 ;  yx  =  -  3 ;  zx  =  1 ;  x.,  =  3 ;  y2  =  4 ;  «2  =  -  2.     Ansr.  r  =  \/59. 


§  46.   COORDINATE  OR  ANALYTICAL  GEOMETRY. 


125 


IV.  Polar  coordinates.  Instead  of  referring  the  point  to  its 
Cartesian  coordinates  in  three  dimen- 
sions, we  may  use  polar  coordinates. 
Let  P  (Fig.  42)  be  the  given  point 
whose  rectangular  coordinates  are  x,  y, 
z ;  and  whose  polar  coordinates  are  r, 
6,  <f>,  as  shown  in  the  figure. 

I.  To   pass   from    rectangular   to 
jiolar  coordinates.     (See  page  96.) 
x  =  OA  =  OMooBcf>  -  rsin#.cos<M 
y  =  AM  =  OMsin<£  =  rsin0.sin<£l    (5) 
z  =  MP  =  r  cos  6.  J 

II.  To  pass  from  polar  to  rectangu-  Fig.  42^ -Polar  Coordinates  in 

r        J  r  v  Three  Dimensions. 

lar  coordinates. 

r  =  J  {a?  +  y*  +  z2) ;  $  -  tan 


^x/C^+j/!) 


<£  =  tan 


-'I    (6) 


Examples.— (1)  Find  the  rectangular  coordinates  of  the  point  (3,  60°,  30°) 
Ansr.  (|,  £n/3,$). 

(2)  Find  the  polar  coordinates  of  the  point  (3,  12,  4).  Ansr.  The  point 
(13,  tan  "  1  i  n/153,  tan  -  H). 

According  to  the  parallelogram  of  velocities,  "  if  two  com- 
ponent velocities  OA,  OB  (Fig.  43)  are  represented  in  direction  and 
magnitude  by  two  sides  of  a  parallelogram  drawn  from  a  point,  0, 
the  resultant  velocity  can  be  represented  in  direction  and  magni- 
tude by  the  diagonal,  OP,  of  the  parallelogram  drawn  from  that 
point".  The  parallelopiped  of  velocities  is 
an  extension  of  the  preceding  result  into  three 
dimensions.  "  If  three  component  velocities 
are  represented  in  direction  and  magnitude  by 
the  adjacent  sides  of  a  parallelopiped,  OA,  OC, 
OB  (Fig.  42),  drawn  from  a  point,  0,  their 
resultant  velocity  can  be  represented  by  the 
diagonal  of  a  parallelopiped  drawn  from  that  point."  Conversely,  if 
the  velocity  of  the  moving  system  is  represented  in  magnitude  and 
direction  by  the  diagonal  OP  (Fig.  42)  of  a  parallelopiped,  this  can 
be  resolved  into  three  component  velocities  represented  in  direction 
and  magnitude  by  three  sides  x,  y,  z  of  the  parellelopiped  drawn 
from  a  point. 

We  assume  that  if  any  contradictory  facts  really  existed   we 


0  x- component 


Fig.  43.— Parallelo- 
gram of  Velocities. 


y 

Bl        A 

17  k 

p, 

/ee\      1 

dx 

y 

/oo\ 

X 

V-    X 

M 

M 

126  HIGHER  MATHEMATICS.  §  46. 

should  have  known  them  long  ago.  A  continental  text-book  has 
forty-five  theoretical  demonstrations  of  this  important  principle. 
But  we  are  slowly  learning  the  lesson  taught  by  John  Stuart  Mill 
that  the  "real  and  only  proof  of  any  law  of  Nature. .  .is  experi- 
ence ".  The  daily  comparison  of  a  new  rule  with 
experience,  and  the  testing  of  its  consequences 
under  the  most  diverse  conditions  is,  after  the 
lapse  of  a  reasonable  period  of  time,  a  more  satis- 
factory proof  than  clumsy  deductions  drawn  from 
obscure  premises.1 

If  the  point  P  travels  along  the  path  APB 
(Fig.  44)  so  as  to  trace  a  path  s  units  long,  then,  when  x  =  OM, 
and  y  =  MP,  let  dx,  dy,  and  ds  be  infinitesimals  such  that 

dx  dy        .  dx     ds  dy     ds  .  ,„. 

5-s  =  oosa;i  =  sma;  or^=sC0So;i=^slIla    <7> 

and 

The  corresponding  formulae  in  three  dimensions  are  very  obvious. 

Examples. — (1)  A  comet  moves  upon  the  parabolic  path  y2=±ax ;  find  its 
rate  of  approach  to  the  sun  which  is  placed  at  the  focus  of  its  orbit.  Let  r 
denote  the  distance  from  the  focus  to  any  point  P(x,  y)  on  the  parabola. 
Hence,  from  the  definition  of  a  parabola  r  =  x  +  a;  .-.  drjdt  =  dx/dt.  Or  its 
rate  of  approach  to  the  sun  is  the  same  as  its  horizontal  velocity.  Let  s  de- 
note the  length  of  the  path,  then  dsjdt  —  velocity  of  motion  =  V,  say.  But  by 
differentiation  of  the  given  equation, 

dy_2a    dx.    ,    _  =  (dsV    (dx\2    ^fdx\2    .    dx_       yV 
dt~  y  '  dt'  "  \dt)      \dt)  +  y*\dt)    **'    dt~~  Jjf^Jjj? 

or  the  comet  approaches  the  sun  with  y{yl  +  4a2)~i  times  its  velocity.  At  the 
vertex  of  the  parabola,  y  =  0,  dx/dt  =  0,  or  the  comet  is  not  approaching  the 
sun  at  all. 

(2)  Show  that  the  ordinate  of  a  point  moving  on  the  parabola  y2  =  4aj 
changes  2/y  times  as  fast  as  the  abscissa ;  and  if,  at  the  point  x  =  4,  the 
abscissa  is  changing  at  the  rate  of  20  ft.  per  second,  at  what  rate  is  the 
ordinate  changing  ?    Hint.  If  x  =  4,  y  =  ±  4  ;  hence 

dy  _  2    dx  >      dy  _     1    dx 
~di~y'  di'  '''1t~-2'  ~dt' 

1  E.  Mach.  Of  course  we  only  deal  with  one  velocity.  The  resolving  of  one 
velocity  into  three  component  velocities  is  a  mathematical  fiction  to  assist  reasoning. 
This  is  not  necessary  in  "Vector  analysis,"  page  116,  which  has  replaced  "coordinate 
geometry  "  in  the  mathematical  treatment  of  many  physical  problems. 


§  47.        OOOKDINATE  OR  ANALYTICAL  GEOMETRY.         127 

Hence  dyjdt  =  ±  %  x  20  =  +  10.  Or  the  ordinate  increases  or  decreases  at 
the  rate  of  10  ft.  per  sec. 

(3)  Let  a  particle  move  with  a  velocity  V  in  space.  From  the  parallelo- 
piped  of  velocities,  I7"  can  be  resolved  into  three  component  velocities  Vlt  V2,  V3, 
along  the  x-,  y-  and  s-axes  respectively.     Hence  show  that 

dx  _  V  •  dy  -  V  ■  dz  -  V  to\ 

di-Vl'di-v*'di-y»'      '      •      '      (y) 

which  may  be  written 

dx  =  d{VJ  +  x0) ;  dy  =  d{V2t  +  yQ) ;  dz  =  d{V,t  +  z0),  .  (10) 
where  x0,  y0,  z0  are  constants.  Hence  we  may  write  the  relation  between  the 
space  described  by  the  particles  in  each  dimension  and  the  time  as 

x  =  V,t  +  x0 ;  y  =  V2t  +  y0 ;  z  =  Vzt  +  z0.  .        .        .      (11) 

Obviously  x0,  y0,  *0  are  the  coordinates  of  the  initial  position  of  the  particle 
when  t  =  0.     Hence  xa,  y0i  z0  are  to  be  regarded  as  constants. 

If  the  reader  cannot  follow  the  steps  taken  in  passing  from  (9)  to  (11),  he 
can  take  Lagrange's  advice  to  the  student  of  a  mathematical  text-book  :  "  Allez 
en  avant,  et  la  foi  vous  viendra,"  in  other  words,  "  go  on  but  return  to  strengthen 
your  powers.     Work  backwards  and  forwards  ". 

Obviously,  xQ,  yQ,  z0  represent  the  positions  of  the  particle  at  the  beginning 
of  the  observation,  when  t  =  0.  Let  s  denote  the  length  of  the  path  traversed 
by  the  particle  at  the  time  t,  when  the  coordinates  of  the  point  are  x,  y  and  z. 
Obviously,  by  the  aid  of  Fig.  41, 


s  .  J(x  -  x0)*  +  (y-  y0)»  +  (z-  z0)*;  or,  s  =  s/Vf+Vf+V* .t, 
from  (11).  s  can  therefore  be  determined  from  the  initial  and  final  positions 
of  the  particle. 

§  57.  Lines  in  Three  Dimensions. 

I.  To  find  the  angle  between  two  straight  lines  whose  direction 
cosines  are  given.  Join  OPl  (Fig.  41)  and  OP2.  Let  if/  be  the 
angle  between  these  two  lines.  In  the  triangle  P2OP1  if  OPx  =  rv 
OP2  =  r2,  PXP2  =  T,  we  get  from  the  properties  of  triangles  given 
on  page  603, 

r2  =  rx2  +  r22  -  2r^2  eos  f. 

Rearranging  terms  and  substituting  for  rY  and  r2  in  (1),  we  obtain 

rx2  =  x*  +  yi*  +  z* ;  r22  =  z22  +  y2*  +  %*, 

xxx2  +  yYy2  +  zxz2 
.'.  cos  f  =  — 

rlr2 

We  can  express  this  another  way  by  substituting, 

xY  =  rY  cos  ttj ;  x2  =  r2  cos  a2 ;  y2  =  r2  cos  fi2  . . .  , 
as  in  (2),  and  we  obtain 

COS  if/  =  COS  ax  .  COS  a2  +  COS  /?x  .  COS  {32  +  COS  yx  .  COS  y2        (12) 


128  HIGHER  MATHEMATICS  §  47. 

or,  cos  \J/  =  lYl2  +  mim.z  +  nxn^         .         .         (13) 

where  \f/  represents  the  angle  between  two  straight  lines  whose 
direction  cosines  are  known. 

(i)  When  the  lines  are  perpendicular  to  one  another,  if/  =  90°, 
.'.  cosi/r  =  oos90°  =  0,  and  therefore 

COS  <*!  .  COS  a2  +  COS  /3X  .  COS  fi2  +  COS  y1  .  COS  y2  =  0,  (14) 

or,  xxx2  +  yYy2  +  zxz2  =  0. 

(ii)  If  the  two  lines  are  parallel, 

«i  =  a2 ;  A  =  & ;  ti  =  y2       •      •      (15) 

Examples. — (1)  Find  the  acute  angle  between  the  lines  whose  direction 
cosines  are  \  V3,  J,  \  V3,  and  \  \/3,  \y  -  \  s/'S.  Hint.  ^  =  l2 = \  \/3 ;  m^ = w2 = £ ; 
n1  =  bs/W\  nt  =  -  %>J3.    Use  (13).     Cos  if,  =  -  £ ;  .-.  »//  =  60°. 

(2)  Let  7lt  72,  73  be  the  velocity  components  (page  125)  of  a  particle 

moving  with  the  velocity  7;  let  o,  £,  7  be  the  angles  which  the  path  described 

by  the  moving  particle  makes  with  the  x-%  y-  and  s-axes   respectively,  then 

show 

,  ,  dxldt    1    dx  ..,_. 

ds  .  00s  a  =  dz;  .-.  tefjrrm  =  cos  a.  .        .        .      (16) 

Hence, 

71  =  ^-7ooBa;7a  =  ^-7oo8iB;    73  =  ^=Fcos7;      .      (17) 
and  consequently,  from  (3), 


7  =  ds/dt  =  s]V*  +  722  +  732 (18) 

The  resolved  part  of  7  along  a  given  line  inclined  at  angles  au  &iy  yx  to  the 
axes  will  be 

7cos^=  71cosa1  +  FaCosjSi  +  730037^  .        .       (19) 

where  ty  denotes  the  angle  which  the  path  described  by  the  particle  makes 
with  the  given  line.     Hint.  Multiply  (12)  by  7,  etc. 

(3)  To  find  the  direction  of  motion  of  the  particle  moving  on  the  line  s 
(i.e.,  r  of  Fig.  40).  Let  a,  £,  7  denote  the  angles  made  by  the  direction  of  s 
with  the  respective  axes  x,  y,  e.     With  the  same  notation, 

008a  =  ^-^-0;    cosj8  =  ^-^°;   cos  7  =  *-ZJ*m    .        .      (20) 
s  s  '  s  x     ' 

Now  introduce  the  values  of  x  -  x0,  y  -  yQ1  and  of  z  -  z0  from  (20),  and  show 
that 

cos  a :  cos  /8 :  cos  7  =  VY :  72 :  73.  .        .        .       (21) 

II.  Projection.  If  a  perpendicular  be  dropped  from  a  given 
point  upon  a  given  plane  the  point  where  the  perpendicular  touches 
the  plane  is  the  projection  of  the  point  P  upon  that  plane.  For 
instance,  in  Fig.  39,  the  projection  of  the  point  P  on  the  plane 
xOy  is  M,  on  the  plane  xOz  is  N,  and  on  the  plane  yOz  is  L. 


§  47.        COORDINATE  OE  ANALYTICAL  GEOMETRY. 


129 


Similarly,  the  projection  of  the  point  P  upon  the  lines  Ox,  Oy,  Oz 
is  at  A,  B  and  G  respectively. 


Fig.  45.— Projecting  Plane. 


Fig.  46. 


In  the  same  way  the  projection  of  a  curve  on  a  given  plane 
is  obtained  by  projecting  every  point  in  the  curve  on  to  the  plane. 
The  plane,  which  contains  all  the  perpendiculars  drawn  from  the 
different  points  of  the  given  curve,  is  called  the  projecting  plane. 
In  Fig.  45,  CD  is  the  projection  of  AB  on  the  plane  EFG;  ABCD 
is  the  projecting  plane. 

12 


Fig.  47. 

Examples. — (1)  The  projection  of  any  given  line  on  an  intersecting  line 
is  equal  to  the  product  of  the  length  of  the  given  line  into  the  cosine  of  the 
angle  of  intersection.  In  Fig.  46,  the  projection  of  AB  on  CD  is  AE,  but 
AE  =  AB  cos  0. 

(2)  In  Fig.  47,  show  that  the  projection  of  OP  on  OQ  is  the  algebraic  sum 
of  the  projections  of  OA,  AM,  MP,  taken  in  this  order,  on  OQ.  Hence,  if 
OA  =  x,OB  =  AM  =  y,  OC  =  PM  =  z  and  OP  =  r,  from  (12) 


r  cos  \p  =  x  cos  a  +  y  cos  &  +  z  cos  y. 


(22) 


III.  The  equation  of  a  straight  line  in  rectangular  coordinates. 
Suppose  a  straight  line  in  space  to  be  formed  by  the  intersection  of 
two  projecting  planes.  The  coordinates  of  any  point  on  the  line 
of  intersection  of  these  planes  will  obviously  satisfy  the  equation  of 

I 


130 


HIGHER  MATHEMATICS. 


§47. 


Fig.  48. 


each  plane.     Let  ab,  a'b'  be  the  projection  of  the  given  line  AB  on 

the  xOz  (where  y  =  0)  and  the  yOz 
(where  #  =  0)  planes,  then  (Fig.  48), 
x  =  mz  +  c  ;  y  =  m'z  +  c'  (23) 
Of  the  fonr  independent  constants, 
m  represents  the  tangent  of  the 
angle  which  the  projection  of  the 
given  line  on  the  xOz  plane  makes 
with  the  #-axis ;  m'  the  tangent 
of  the  angle  made  by  the  line  pro- 
jected on  the  yOz  plane  with  the 
2/-axis ;  c  is  the  distance  intercepted 
by  the  projection  of  the  given  line 
along  the  #-axis ;  c'  a  similar  intersection  along  the  ?/-axis.  Hence 
we  infer  that  two  simultaneous  equations  of  the  first  degree  re- 
present a  straight  line. 

Example. — The  equations  of  the  projections  of  a  straight  line  on  the 
coordinate  planes  xz  and  zy  are  x  =  2z  +  3,  and  y  =  Sz  -  5.  Show  that  the 
equation  of  the  projection  on  the  xy  plane  is  2y  =  3x  -  19.  c'  =  -  5  ;  c  =  3 ; 
m  =  2 ;  m'  =  3  ;  eliminate  z ;  etc.     Ansr.  2y  -  3x  +  19  =  0. 

If,  now,  a  particular  value  be  assigned  to  either  variable  in 
either  of  these  equations,  the  value  of  the  other  two  can  be  readily 
calculated.  These  two  equations,  therefore,  represent  a  straight 
line  in  space. 

The  difficulties  of  three-dimensional  geometry  are  greatly  les- 
sened if  we  bear  in  mind  the  relations  previously  developed  for 
simple  curves  in  two  dimensions.  It  will  be  obvious,  for  instance, 
from  page  94,  that  if  the  straight  line  is  to  pass  through  a  given 
point  (xv  yv  i J,  the  coordinates  of  the  given  point  must  satisfy  the 
equations  of  the  curve.     Hence,  from  (23),  we  must  also  have 

iPj  =  mzl  +  c ;  yx  =  m!zx  +  c'.       .        .         (24) 
Subtracting  (24)  from  (23),  we  get 

x  -  xx  =  m(z  -  zx) ;  y  -  y1  =  m'(z  -  zj  .  (25) 
which  are  the  equations  of  a  straight  line  passing  through  the 
point  xv  yv  zx. 

If  the  line  is  to  pass  through  two  points  xv  yv  zv  and 
#2>  y2>  zv  we  &et>  by  tlie  metn°d  of  page  94, 

s  -  xi  =  x2  ~  xi .  y  ~  Vi  =  V2  -Vi  t         \         /26) 


§  47.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  131 

which  are  the  equations  of  the  straight  line  passing  through  the 
two  given  points. 

Example. — Show  that  the  equations  x  +  8z  =  19  ;  y  =  lOz  -  24  pass 
through  the  points  (3,  -  4,  2)  and  ( -  5,  6,  3). 

If  x,  y,  z  denote  the  coordinates  of  any  point  A  on  a  given 
straight  line ;  and  xvyvzv  the  known  coordinates  of  another  point 
P  on  the  straight  line  such  that  the  distance  between  A  and  P  is  r, 
then  it.  can  be  shown  that  the  equation  of  the  line  assumes  the 
symmetrical  form  :  r  = 

x-x1  =  y_1y1  =  z_^zi  _  ■  * 

I  m  n  v 

where  Z,  m  and  n  are  the  direction  cosines  of  the  line.  This  equa- 
tion gives  us  the  equation  of  a  straight  line  in  terms  of  its  direction 
cosines  and  any  known  point  upon  it.  (27)  is  called  the  sym- 
metrical equation  of  a  straight  line. 

Example. — If  a  line  makes  angles  of  60°,  45°,  and  60°  respectively  with 
the  three  axes  x,  y  and  z,  and  passes  through  the  point  (1,  -  3,  2),  show  that 
the  equation  of  the  line  is  x  -  1  =  sl\{y  +  3)  =  z  -  2.  Hint.  Cos  60°  =  £ ; 
cos45°  =  \/J7 

If  the  two  lines 

x  =  mxz  +  c1;  y  =  m{z  +  c/ ;  .  .  (28) 
x  =  m2z  +  c2 ;  y  =  m2'z  +  c2',  .  .  (29) 
intersect,  they  must  have  a  point  in  common,  and  the  coordinates 
of  this  point  must  satisfy  both  equations.  In  other  words,  x,  y 
and  z  will  be  the  same  in  both  equations — x  of  the  one  line  is  equal 
to  x  of  the  other. 

.-.  (m1  -  m2)z  +  gx  -  c2  =  0,         .         .         (30) 
(mx'  -  m2)z  +  Cj'  -  c2  =  0.         .         .         (31) 
But  the  z  of  one  line  is  also  equal  to  z  of  the  other,  hence,  if 
the  relation 

W  -  G2)  K  -  ™2)  =  K  -  c2)  K'  -  ™>2)>    •      (32) 

subsists  the  two  lines  will  intersect. 

Example. — Show  that  the  two  lines  x  =  Sz  +  7,  y=5z+8;  &ndx=2z  +  3. 
y  =  ±z  +  4  intersect.     Hint.  (8  -  4)  (3  -  2)  =  (7  -  3)  (4  -  3). 

The  coordinates  of  the  point  of  intersection  are  obtained  by 
substituting  (30)  or  (31)  in  (28),  or  (29).  Note  that  if  m1  =  m{  or 
m2  =  m2,  the  values  of  x,  y  and  z  then  become  infinite,  and  the 
two  lines  will  be  in  parallel  planes ;  if  both  ml  =  m±  and  m2  =  m2', 
they  will  be  parallel. 


132 


HIGHER  MATHEMATICS. 


§48. 


§  58.    Surfaces  and  Planes. 

L  To  find  the  equation  of  a  plane  surface  in  rectangular  co- 
ordinates. Let  ABC  (Fig. 
49)  be  the  given  plane 
whose  equation  is  to  be 
determined.  Let  the 
given  plane  cut  the  co- 
ordinate axes  at  points 
A,  B,  C  such  that  OA  =  a, 
OB  =  b,  OC  =  c.  From 
.*•  any  point  P(x,  y,  z)  drop 
the  perpendicular  PM  on 
to  the  yOx  plane.  Then 
OA'  =  xy  MA'  =  y  and 
MP  =  z.  It  is  required  to  find  an  equation  connecting  the  co- 
ordinates x,  y  and  z  respectively  with  the  intercepts  a,  b,  c.  From 
the  similar  triangles  AOB,  AA'B\ 

OA:BO  =  A' A  :  B'A' ;  or,  a  :  b  =  a  -  x  :  B'A', 


Fig.  49. 


B'A'  =  b  -  -  ;  also  B'M  =  B'A' 
a 


MA'  =  b 


V 


bx 

a ' 


Again,  from  the  similar  triangles  COB,  C'A'B',  PMB',  page  603, 
OC'.BO  =  MP:  B'M; 


or, 


c:b  =  z  \b  -  y 


bx         _         _  bcx 

-;  .:bz  =  bc-cy-—. 


Divide  through  by  be ;  rearrange  terms  and  we  get  the  intercept 
equation  of  the  plane,  i.e.,  the  equation  of  a  plane  expressed  in 
terms  of  its  intercepts  upon  the  three  axes : 


x      y      z 
-  +  |  +  -  =  l 
a      b       c        ' 


(33) 


an  equation  similar  to  that  developed  on  page  90.  In  other  words, 
equation  (33)  represents  a  plane  passing  through  the  points  (a,  0,  0), 
(0,  b,  0),  (0,  0,  c). 

If  ABC  (Fig.  49)  represents  the  face,  or  plane  of  a  crystal,  the  intercepts 
a,  b,  c  on  the  x-,  y-  and  s-axes  are  called  the  parameters  of  that  plane.  The 
parameters  in  crystallography  are  usually  expressed  in  terms  of  certain  axial 
lengths  assumed  unity.       If  OA  =  a,  OB=  b,  OC  =  c,  any  other  plane,  whose 


§  48.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  133 

intercepts  on  the  x-,  y-  and  s-axes  are  respectively  p,  2  and  r,  is  defined  by 
the  ratios 

a    b    c 

p'q-r 
These  quotients  are  called  the  parameters  of  the  new  plane.  The  reciprocals 
of  the  parameters  are  the  indices  of  a  crystal  face.  The  several  systems  of 
cry  stall  ographic  notation,  which  determine  the  position  of  the  faces  of  a 
crystal  with  reference  to  the  axes  of  the  crystal,  are  based  on  the  use  of 
parameters  and  indices. 

We  may  write  equation  (33)  in  the  form, 

Ax  +  By  +  Gz  +  D  =  0,  .        .        (34) 

which  is  the  most  general  equation  of  the  first  degree  between 
three  variables.  Equation  (33)  is  the  general  equation  of  a 
plane  surface.  It  is  easily  converted  into  (34)  by  substituting 
Aa  +  D  -  0,  Bb  +  D  -  0,  Cc  +  D  =  0. 

Examples. — (1)  Find  the  equation  of  the  plane  passing  through  the  three 
points  (3,  2,  4),  (0,  4,  1),  and  (-  2,  1,  0).  Ansr.  11a;  -  Sy  -  ldz  +  25  =  0. 
Hint.  From  (33), 

3       2      4  4       1  2       1^  25     ,       25  25 

(2)  Find  the  equation  of  the  plane  through  the  three  points  (1,  0,  0), 
(0,  2,  0),  (0,  0,  3).     Ansr.  0  +  \y  +  \z  =  1.     Use  (33)  or  (34). 

If  OQ  =  r  (Fig.  49)  be  normal,  that  is,  perpendicular  to  the 
plane  ABG,  the  projection  of  OP  on  OQ  is  equal  to  the  sum  of  the 
projections  of  OA\  PM,  MA'  on  OQ,  Ex.  (2),  page  129.  Hence,  the 
perpendicular  distance  of  the  plane  from  the  origin  is 

x  cos  a  +  y  cos  p  +  z  cos  y  =  r.       .         .        (35) 
This  is  called  the  normal  equation  of  the  plane,  that  is,  the 
equation  of  the  plane  in  terms  of  the  length  and  direction  cosines 
of  the  normal  from  the  origin.     From  (34),  we  get 
cos2a  :  cos2£  :  cos2y  =  A2 :  B2 :  C2 ; 
and  by  componendo,1 

(C0S2a  +  COS2/?  +  COS2y)  !  COS2a  =  A2  +  B2  +  C2  I  A2. 

But  by  (3),  the  term  in  brackets  on  the  left  is  unity,  consequently 


1  If  a,  b,  c  and  d  are  proportional,  the  text-books  on  algebra  tell  us  that 
a:  b  =  c  :  d\  and  it  therefore  follows  by  "invertendo"  :  b  :  a  =  d  :  c;  and  by 
"alternando "  :  a  :  c  =  b  :  d ;  and  by  " componendo"  :  a  +  b  :  b  =  c  +  d  :  d;  and 
by  "dividendo"  :  a—b:b  =  c-d:d;  and  by  "  convertendo  "  :  a  :  a-b  =  c:  c-d 
and  by  "  componendo  et  dividendo  ":  a±b:a+b  =  c±d:  c  +  d. 


134  HIGHER  MATHEMATICS.  §  48. 

the  direction  cosines  of  the  normal  to  the  plane  are 

A 
COS  a  =  — ,  -  ; 

J  A2  +  B2  +  G2 ' 

B 

cos  3  =      ,  ; 

H       JA2  +  B2  +  G2' 
C 

COS  y  =       , 

1       si  A2  +  B2  +  C2 
The  ambiguity  of  sign  is  removed  by  comparing  the  sign  of  the 
absolute  term  in  (34)  and  (35).     Dividing  equation  (34)  through 
with  +  J  A2  +  B2  +  G2,  we  can  write 

r  = D  .        .        .      (36) 

JA2{  +B2  +  C2 

Example. — Find  the  length  of  the  perpendicular  from  the  origin  to  the 
plane  whose  equation  is  2x  -  Ay  +  z  -  8  =  0.  Ansr.  8  */jf.  Hint.  A  =  2, 
B  =  -  4,  C  =  1,  D  =  8.     Use  the  right  member  of  the  equation  (36). 

II.  Surfaces  of  revolution.  Just  as  it  is  sometimes  convenient 
to  suppose  a  line  to  have  been  generated  by  the  motion  of  a  point, 
so  surfaces  may  be  produced  by  a  straight  or  curved  line  moving 
according  to  a  fixed  law  represented  by  the  equation  of  the  curve. 
The  moving  line  is  called  the  generator.  Surfaces  produced  by  the 
motion  of  straight  lines  are  called  ruled  surfaces.  When  the 
straight  line  is  continually  changing  the  plane  of  its  motion,  twisted 
or  skew  surfaces — surfaces  gauches — are  produced.  Such  is  the 
helix,  the  thread  of  a  screw,  or  a  spiral  staircase.  On  the  other 
hand,  if  the  plane  of  the  motion  of  a  generator  remains  constant,  a 
developable  surface  is  produced.  Thus,  if  the  line  rotates  round 
a  fixed  axis,  the  surface  cut  out  is  called  a  surface  of  revolution. 
A  sphere  may  be  formed  by  the  rotation  of  a  circle  about  a  diameter ; 
a  cylinder  may  be  formed  by  the  rotation  of  a  rectangle  about  one 
of  its  sides  as  axis ;  a  cone  may  be  generated  by  the  revolution 
of  a  triangle  about  its  axis  ;  an  ellipsoid  of  revolution,  by  the  rota- 
tion of  an  ellipse  about  its  major  or  minor  axes  ;  a  paraboloid,  by 
the  rotation  of  a  parabola  about  its  axis.  If  a  hyperbola  rotates 
about  its  transverse  axis,  two  hyperboloids  will  be  formed  by  the 
revolution  of  both  branches  of  the  hyperbola.  On  the  other  hand, 
only  one  hyperboloid  is  formed  by  rotating  the  hyperbolas  about 
their  conjugate  axes.  In  the  'former  case,  the  hyperboloid  is  said 
to  be  of  two  sheets,  in  the  latter,  of  one  sheet. 


§  49.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  135 


III.  To  find  the  equation  of  the  surface  of  a  right  cylinder. 
Let  one  side  of  a  rectangle  rotate  about 
Oz  as  axis.  Any  point  on  the  outer  edge 
will  describe  the  circumference  of  a  circle. 
If  P(x,  y,  z)  (Fig.  50)  be  any  point  on 
the  surface,  r  the  radius  of  the  cylinder, 
then  the  required  equation  is 


.2  _ 


x2  +  y* 


(37) 


The  equation  of  a  right  cylinder  is  thus 
independent  of  z.  This  means  that  z  may 
have  any  value  whatever  assigned  to  it. 

Examples. — (1)  Show  that  the  equation  of 
a  right  cone  is  x2  +  y2  -  z2t&n2<p  =  0,  where  <p 
represents  half  the  angle  at  the  apex  of  the 
cone.  Hint.  Origin  of  axes  is  at  apex  of  cone  ; 
let  the  z-axis  coincide  with  the  axis  of  the  cone. 
Find  O'P'A'  on  the  base  of  the  cone  resembling 
OP  A  (Fig.  50).     Hence  show  O'F  =  z  tan  <p.     But  O'P'  =  Jx*  + 

(2)  The  equation  of  a  sphere  is  x2  +  y2  +  z2  =  r2.      Prove  this.     Centre  of 
sphere  at  origin  of  axes.     Take  a  section  across  z-axis.     Find 

O'P'A';  OP'=r;  {OP')2={00')2+(0'P')2,  {0'P')2=x2  +  y2;  (00')2  =  z2;  etc. 

The  subject  will  be  taken  up  again  at  different  stages  of  our 
work. 

§  49.    Periodic  or  Harmonic  Motion. 

Let  P  (Fig.  51)  be  a  point  which  starts  to  move  from  a  position 
of  rest  with  a  uniform 
velocity  on  the  perimeter 
of  a  circle.  Let  xOx',  yOy' 
be  coordinate  axes  about 
the  centre  0.  Let PVP2... 
be  positions  occupied  by  the 
point  after  the  elapse  of 
intervals  of  tv  t2 . . .  From 
Pj  drop  the  perpendicular 
M^-l  on  to  the  rc-axis. 
Remembering  that  if  the 
direction  of  M^,  M2P2 . . . 
be  positive,  that  of  MZPV 
M^P^  is  negative,  and  the 
motion  of  OP  as  P  revolves 


Fig.  51. — Periodic  or  Harmonic  Motion. 


136 


HIGHER  MATHEMATICS. 


49. 


about  the  centre  0  in  the  opposite  direction  to  the  hands  of  a  clock 
is  conventionally  reckoned  positive,  then 


sin  a,  = 


+  OPl 


sin  an 


+  M2P, 
+  OP, 


sin  a,  = 


M,P, 


3-+  OP, 


•:  sm  a/1  = 


-MAPA 


+  0PA' 
sina4=  -M4P±. 


Or,  if  the  circle  have  unit  radius  r  =  1, 
sin  a2  =  +  i^ ;  sin  a2  =  +  2lf2P2 ;  sin  a3  =  -  M3P. 

If  the  point  continues  in  motion  after  the  first  revolution,  this 
series  of  changes,  is  repeated  over  and  over  again.  During  the 
first  revolution,  if  we  put  tt-  =  180°,  and  let  6V  02, .  . .  represent 
certain  angles  described  in  the  respective  quadrants, 


=  7r  +  a. 


=    2tT    - 


During  the  second  revolution, 

01  =  2ir  +  a1;    02  =  2tt  +  (tt  -  a2)  J    0,  =  2tt  +  (tt  +  a3),  etc. 

We  may  now  plot  the  curve 


a) 


y  =  sm  a 

by  giving  a  series  of  values  0,  Jir,  f  tt  . . .  to  a  and  finding  the  cor- 
responding values  of  y.     Thus  if 


#  =  a  =  0, 

y  =  sin  0, 


sin  fir, 


sm  7r, 


2"> 

sin 


3^- 

2^, 


2-, 

sin  27r, 


sm  -g-ir, 


2/  =  sin  0°,  sin  90°,  sin  180°,  sin  270°,  sin  360°,  sin  90°, . . . ; 

2/  =  0,  1,  0,  -1,         0,  1,... 

3rmediate  values  are  sin  J?r  =  sin  45°  =  -707,  sin  f  ?r  =  -707  . . . 
The  curve  so  obtained  has  the  wavy  or  undulatory  appearance 


+y . 

/        \                               /        v                               /        V 

f            y                             £_          ^                             t_        _5 

2          t              ^v          t              v!          t 

£       tt      3         \tt    yJ      K     a\       Ttt    *//     a7r    5A       *-«*«>• 

U          2          "  C^         <?            /           2           3"V^        2"      *"4  2         J     'P                     T 

1           7.  «               \           7_                [5           ' 

C       i                     C       j     .                 c      ^7 

>          r                                      \          /                                     \        .r 

<;>                                         ^^                                        >=5^ 

-jr 

Fig.  52.— Curve  of  Sines,  or  Harmonic  Curve. 

shown  in  Fig.  52.     It  is  called  the  curve  of  sines  or  the  har- 
monic curve. 

A  function  whose  value   recurs   at  fixed  intervals  when   the 
variable  uniformly  increases  in  magnitude  is  said  to  be  a  periodic 


§  49.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  137 

function.     Its  mathematical  expression  is 

f(t)=f(t  +  qt)        ...  (2) 

where  q  may  be  any  positive  or  negative  integer.  In  the  present 
case  q  =  2-rr.  The  motion  of  the  point  P  is  said  to  be  a  simple 
harmonic  motion.  Equation  (1)  thus  represents  a  simple  harmonic 
motion. 

If  we  are  given  a  particular  value  of  a  periodic  function  of, 
say,  t,  we  can  find  an  unlimited  number  of  different  values  of  t 
which  satisfy  the  original  function.  Thus  2t,  3t,  ££7.'.,  all  satisfy 
equation  (2). 

Examples.— (1)  Show  that  the  graph  of  y  =  cos  o  has  the  same  form  as 
the  sine  curve  and  would  be  identical  with  it  if  the  y-axis  of  the  sine  curve 
were  shifted  a  distance  of  £»r  to  the  right.  [Proof :  sin  (£tt  +  x)  =  cos  x,  etc.] 
The  physical  meaning  of  this  is  that  a  point  moving  round  the  perimeter  of 
the  circle  according  to  the  equation  y  =  cos  a  is  just  £ir,  or  90°  in  advance  of 
one  moving  according  to  y  =  sin  a. 

(2)  Illustrate  graphically  the  periodicity  of  the  function  y  =  tan  a.  (Note 
the  passage  through  +'  oo.)     Keep  your  graph  for  reference  later  on. 

Instead  of  taking  a  circle  of  unit  radius,  let  r  denote  the  mag- 
nitude of  the  radius,  then 

y  =  r  sin  a (3) 

Since  sin  a  can  never  exceed  the  limits  +  1,  the  greatest  and  least 
values  y  can  assume  are  -  r  and  +  r ;  r  is  called  the  amplitude 
of  the  curve.  The  velocity  of  the  motion  of  P  determines  the  rate 
at  which  the  angle  a  is  described  by  OP,  the  so-called  angular 
velocity.     Let  t  denote  the  time,  q  the  angular  velocity, 

■^  =  q  ;  or  a  =  qt,  .  .  .  (4) 

and  the  time  required  for  a  complete  revolution  is 

t  =  27r/g,  .         .  .         (5) 

which  is  called  the  period  of  oscillation,  the  periodic  value,  or 
the  periodic  time  ;  2-rr  is  the  wave  length.  If  E  (Fig.  51)  denotes 
some  arbitrary  fixed  point  such  that  the  periodic  time  is  counted 
from  the  instant  P  passes  through  E,  the  angle  xOE  =  e,  is  called 
the  epoch  or  phase  constant,  and  the  angle  described  by  OP  in 
the  time  t  =  qt  +  e  =  a,  or 

y  =  r  sin  [qt  +  c). .  .  .  .  (6) 
Electrical  engineers  call  c  the  "lead"  or,  if  negative,  the  "lag" 
of  the  electric  current. 


138  HIGHEB  MATHEMATICS.  §  49. 

Examples. — (1)  Plot  (6),  note  that  the  angles  are  to  be  measured  in 
radians  (page  606),  and  that  one  radian  is  57'3°.  Now  let  r=10,  e  =  30°  =  O52 
radians.     Let  q  denote  0-5°,  or  ^Ve  radians. 

.-.  y  =  10  sin  (0-0087*  +  0-52). 
If  t  =  10,  y  =  10  sin  0-61  =  10  sin  35°,  from  a  Table  of  Radians  (Table   XIII.) . 
From  a  Table  of  Trigonometrical  Sines,  10  sin  35°  =  10  x  0*576  =  5-76  we  get 
the  same  result  more  directly  by  working  in  degrees.     In  this  case, 

y  =  10  sin  (£*°  +  30°). 
If  £=10,  we  have  y =10  sin  35°  as  before.     Then  we  find  if  r=10,  t= 30 =0-52 
radians,  and  if 

t  =  0,        120,        300,        480,        720 ; 
y  =  5,  10,  0,       - 10,         -  5. 

Intermediate  values  are  obtained  in  the  same  way.  The  curve  is  shown  in 
Fig.  53.  Now  try  the  effect  of  altering  the  value  of 
e  upon  the  value  of  y,  say,  you  put  e  =  0,  45°,  60°,  90°, 
and  note  the  effect  on  Oy  (Fig.  52). 

(2)  It  is  easy  to  show  that  the  function 

a  sin  (qt  +  e)  +  b  cos  (qt  +  e)         .         (7) 
is  equal  to  A  sin  (qt  +  ex)  by  expanding  (7)  as  indi- 
um    go  cated  in  formulae  (23)  and  (24),  page  612.     Thus  we 
get 
sin  qt(a  cos  e  -  b  sin  e)  +  cos  qt(b  cos  e  +  a  sin  «)  =  A  sin  (qt  +  ej, 
provided  we  collect  the  constant  terms  as  indicated  below. 

A  cos  ex  =  a  cos  e  -  b  sin  e  ;  A  sin  e2  =  b  cos  e  +  a  sin  e.        .        (8) 
Square  equations  (8)  and  add 

.-.  42  =  <z2+  bV (9) 

Divide  equations  (8),  rearrange  terms  and  show  that 

■fr  ['-*>  =  to  («-„)  —  | .        .        .        .      (10) 
cos  (e  -  cj)  a  v    ' 

(3)  Draw  the  graphs  of  the  two  curves, 

y  =  a  sin  (qt  +  e) ;  and  yl=a1  sin  (qt  +  cj. 
Compare  the  result  with  the  graph  of 

y2  =  a  sin  (qt  +  e)  +  a^  sin  (qt  +  ej. 

(4)  Draw  the  graphs  of 

yx  =  sin  x  ;  y2  =  £  sin  Bx ;  ys  =  £  sin  5x  ;  y  =  sin  x  +  |  sin  3x  +  ±  sin  5x. 

(5)  There  is  an  interesting  relation  between  sin  x  and  ex.  Thus,  show 
that  if 

y  =  a  sin  qt  +  b  sin  qt ;  -^  =  -  q*y  ;  ^  =  gfy  ; . . . 

The  motion  of  .M"  (Fig.  51),  that  is  to  say,  the  projection  of  the 
moving  point  on  the  diameter  of  the  circle  xOx'  is  a  good  illustra- 
tion of  the  periodic  motion  discussed  in  §21.     The  motion  of  an 


§50. 


COORDINATE  OR  ANALYTICAL  GEOMETRY. 


139 


oscillating  pendulum,  of  a  galvanometer  needle,  of  a  tuning  fork, 
the  up  and  down  motion  of  a  water  wave,  the  alternating  electric 
current,  sound,  light,  and  electromagnetic  waves  are  all  periodic 
motions.  Many  of  the  properties  of  the  chemical  elements  are  also 
periodic  functions  of  their  atomic  weights  (Newlands-Mendeleeff 
law). 

Some  interesting  phenomena  have  recently  come  to  light  which 
indicate  that  chemical  action  may  assume  a  periodic  character. 
The  evolution  of  hydrogen  gas,  when  hydrochloric .  acid  acts  on 
one  of  the  allotropic  forms  of  chromium,  has  recently  been  studied 
by  W.  Ostwald  (Zeit. 
phys.  Chem.,  35,  33, 
204, 1900).  He  found 
that  if"  the  rate  of  evo- 
lution of  gas  evolved 
during  the  action  be 
plotted  as  ordinate 
against  the  time  as 
abscissa,  a  curve  is 
obtained  which  shows  regularly  alternating  periods  of  slow  and 
rapid  evolution  of  hydrogen.  The  particular  form  of  these  "  waves  " 
varies  with  the  conditions  of  the  experiment.  One  of  Ostwald' s 
curves  is  shown  in  Fig.  54  (see  J.  W.  Mellor's  Chemical  Statics 
and  Dynamics,  London,  348,  1904). 


Ostwald's  Curve  of  Chemical  Action. 


§  50.    Generalized  Forces  and  Coordinates. 

When  a  mass  of  any  substance  is  subject  to  some  physical 
change,  certain  properties — mass,  chemical  composition — remain 
fixed  and  invariable,  while  other  properties — temperature,  pressure, 
volume — vary.  When  the  value  these  variables  assume  in  any 
given  condition  of  the  substance  is  known,  we  are  said  to  have  a 
complete  knowledge  of  the  state  of  the  system.  These  variable 
properties  are  not  necessarily  independent  of  one  another.  We 
have  just  seen,  for  instance,  that  if  two  of  the  three  variables 
defining  the  state  of  a  perfect  gas  are  known,  the  third  variable 
can  be  determined  from  the  equation 

pv  =  RT, 
where  B  is  a  constant.     In  such  a  case  as  this,  the  third  variable 
is  said  to  be  a  dependent  variable,  the  other  two,  independent  vari- 


140  HIGHER  MATHEMATICS.  §  50. 

ables.  When  the  state  of  any  material  system  can  be  denned  in 
terms  of  n  independent  variables,  the  system  is  said  to  possess  n 
degrees  of  freedom,  and  the  n  independent  variables  are  called 
generalized  coordinates.  For  the  system  just  considered  n  =  2, 
and  the  system  possesses  two  degrees  of  freedom. 

Again,  in  order  that  we  may  possess  a  knowledge  of  some 
systems,  say  gaseous  nitrogen  peroxide,  not  only  must  the  vari- 
ables given  by  the  gas  equation 

<f>(p,  v,T)  =  0 

be  known,  but  also  the  mass  of  the  N204  and  of  the  N02  present. 
If  these  masses  be  respectively  m1  and  ra2,  there  are  five  variables 
to  be  considered,  namely, 

^(p,  v,  T,  mv  m2)  =  0, 

but  these  are  not  all  independent.     The  pressure,  for  instance,  may 
be  fixed  by  assigning  values  to  v,  T,  mv  m2 ;  p  is  thus  a  dependent 
variable,  v,  T,  mlf  m2  are  independent  variables.     Thus 
p  =  f(vt  T}  mv  m2). 

We  know  that  the  dissociation  of  N204  into  2N02  depends  on  the 
volume,  temperature  and  amount  of  N02  present  in  the  system 
under  consideration.     At  ordinary  temperatures 

and  the  number  of  independent  variables  is  reduced  to  three.  In 
this  case  the  system  is  said  to  possess  three  degrees  of  freedom. 
At  temperatures  over  135° — 138°  the  system  contains  N02  alone, 
and  behaves  as  a  perfect  gas  with  two  degrees  of  freedom. 

In  general,  if  a  system  contains  m  dependent  and  n  independent 
variables,  say 

fl/j,  X2t  #3>  •  •  •  %n  +  m 

variables,  the  state  of  the  system  can  be  determined  by  m  +  n 
equations.  As  in  the  familiar  condition  for  the  solution  of  simul- 
taneous equations  in  algebra,  n  independent  equations  are  required 
for  finding  the  value  of  n  unknown  quantities.  But  the  state  of 
the  system  is  defined  by  the  m  dependent  variables  ;  the  remaining 
n  independent  variables  can  therefore  be  determined  from  n  inde- 
pendent equations. 

Let  a  given  system  with  n  degrees  of  freedom  be  subject  to 
external  forces 

Xv  X2,  X3, . . .  Xnt 


§  50.        COORDINATE  OR  ANALYTICAL  GEOMETRY.  141 

so  that  no  energy  enters  or  leaves  the  system  except  in  the  form 
of  heat  or  work,  and  such  that  the  n  independent  variables  are 
displaced  by  amounts 

dxv  dx2,  dxz, . . .  dxn. 
Since  the  amount  of  work  done  on  or  by  a  system  is  measured  by 
the  product  of  the  force  and  the  displacement,  these  external  forces 
XYX2 ...  perform  a  quantity  of  work  dW  which  depends  on  the 
nature  of  the  transformation.     Hence 

dW  =  X1dx1  -f  X2dx2  + . . .  Xndxn 
where  the  coefficients  Xv  X2,  X3...are  called  the  generalized 
forces  acting  on  the  system.  P.  Duhem,  in  his  work,  Traite  $U- 
mentaire  de  Micanique  Ghimique  fondee  sur  la  Thermodynamique, 
Paris,  1897-99,  makes  use  of  generalized  forces  and  generalized 
coordinates. 


CHAPTER  III. 

FUNCTIONS  WITH  SINGULAR  PROPERTIES. 

"  Although  a  physical  law  may  never  admit  of  a  perfectly  abrupt 
change,  there  is  no  limit  to  the  approach  which  it  may  make 
to  abruptness." — W.  Stanley  Jevons. 

§  51.    Continuous  and  Discontinuous  Functions. 

The  law  of  continuity  affirms  that  no  change  can  .  take  place 
abruptly.  The  conception  involved  will  have  been  familiar  to  the 
reader  from  the  second  section  of  this  work.  It  was  there 
shown  that  the  amount  of  substance,  x,  transformed  in  a  chemical 
reaction  in  a  given  time  becomes  smaller  as  the  interval  of  time,  t, 
during  which  the  change  occurs,  is  diminished,  until  finally,  when 
the  interval  of  time  approaches  zero,  the  amount  of  substance 
transformed  also  approaches  zero.  In  such  a  case  x  is  not  only  a 
function  of  t,  but  it  is  a  continuous  function  of  t.  The  course  of 
such  a  reaction  may  be  represented  by  the  motion  of  a  point  along 
the  curve 

If  the  two  states  of  a  substance  subjected  to  the  influence  of  two 
different  conditions  of  temperature  be  represented,  say,  by  two 
neighbouring  points  on  a  plane,  the  principle  of  continuity  affirms 
that  the  state  of  the  substance  at  any  intermediate  temperature 
will  be  represented  by  a  point  lying  between  the  two  points  just 
mentioned ;  and  in  order  that  the  moving  point  may  pass  from  one 
point,  a,  on  the  curve  to  another  point,  b,  on  the  same  curve,  it 
must  successively  assume  all  values  intermediate  between  a  and  b, 
and  never  move  off  the  curve.  This  is  a  characteristic  property  of 
continuous  functions.  Several  examples  have  been  considered  in 
the  preceding  chapters.  Most  natural  processes,  perhaps  all,  can 
be  represented  by  continuous  functions.  Hence  the  old  empiricism  : 
Natura  non  agit  per  saltum. 

The  law  of  continuity,  though  tacitly  implied  up  to  the  present, 

142 


§52. 


FUNCTIONS  WITH  SINGULAR  PROPERTIES. 


143 


does  not  appear  to  be  always  true.  Even  in  some  of  the  simplest 
phenomena  exceptions  seem  to  arise.  In  a  general  way,  we  can 
divide  discontinuous  functions  into  two  classes  :  first,  those  in  which 
the  graph  of  the  function  suddenly  stops  to  reappear  in  some  other 
part  of  the  plane — in  other  words  a  break  occurs  ;  second,  those 
in  which  the  graph  suddenly  changes  its  direction  without  exhibit- 
ing a  break 1 — in  that  case  a  turning  point  or  point  of  inflexion 
appears. 

Other  kinds  of  discontinuity  may  occur,  but  do  not  commonly 
arise  in  physical  work.  For  example,  a  function  is  said  to  be  dis- 
continuous when  the  value  of  the  function  y  =  f(x)  becomes 
infinite  for  some  particular  value  of  x.  Such  a  discontinuity 
occurs  when  x  =  0  in  the  expression  y  =  ljx.  The  differential 
coefficient  of  this  expression, 

^  =  -1, 
dx  x2' 

is  also  discontinuous  for  x  =  0.     Other  examples,  which  should  be 

verified  by  the  reader  are,  log  x,  when  x  =  0 ;  tan  x,  when  x  =  \tt,  ... 

The  graph  for  Boyle's  equation,  pv  =  constant,  is  also  said  to  be 

discontinuous  at  an  infinite  distance  along  both  axes. 

§  52.    Discontinuity  accompanied  by  "Breaks". 

If  a  cold  solid  be  exposed  to  a  source  of  heat,  heat  appears 
to  be  absorbed,  and  the 
temperature,  0,  of  the 
solid  is  a  function  of 
the  amount  of  heat,  Q, 
apparently  absorbed  by 
the  solid.  As  soon 
as  the  solid  begins  to 
melt,  it  absorbs  a  great 
amount  of  heat  (latent 
heat  of  fusion),  unac- 
companied by  any  rise  of 
temperature.  When  the 
substance  has  assumed  the  fluid  state  of  aggregation,  the  tem- 

1  Sometimes  the  word  "break"  is  used  indiscriminately  for  both  kinds  of 
discontinuity.  It  is,  indeed,  questionable  if  ever  the  "break"  is  real  in  natural 
phenomena.  I  suppose  we  ought  to  call  turning  points  "singularities,"  not 
"discontinuities"  (see  S.  Jevon's  Principles  qf  Science,  London,  1877). 


y            $/* 

;  \  boiling  point 

U^c 

,<•' TJ  melting  point 

'R          0 

Vr 

Pig.  55. 


144  HIGHER  MATHEMATICS.  §  52. 

perature  is  a  function  of  the  amount  of  heat  absorbed  by  the 
fluid,  until,  at  the  boiling  point,  similar  phenomena  recur.  Heat 
is  absorbed  unaccompanied  by  any  rise  of  temperature  (latent  heat 
of  vaporization)  until  the  liquid  is  completely  vaporized.  The 
phenomena  are  illustrated  graphically  by  the  curve  0 ABODE  (Eig. 
55).  If  the  quantity  of  heat,  Q,  supplied  be  regarded  as  a  function 
of  the  temperature,  6,  the  equation  of  the  curve  OABGED  (Fig. 
55),  will  be 

This  function  is  said  to  be  discontinuous  between  the  points  A  and 
B,  and  between  G  and  D.  Breaks  occur  in  these  positions.  f(6) 
is  accordingly  said  to  be  a  discontinuous  function,  for,  if  a 
small  quantity  of  heat  be  added  to  a  substance,  whose  state  is 
represented  by  a  point,  between  A  and  i?,  or  G  and  D,  the  tem- 
perature is  not  affected  in  a  perceptible  manner.  The  geometrical 
signification  of  the  phenomena  is  as  follows :  There  are  two 
generally  different,  tangents  to  the  curve  at  the  points  A  and  B 
corresponding  to  the  one  abscissa,  namely,  tan  a  and  tan  a.  In 
other  words,  see  page  102,  we  have 

dQ 

■jq  =f{0)  =  tan  a  =  tan  angle  6BA  ; 

dQ 

~Tq  =  f(0)  =  tan  a'  =  tan  angle  6B'A,\ 

that  is  to  say,  the  function  f'(0)  is  discontinuous  because  the 
differential  coefficient  has  two  distinct  values  determined  by  the 
slope  of  the  tangent  to  each  curve  at  the  point  where  the  discon- 
tinuity occurs. 

The  physical  meaning  of  the  discontinuity  in  this  example,  is 
that  the  substance  may  have  two  values  for  its  specific  heat — the 
amount  of  heat  required  to  raise  the  temperature  of  one  gram  of 
the  solid  one  degree — at  the  melting  point,  the  one  corresponding 
to  the  solid  and  the  other  to  the  liquid  state  of  aggregation.  The 
tangent  of  the  angle  represented  by  the  ratio  dQ/dO  obviously 
represents  the  specific  heat  of  the  substance.  An  analogous  set  of 
changes  occurs  at  the  boiling  point. 

It  is  necessary  to  point  out  that  the  alleged  discontinuity  in 
the  curve  OABG  may  be  only  apparent.  The  "  corners  "  may  be 
rounded  off.     It  would  perhaps  be  more  correct  to  say  that  the 


§53. 


FUNCTIONS  WITH  SINGULAR  PROPERTIES. 


145 


curve  is  really  continuous  between  A  and  B,  but  that  the  change 
of  temperature  with  the  addition  of 
heat  is  discontinuous. 

Again,  Fig.  56  shows  the  result  of 
plotting  the  variations  in  the  volume 
of  phosphorus  with  temperatures  in 
the  neighbourhood  of  its  melting  point. 
AB  represents  the  expansion  curve  of 
the  solid,  CD  that  of  the  liquid.  A 
break  occurs  between  B  and  G.  Phos- 
phorus at  its  melting  point  may  thus 
have  two  distinct  coefficients  of  ex- 
pansion, the  one  corresponding  to  the  solid  and  the  other  to  the 
liquid  state  of  aggregation.  Similar  changes  take  place  during  the 
passage  of  a  system  from  one  state  to  another,  say  of  rhombic  to 
monoclinic  sulphur ;  of  a  mixture  of  magnesium  and  sodium  sul- 
phates to  astracanite,  etc.  The  temperature  at  which  this  change 
occurs  is  called  the  "  transition  point ". 


Fig.  56. 


§  53.   The  Existence  of  Hydrates  in  Solution. 

Another  illustration.  If  p  denotes  the  percentage  composition 
of  an  aqueous  solution  of  ethyl  alcohol  and  s  the  corresponding 
specific  gravity  in  vacuo  at  15°  (sp.  gr.  H20  at  15°  =  9991*6),  we 
have  the  following  table  compiled  by  Mendeleeff : — 


p 

s 

P 

s 

P 

s 

P 

s 

5 

9904-1 

30 

9570-2 

55 

9067-4 

80 

8479-8 

10 

9831-2 

35 

9484-5 

60 

8953-8 

85 

8354-8 

15 

9768-4 

40 

9389-6 

65 

8838-6 

90 

82250 

20 

9707-9 

45 

9287-8 

70 

8714-5 

95 

8086-9 

25 

9644-3 

50 

9179-0 

75 

8601-4 

100 

7936-6 

It  was  found  empirically  that  the  experimental  results  are  fairly 
well  represented  by  the  equation 

s  =  a  +  bp  +  cp2,  ...         (1) 

which  is  the  general  expression  for  a  parabolic  curve,  a,  b  and  c 
being  constants,  page  99,  or  the  equation  may  embody  two  straight 
lines,  page  121.  By  plotting  the  experimental  data  the  curve  shown 
in  Fig.  57  is  obtained. 

It  is  urged  that  just  as  compounds  may  be  formed  and  decom- 

K 


146 


HIGHER  MATHEMATICS. 


§53. 


posed  at  temperatures  higher  than  that  at  which  their  dissociation 
commences,  and  that  for  any  given  temperature  a  definite  relation 
exists  between  the  amounts  of  the  original  compound  and  of  the 
products  of  its  dissociation,  so  may  definite  but  unstable  hydrates 
exist  in  solutions  at  temperatures  above  their  dissociation  tempera- 
ture. If  the  dissolved  substance  really  enters  into  combination 
with  the  solvent  to  form  different  compounds  according  to  the 
nature  of  the  solution,  many  of  the  physical  properties  of  the 
solution — density,  thermal  conductivity  and  such  like — will  natur- 
ally depend  on  the  amount  and  nature  of  these  compounds, 
because  chemical  combination  is  usually  accompanied  by  volume, 
density,  thermal  and  other  changes. 


fo.ooo 

9.000 

8.0O0 

0 

'0 

4i 

0 

3 

O 

6 

0 

WO 

Fig.  57. 
Assuming  that  the  amount  of  such  a  definite  compound  is  pro- 
portional to  the  concentration  of  the  solution,  the  rate  of  change  of, 
say,  the  density,  s,  with  change  of  concentration,  p,  will  be  a  linear 
function  of  p,  in  other  words,  ds/dp  will  be  represented  by  the  equa- 
tion for  a  straight  line.     From  the  differentiation  of  (1),  we  obtain, 

•         •         •         (2) 


*-»  +  *» 


where  ds  is  the  difference  in  the  density  of  two  experimental 
values  corresponding  with  a  difference  dp  in  the  percentage  com- 
position of  the  two  solutions.  The  second  member  of  (2)  cor- 
responds with  the  equation  of  a  straight  line,  page  90.  On  treating 
the  experimental  data  by  this  method,  Mendeleeff l  found  that  ds/dp 

1  D.  Mendeleeff,  Journ.  Ohem.  Soc,  31,  778,  1887  ;  S.  U.  Pickering,  ib.,  57,  64, 
331,  1890;  Phil.  Mag.  [5],  29,427,  1890;  Watt's  Diet.  Chem.,  art.  " Solutions"  ii., 
1894 ;  H.  Crompton,  Journ.  Chem.  Soc,  53,  116,  1888  ;  S.  Arrhenius,  Phil.  Mag.  [5], 
28,  36,  1889 ;  E.  H.  Hayes,  ib.  [5],  32,  99,  1891 ;  A.  W.  Riicker,  ib.  [5],  32,  306,  1891 ; 
S.  Lupton,  ib.  [5],  31,  418,  1891 ;  T.  M.  Lowry,  Science  Progress,  3, 124,  1908. 


§  53.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  147 


was  discontinuous, 
ordinates  against  gy 
abscissa    p    for  dj* 
concentrations 
corresponding  to 
17-56,  46-00  and 
88-46  per  cent, 
of  ethyl  alcohol. 
These  concen- 
trations coincide 


Breaks  were  obtained  by  plotting  dsjdp  as 


wo 


After  Mendeleeff. 


with  chemical  compounds  having  the  composition  C2H5OH  .  12H20, 
C2H5OH  .  3H20  and  3C2H6OH .  H20  as  shown  in  Fig.  58.  The 
curves  between  the  breaks  are  supposed  to  represent  the  "zone" 
in  which  the  corresponding  hydrates  are  present  in  the  solution. 

The  mathematical  argument  is  that  the  differential  coefficient 
of  a  continuous  curve  will  differentiate  into  a  straight  line  or 
another  continuous  curve  ;  while  if  a  curve  is  really  discontinuous, 
or  made  up  of  a  number  of  different  curvea,  it  will  yield  a  series  of 
straight  lines.  Each  line  represents  the  rate  of  change  of  the 
particular  physical  property  under  investigation  with  the  amount 
of  hypothetical  unstable  compound  existing  in  solution  at  that 
concentration.  An  abrupt  change  in  the  direction  of  the  curve 
leads  to  a  breaking  up  of  the  first  differential  coefficient  of  that 
curve  into  two  curves  which  do  not  meet.  This  argument  has  been 
extensively  used  by  Pickering  in  the  treatment  of  an  elaborate  and 
painstaking  series  of  determinations  of  the  physical  properties  of 
solutions.  Crompton  found  that  if  the  electrical  conductivity  of  a 
solution  is  regarded  as  a  function  of  its  percentage  composition, 
such  that 

K  =  a  +  bp  +  cp2  +  fpz,     .        .        .         (3) 

the  first  differential  coefficient  gives  a  parabolic  curve  of  the  type 
of  (1)  above,  while  the  second  differential  coefficient,  instead  of 
being  a  continuous  function  of  p, 


dtf 


=  A+Bp, 


W 


was  found  to  consist  of  a  series  of  straight  lines,  the  position  of  the 
breaks  being  identical  with  those  obtained  from  the  first  differential 
coefficient  dsjdp.  The  values  of  the  constants  A  and  B  are  readily 
obtained  if  c  and  p  are  known.     If  the  slope  of  the  (p,  s)-curve 


K 


148 


HIGHER  MATHEMATICS. 


§54. 


changes  abruptly,  ds/dp  is  discontinuous  ;  if  the  slope  of  the 
(ds/dp,  p)-Guxve  changes  abruptly,  dh/dp2  is  discontinuous. 

But  after  all  we  are  only  working  with  empirical  formulae,  and 
"no  juggling  with  feeble  empirical  expressions,  and  no  appeal  to 
the  mysteries  of  elementary  mathematics  can  legitimately  make  ex- 
perimental results  any  more  really  discontinuous  than  they  them- 
selves are  able  to  declare  themselves  to  be  when  properly  plotted  ".1 

It  must  be  pointed  out  that  the  differentiation  of  experimental 
results  very  often  furnishes  quantities  of  the  same  order  of  magni- 
tude as  the  experimental  errors  themselves.2  This  is  a  very 
serious  objection.  Pickering  has  tried  to  eliminate  the  experi- 
mental errors,  to  some  extent,  by  differentiating  the  results  obtained 
by  "smoothing"  the  curve  obtained  by  plotting  the  experimental 
results.  On  the  face  of  it  this  "  smoothing "  of  experimental 
results  is  a  dangerous  operation  even  in  the  hands  of  the  most 
experienced  workers.  Indeed,  it  is  supposed  that  that  prince  of 
experimenters,  Regnault,  overlooked  an  important  phenomenon  in 
applying  this  very  smoothing  process  to  his  observations  on  the 
vapour  pressure  of  saturated  steam.  Regnault  supposed  that  the 
curve  OPQ  (Fig.  64)  showed  no  singular  point  at  P  (Fig.  64)  when 
water  passed  from  the  liquid  to  the  solid  state  at  0°.  It  was  re- 
served for  J.  Thomson  to  prove  that  the  ice-steam  curve  has  a 
different  slope  from  the  water-steam  curve. 


§  5$.  The  Smoothing  of  Curves. 

The  results  of  observations  of  a  series  of  corresponding  changes 
in  two  variables  are  represented  by  light 
dots  on  a  sheet  of  squared  paper.  The 
dots  in  Fig.  59  represent  the  vapour 
pressures  of  dissociating  ammonium 
carbonate  at  different  temperatures.  A 
curve  is  drawn  to  pass  as  nearly  as  pos- 
sible through  all  these  points.  The  re- 
sulting curve  is  assumed  to  be  a  graphic 
representation  of  the  general  formula 
(known  or  unknown)  connecting  the  two  variables.     Points  devi- 


\» 

\  • 

V 

> 

• 

J 

•    • 

• 

Fig.  59. — Smoothed  Curve. 


i  0.  J.  Lodge,  Nature,  40,  273,  1889  ;  S.  U.  Pickering,  ib.,  40,  343,  1889. 
2  This   paragraph,  will  be  better  understood  after  Chapter  V.,  §  106,  has  been 
studied.     The  reader  may  then  return  to  this  section. 


§  55.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  149 

ating  from  the  curve  are  assumed  to  be  affected  with  errors  of 
observation.  As  a  general  rule  the  curve  with  the  least  curvature 
is  chosen  to  pass  through  or  within  a  short  distance  of  the  greatest 
number  of  dots,  so  that  an  equal  number  of  these  dots  (representing 
experimental  observations)  lies  on  each  side  of  the  curve.  Such 
a  curve  is  said  to  be  a  smoothed  curve  (see  also  page  320). 

One  of  the  commonest  methods  of  smoothing  a  curve  is  to  pin 
down  a  flexible  lath  to  points  through  which  the  curve  is  to  be 
drawn  and  draw  the  pen  along  the  lath.  It  is  found  impossible  in 
practice  to  use  similar  laths  for  all  curves.  The  lath  is  weakest 
where  the  curvature  is  greatest.  The  selection  and  use  of  the  lath 
is  a  matter  of  taste  and  opinion.  The  use  of  "  French  curves  "  is 
still  more  arbitrary.  Pickering  used  a  bent  spring  or  steel  lath  held 
near  its  ends.  Such  a  lath  is  shown  in  statical  works  to  give  a 
line  of  constant  curvature.  The  line  is  called  an  "  elastic  curve  " 
(see  G.  M.  Minchin's  A  Treatise  on  Statics,  Oxford,  2,  204,  1886). 

§  55.   Discontinuity  accompanied  by  a  Sudden  Change  of 

Direction. 

The  vapour  pressure  of  a  solid  increases  continuously  with 
rising  temperature  until,  at  its  melting  point,  the  vapour  pressure 
11  suddenly  "  begins  to  increase  more  rapidly  than  before.  This  is 
shown  graphically  in  Fig.  60.  The  substance  melts  at  the  point  of 
intersection  of  the  "  solid  "  and  u  liquid  " 
curves.  The  vapour  pressure  itself  is  not 
discontinuous.  It  has  the  same  value  at 
the  melting  point  for  both  solid  and  liquid 
states  of  aggregation.  It  is,  however,  quite 
clear  that  the  tangents  of  the  two  curves 
differ  from  each  other  at  the  transition 
point,  because  .-•g]R  /R' 

tan  a  =f(6)  =  ^|  is  less  than  tan  a'  =/(#)  =  %  Fig.  60. 

There  are  two  tangents  to  the  _p#-curve  at  the  transition  point. 
The  value  of  dp/dd  for  solid  benzene,  for  example,  is  greater  than 
for  the  liquid.     The  numbers  are  2-48  and  1*98  respectively. 

If  the  equations  of  the  two  curves  were  respectively  ax  +  by  =  1 ; 
and  bx  +  ay  =  1,  the  roots  of  these  two  equations, 

1  1 

x  =  r  ;  y  =  r> 

a  +  b     v       a  +  b 


"<x\ &\         0 


150 


HIGHER  MATHEMATICS. 


§55. 


Fig.  61. 


would  represent  the  coordinates  of  the  point  of  intersection,  as 
indicated  on  page  94.  To  illustrate  this  kind  of  discontinuity  we 
shall  examine  the  following  phenomena  : — 

I.  Critical  temperature.  Cailletet  and  Collardeau  have  an 
ingenious  method  for  finding  the  critical  temperature  of  a 
substance  without  seeing  the  liquid.1  By  plotting  temperatures 
as  abscissae  against  the  vapour  pressures  of  different  weights  of 
the  same  substance  heated  at  constant  volume,  a  series  of  curves 
are  obtained  which  are  coincident  as  long  as  part  of  the  substance 
is  liquid,  for  "the  pressure  exerted  by  a  saturated  vapour  depends 

on  temperature  only  and  is  independent  of  the 
quantity  of  liquid  with  which  it  is  in  contact ". 
Above  the  critical  temperature  the  different  masses 
of  the  substance  occupying  the  same  volume  give 
different  pressures.  From  this  point  upwards  the 
pressure-temperature  curves  are  no  longer  super- 
posable.  A  series  of  curves  are  thus  obtained 
which  coincide  at  a  certain  point  P  (Fig.  61),  the  abscissa,  OK, 
of  which  denotes  the  critical  temperature.  As  before,  the  tangent 
of  each  curve  Pa,  Pb  . .  .  is  different  from  that  of  OP  at  the  point  P. 

II.  Cooling  curves.  If  the  temperature  of  cooling  of  pure  liquid 
bismuth  be  plotted  against  time,  the  resulting  curve,  called  a  cooling 
curve  (ab,  Fig.  62),  is  continuous,  but  the  moment  a  part  of  the 

metal  solidifies,  the  curve  will  take 
another  direction  be,  and  continue 
so  until  all  the  metal  is  solidified, 
when  the  direction  of  the  curve 
again  changes,  and  then  continues 
quite  regularly  along  cd.  For  bis- 
muth the  point  b  is  at  268°. 

If  the  cooling  curve  of  an  alloy 
of  bismuth,  lead  and  tin  (Bi,  21 ; 
timt   Pb,    5*5 ;    Sn,    75*5)    is    similarly 
plotted,  the  first  change  of  direction 
Pic.  62._Oooling  Carves.  h  observed  at  175°,  when  solid  bis- 

muth  is  deposited ;  at  125°  the  curve  again  changes  its  direction, 


& 

\ 

ZSO' 

\- 

""X 

ZOO" 

%*v 

ISO' 

v^^ 

WO" 

t 

0 

i  L.  P.  Cailletet  and  E.  Collardeau,  Ann.  Chim.  Phys.,  [6],  25,  522,  1891.  Note 
lhat  the  critical  temperature  is  the  temperature  above  which  a  substance  cannot  exist 
other  than  in  the  gaseous  state. 


§56. 


FUNCTIONS  WITH  SINGULAR  PROPERTIES. 


151 


with  a  simultaneous  deposition  of  solid  bismuth  and  tin ;  and 
finally  at  96°  another  change  occurs  corresponding  to  the  solidifi- 
cation of  the  eutectic  alloy  of  these  three  metals. 

These  cooling  curves  are  of  great  importance  in  investigations 
on  the  constitution  of  metals  and  alloys.  The  cooling  curve  of  iron 
from  a  white  heat  is  particularly  interesting,  and  has  given  rise  to 
much  discussion.  The  curve  shows  changes  of  direction  at  about 
1,130°,  at  about  850°  (called  Arz  critical  point),  at  about  770° 
(called  Ar2  critical  point),  at 
about  500°  (called  the  ArY  critical 
point),  at  about  450°— 500°  C, 
and  at  about  400°  C.  The  mag- 
nitude of  these  changes  varies 
according  to  the  purity  of  the 
iron.  Some  are  very  marked 
even  with  the  purest  iron.  This 
sudden  evolution  of  heat  (recal- 
escence)  at  different  points  of  the 
cooling  curve  has  led  many  to 
believe  that  iron  exists  in  some 
allotropic  state  in  the  neighbourhood  of  these  temperatures.1  Fig. 
63  shows  part  of  a  cooling  curve  of  iron  in  the  most  interesting 
region,  namely,  the  Ar3  and  Ar2  critical  points. 


n/ne 


Fig.  63. — Diagrammatic. 


§  56.  The  Triple  Point. 

Another  example,  which  is  also  a  good  illustration  of  the  beauty 
and  comprehensive  nature  of  the  graphic  method 
of  representing  natural  processes,  may  be  given 
here. 

(a)  When  water,  partly  liquid,  partly  vapour, 
is  enclosed  in  a  vessel,  the  relation  between  the 
pressure  and  the  temperature  can  be  represented 
by  the  curve  PQ  (Fig.  64),  which  gives  the 
pressure  corresponding  with  any  given  tempera- 
ture when  the  liquid  and  vapour  are  in  contact  and  in  equilibrium. 
This  curve  is  called  the  steam  line. 

(b)  In  the  same  way  if  the  enclosure  were  filled  with  solid  ice, 


Fig.  64.— Tripk 
Point. 


1  W.  C.  Roberts- Austen's  papers  in  the  Proc.  Soc.  Mechanical  Engineers,  543, 
1891  ;  102,  1893  ;  238,  1895  ;  31,  1897  ;  35,  1899,  may  be  consulted  for  fuller  details. 


152  HIGHER  MATHEMATICS.  §  56. 

and  liquid  water,  the  pressure  of  the  mixture  would  be  completely- 
determined  by  the  temperature.  The  relation  between  pressure 
and  temperature  is  represented  by  the  curve  PN,  called  the  ice 
line. 

(c)  Ice  may  be  in  stable  equilibrium  with  its  vapour,  and  we 
can  plot  the  variation  of  the  vapour  pressure  of  ice  with  its  tem- 
perature. The  curve  OP  so  obtained  represents  the  variation  of 
the  vapour  pressure  of  ice  with  temperature.  It  is  called  the  hoar 
frost  line. 

The  plane  of  the  paper  is  thus  divided  into  three  parts  bounded 
by  the  three  curves  OP,  PN,  PQ.  If  a  point  falls  within  one  of 
these  three  parts  of  the  plane,  it  represents  water  in  one  particular 
state  of  aggregation,  ice,  liquid  or  steam.1  When  a  point  falls  on  a 
boundary  line  it  corresponds  with  the  coexistence  of  two  states  of 
aggregation.  Finally,  at  the  point  P,  and  only  at  this  point,  the 
three  states  of  aggregation,  ice,  water,  and  steam  may  coexist  to- 
gether. This  point  is  called  the  triple  point.  For  water  the 
coordinates  of  the  triple  point  are 

p  =  4-58  mm.,  T  =  0-0076°  C. 

1.  Influence  of  pressure  on  the  melting-point  of  a  solid.  The 
two  formulae,  dQ  =  Td<f> ;  Q)Qfdv)T  =  T(bp[dT)vt  were  discussed 
on  pages  81  and  82.  Divide  the  former  by  dv  and  substitute  the 
result  in  the  latter.     We  thus  obtain, 

C!),-(8>;  •  •  ■  ■  » 

which  states  that  the  change  of  entropy,  <£,  per  unit  change  of 
volume,  v,  at  a  constant  temperature  (T°  absolute),  is  equal  to  the 
change  of  pressure  per  unit  change  of  temperature  at  constant 
volume.  If  a  small  amount  of  heat,  dQ,  be  added  to  a  substance 
existing  partly  in  one  state,  "  1,"  and  partly  in  another  state,  "2," 
a  proportional  quantity,  dm,  of  the  mass  changes  its  state,  such 
that 

dQ  =  L12dm, 

where  L12  is  a  constant  representing  the  latent  heat  of  the  change 
from  state  "  1 "  to  state  "  2  ".     From  the  definition  of  entropy,  0, 

1  Certain  unstable  conditions  (metastable  states)  are  known  in  which  a  liquid  may- 
be found  in  the  solid  region.  A  supercooled  liquid,  for  instance,  may  continue  the  QP 
curve  along  to  S  instead  of  changing  its  direction  along  PM, 


§  56.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  153 

dQ  =  Td<f> ;  hence  d<j>  =  -j?dm.      .        .        (2) 

If  vv  v2  be  the  specific  volumes  of  the  substance  in  the  first  and 
second  states  respectively 

dv  =  v2dm  -  vxdm  =  (v2  -  v-^dm. 
From  (2)  and  (1) 

'  '  Wr"  T(v2  -  v,)  >  \7>f)9  T(v2  -  SJ  '  V  ; 
This  last  equation  tells  us  at  once  how  a  change  of  pressure 
will  change  the  temperature  at  which  two  states  of  a  substance 
can  coexist,  provided  that  we  know  vv  v2,  T  and  L12. 

Examples. — (1)  If  the  specific  volume  of  ice  is  1-087,  and  that  of  water 
unity,  find  the  lowering  of  the  freezing  point  of  water  when  the  pressure 
increases  one  atmosphere  (latent  heat  of  ice  =  80  cal.).  Here  v2  -  vx  =  0*087, 
T  =  273,  dp  =  76  cm.  mercury.  The  specific  gravity  of  mercury  is  13*5,  and 
the  weight  of  a  column  of  mercury  of  one  square  cm.  cross  section  is 
76  x  13-5  =  1,033  grams.  Hence  dp  =  1,033  grams,  L12  =  80  cal.  =  80  x  47,600 
C.G.S.  or  dynamical  units.     From  (3),  dT  =  0*0064°  C.  per  atmosphere. 

(2)  For  naphthalene  T  =  352-2,  v%  -  vt  =  0*146  ;  L12  =  35*46  cal.  Find 
the  change  of  melting  point  per  atmosphere  increase  of  pressure.    dT=  0*031. 

II.  The  slopes  of  the  pT-curves  at  the  triple  point.  Let  L12, 
L23,  L31  be  the  latent  heats  of  conversion  of  a  substance  from  states 
1  to  2  ;  2  to  3  ;  3  to  1  respectively  ;  vv  v2,  vz  the  respective  volumes 
of  the  substance  in  states  1,  2,  3  respectively ;  let  T  denote  the 
absolute  temperature  at  the  triple  point.  Then  dp/dT  is  the  slope 
of  the  tangent  to  these  curves  at  the  triple  point,  and 

/ty\    _       L12       .  /ty\  L2S      m  /ty\    _       £31         U\ 

\dTj12  T{v2-Viy  \7>T)n-T(vt-vJ'  \7>T/n  Tfa-vJ  ™ 
The  specific  volumes  and  the  latent  heats  are  generally  quite 
different  from  the  three  changes  of  state,  and  therefore  the  slopes 
of  the  three  curves  at  the  triple  point  are  also  different.  The 
difference  in  the  slopes  of  the  tangents  of  the  solid-vapour  (hoar 
frost  line),  and  the  liquid-vapour  (steam  line)  curves  of  water 
(Fig.  39)  is 

\7>TJ1Z      \dTj23      T\v3  -Vl      vt-  vj'       *        W 

At  the  triple  point 

Lu  =  LU  +  L23;  and  (v,  -  vj  =  {v2  -  vj  +  (vz  -  v2).       (6) 

Example. — As  a  general  rule,  the  change  of  volume  on  melting,  (v2-  v{j, 
is  very  small  compared  with  the  change  in  volume  on  evaporation,  (vs-v^), 


154 


HIGHER  MATHEMATICS. 


§57. 


or  sublimation,  (v3  -  vj  ;  hence  v%  -  vx  may  be  neglected  in  comparison  with 
the  other  volume  changes.     Then,  from  (5)  and  (6), 


ma-(m 


(7) 


Hence  calculate  the  difference  in  the  slope  of  the  hoar  frost  and  steam  lines 
for  water  at  the  triple  point.  Latent  heat  of  water  =  80  ;  L12  =  80  x  42,700; 
T  =  273,  vs  -  v2  =  209,400  c.c.  Substitute  these  values  on  the  right-hand  side 
of  the  last  equation.     Ansr.  0*059. 

The  above  deductions  have  been  tested  experimentally  in  the 
case  of  water,  sulphur  and  phosphorus  ;  the  results  are  in  close 
agreement  with  theory. 


Fig.  65. 


§  57.  Maximum  and  Minimum  Values  of  a  Function. 

By  plotting  the  rates,  7,  at  which  illuminating  gas  flows  through 
the  gasometer  of  a  building  as  ordinates,  with  time,  t,  as  abscissae, 
a  curve  resembling  the  adjoining  diagram  (Fig.  65)  is  obtained. 

It  will  be  seen  that  very  little  gas 
is  consumed  in  the  day  time,  while 
at  night  there  is  a  relatively  great 
demand.  Observation  shows  that 
as  t  changes  from  one  value  to 
another,  V  changes  in  such  a  way 
that  it  is  sometimes  increasing  and 
sometimes  decreasing.  In  conse- 
quence, there  must  be  certain  values  of  the  function  for  which  V, 
which  had  previously  been  increasing,  begins  to  decrease,  that  is 
to  say,  V  is  greater  for  this  particular  value  of  t  than  for  any 
adjacent  value  ;  in  this  case  V  is  said  to  have  a  maximum  value. 
Conversely,  there  must  be  certain  values  oif(t)  for  which  V,  having 
been  decreasing,  begins  to  increase.  When  the  value  of  V,  for 
some  particular  value  of  t,  is  less  than  for  any  adjacent  value  of  t, 
V  is  said  to  be  a  minimum  Yalue. 

Imagine  a  variable  ordinate  of  the  curve  to  move  perpendicu- 
larly along  Ot,  gradually  increasing  until  it  arrives  at  the  position 
M1PV  and  afterwards  gradually  decreasing.  The  ordinate  at  M1P1 
is  said  to  have  a  maximum  value.  The  decreasing  ordinate,  con- 
tinuing its  motion,  arrives  at  the  position  NYQlf  and  after  that 
gradually  increases.  In  this  case  the  ordinate  at  J^Qj  is  said  to 
have  a  minimum  value. 

The  terms  "maximum"  and  "minimum"  do  not  necessarily 


§  58.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  155 

denote  the  greatest  and  least  possible  values  which  the  function 
can  assume,  for  the  same  function  may  have  several  maximum  and 
several  minimum  values,  any  particular  one  of  which  may  be 
greater  or  less  than  another  value  of  the  same  function.  In 
walking  across  a  mountainous  district  every  hill-top  would  repre- 
sent a  maximum,  every  valley  a  minimum. 

The  mathematical  form  of  the  function  employed  in  the  above 
illustration  is  unknown,  the  curve  is  an  approximate  representation 
of  corresponding  values  of  the  two  variables  determined  by  actual 
measurements. 

Example. — Plot  the  curve  represented  by  the  equation  y  =  sin  x.  Give 
x  a  series  of  values  %irt  ir,  f  t,  2ir,  and  so  on.     Show  that 

Maximum  values  of  y  occur  for  x  =  fir,  fr,  fx, . . . 

Minimum  values  of  y  occur  for  x  =  -  \ir,  fir,  £>r, . .  . 
The  resulting  curve  is  the  harmonic  or  sine  curve  shown  in  Fig.  52,  page  136. 

One  of  the  most  important  applications  of  the  differential  cal- 
culus is  the  determination  of  maximum  and  minimum  values  of  a 
function.  Many  of  the  following  examples  can  be  solved  by  special 
algebraic  or  geometric  devices.  The  calculus,  however,  offers  a  sure 
and  easy  method  for  the  solution  of  these  problems. 

§  58.  How  to  find  Maximum  and  Minimum  Yalues  of  a 
Function. 

If  a  cricket  ball  be  thrown  up  into  the  air,  its  velocity,  ds/dt, 
will  go  on  diminishing  until  the  ball  reaches  the  highest  point  of 
its  ascent.  Its  velocity  will  then  be  zero.  After  this,  the  velocity 
of  the  ball  will  increase  until  it  is  caught  in  the  hand.  In  other 
words,  ds/dt  is  first  positive,  then  zero,  and  then  negative.  This 
means  that  the  distance,  s,  of  the  ball  from  the  ground  will  be 
greatest  when  ds/dt  is  least ;  s  will  be  a  maximum  when  ds/dt  is 
zero. 

We  generally  reckon  distances  up  as  positive,  and  distances 
down  as  negative.  We  naturally  extend  this  to  velocities  by 
making  velocities  directed  upwards  positive,  and  velocities  directed 
downwards  negative.  Thus  the  velocity  of  a  falling  stone  is 
negative  although  it  is  constantly  getting  numerically  greater  (i.e., 
algebraically  less).  We  also  extend  this  convention  to  directed 
acceleration ;  but  we  frequently  call  an  increasing  velocity  positive, 
and  a  decreasing  velocity  negative  as  indicated  on  page  18. 


156 


HIGHER  MATHEMATICS. 


58. 


Numerical  Illustration. — The  distance,  s,  of  a  body  from  the  ground  at 
any  instant,  t,  is  given  by  the  expression 

s  =  \g&  +  v0t, 
where  v0  represents  the  velocity  of  the  body  when  it  started  its  upward  or 
downward  journey ;  g  is  a  constant  equal  to  -  32  when  the  body  is  going 
upwards,  and  to  +  32  when  the  body  is  coming  down.  (  Now  let  a  cricket  ball 
be  sent  up  from  the  hand  with  a  velocity  of  64  feet  per  second,  it  will  attain 
its  highest  point  when  dsjdt  is  zero,  but 


=  -  32*  +  v0 


.       v0      64      ,       ds 
•'=32  =  32'wlien^ 


=  0. 


Let  us  now  trace  the  different  values  which  the  tangent  to  the 
curve  at  any  point  X  (Fig.  66)  assumes  as  X  travels  from  A  to  P ; 

from  P  to  B  ;  from  B  to  Q  ;  and 

from  Q  to  G;    let  a  denote  the 

angle   made    by   the   tangent   at 

any  point  on  the  curve  with  the 

sc-axis.    Eemember  that  tan  0°  =  0 ; 

tan  90°  =  oo  ;  when  a  is  less  than 

90°,  tan  a  is  positive  ;  and  when  a 

is  greater  than  90°  and  less  than 

180°  tan  a  is  negative. 

First,  as  P  travels  from  A  to  P,  x  increases,  y  increases.     The 

tangent  to  the  curve  makes  an  acute  angle,  alf  with  the  rc-axis. 

In  this  case,  tan  a  is  positive,  and  also 


M  N 

Fig.  66. — Maximum  and  Minimum 


dy  _ 


(1) 


At  P,  the  tangent  is  parallel  to  the  a;-axis ;  y  is  a  maximum,  that 
is  to  say,  tan  a  is  zero,  and 

dy 


dx 


=  0 


(2) 


Secondly,  immediately  after  passing  P,  the  tangent  to  the  curve 
makes  an  obtuse  angle,  a2,  with  the  ic-axis,  that  is  to  say,  tan  a  is 
negative,  and 

dx~ W 

The  tangent  to  the  curve  reaches  a  minimum  value  at  NQ ;  at  Q 
the  tangent  is  again  parallel  to  #-axis,  y  is  a  minimum  and  tan  a, 
as  well  as 


dx 


=  0. 


(4) 


;/ 


§  58.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  157 

After  passing  Qy  again  we  have  an  acute  angle,  a3,  and, 

l  =  + (*) 

Thus  we  see  that  every  time  dyjdx  becomes  zero,  y  is  either  a 
maximum  or  a  minimum.  Hence  the  rule  :  When  the  first  differ- 
ential coefficient  changes  its  sign  from  a  positive  to  a  negative 
value  the  function  has  a  maximum  value,  and  when  the  first 
differential  coefficient  changes  its  sign  from  a  negative  to  a 
positive  value  the  function  has  a  minimum  value. 

There  are  some  curves  which  have  maximum  and  minimum 
values  very  much  resembling  P'  and  Q'  (Fig.  67).  These  curves 
are  said  to  have  cusps  at  F  and  Q. 

It  will  be  observed,  in  Fig.  67,  that  x 
increases  and  y  approaches    a  maximum 
value  while  the  tangent  MP'  makes  an 
acute  angle  with  the  #-axis,  that  is  to  say, 
dyjdx  is  positive.     At  ¥  the  tangent  be-  ■ 
comes  perpendicular  to  the  a;-axis,  and  in  -q- 
consequence  the  ratio  dyjdx  becomes  in- 
finite.     The  point  F  is   called    a   cusp.    FlG#  67 —Maximum  and 
After  passing  P',  dyjdx  is  negative.      In         Minimum  Cusps, 
the  same  way  it  can  be  shown  that  as  the  tangent  approaches  NQ\ 
dyjdx  is  negative,  at  Q'f  dyjdx  becomes  infinite,  and  after  passing 
Q',  dyjdx  is  positive.    Now  plot  y  =  x%,  and  you  will  get  a  cusp  at  0. 

A  function  may  thus  change  its  sign  by  becoming  zero  or  in- 
finity, it  is  therefore  necessary  for  the  first  differential  coefficient  of 
the  function  to  assume  either  of  these  values  in  order  that  it  may 
have  a  maximum  or  a  minimum  value.  Consequently,  in  order  to 
find  all  the  values  of  x  for  which  y  possesses  a  maximum  or  a 
minimum  value,  the  first  differential  coefficient  must  be  equated 
to  zero  or  infinity  and  the  values  of  x  which  satisfy  these  condi- 
tions determined. 

Examples. — (1)  Consider  the  equation  y  =  x2  -  Sx,  .*.  dyjdx  =  2x  -  8. 
Equating  the  first  differential  coefficient  to  zero,  we  have  2x  -  8  =  0 ;  or  x  =  4. 
Add  +  1  to  this  root  and  substitute  for  x  in  the  original  equation, 
when  z  =  3,y=    9  -  24  =  -  15 ; 
x  =  4,  y  =  16  -  32  =  -  16 ; 
x  =  5,  y  =  25  -  40  =  -  15. 
y  is  therefore  a  minimum  when  x  =  4,  since  a  slightly  greater  or  a  slightly 
less  value  of  x  makes  y  assume  a  greater  value.     The  addition  of  +  1  to  the 
root  gives  only  a  first  approximation.     The  minimum  value  of  the  function 


158 


HIGHER  MATHEMATICS. 


§59 


might  have  been  between  3  and  4 ;  or  between  4  and  5.     The  approximation 

may  be  carried  as  close  as  we  please  by  using  less  and  less  numerical  values 

in  the  above  substitution.     Suppose  we  substitute  in  place  of  +  1,  +  h,  then 

when  x  =  4  -  h,  y  =  h?  -  16 ; 

x  =4,  y  =       -16; 

x  =  4  +  h,  y  =  W  -  16. 
Therefore,  however  small  h  may  be,  the  corresponding  value  of  y  is  greater 
than -16.     That  is  to  say,  a;  =  4  makes  the  function  a  minimum,  Q  (Fig.  68). 
You  can  easily  see  that  this  is  so  by  plotting  the  original  equation  as  in  Fig.  68. 


V    J 

4     t 

A     k 

W 

H- 

1 


r  J 

j 

J&-* 

7 

L 

Fig.  68. 


Fig.  69. 


Fig.  70. 


(2)  Show  that  y  =  1  +  8x  -  2x2,  has  a  maximum  value,  P  (Fig.  69),  for 
X  m  2.    Plot  the  original  equation  as  in  Fig.  69. 

(3)  Show  that  y  has  neither  a  maximum  nor  a  minimum  when 
y  =  2  +  (x  -  If.  Here  dyjdx  =  3{x  -  l)2  =  0 ;  .-.  x  =  1.  But  x  =  1  does 
not  make  y  a  maximum  nor  a  minimum.  If  x  =  1,  y  =  2 ;  if  x  =  0,  y  =  1 ; 
if  x  =  2,  y  =  3,  the  graph  is  shown  in  Fig.  70.    The  critical  point  is  at  P. 


§  59.    Turning  Points  or  Points  of  Inflexion. 

Let  us  now  return  to  the  subject  of  §  58.     The  fact  that 
dy  dy 

is  not  a  sufficient  condition  to  establish  the  existence  of  maximum 
and  minimum  values  of  a  function,  although  it  is  a  rough  practical 

test.  Some  of  the  values  thus  obtained 
do  not  necessarily  make  the  function  a 
maximum  or  a  minimum,  since  a  vari- 
able may  become  zero  or  infinite  without 
changing  its  sign.  This  will  be  obvious 
from  a  simple  inspection  of  Fig.  71, 
where 

dy    n  «.  t?  .   a  dy 

^-  =  0ati?,   and,^  = 

Yet   neither   maximum    nor    minimum 
Fig.  71.-Points  of  Inflexion.  yalues  q{  ^  function  exisi     A  mrther 

test  is  therefore  required  in  order  to  decide  whether  individual 


atflf. 


§60.        FUNCTIONS  WITH  SINGULAR  PROPERTIES. 


159 


values  of  x  correspond  to  maximum  or  minimum  values  of  the 
function.  This  is  all  the  more  essential  in  practical  work  where 
the  function,  not  the  curve,  is  to  be  operated  upon. 

By  reference  to  Fig.  71  it  will  be  noticed  that  the  tangent 
crosses  the  curve  at  the  points  R  and  S.  Such  a  point  is  called  a 
turning  point  or  point  of  inflexion.  You  will  get  a  point  of 
inflexion  by  plotting  y  =  xz.  The  point  of  inflexion  marks  the 
spot  where  the  curve  passes  from  a  convex  to  a  concave,  or  from  a 
concave  to  a  convex  configuration  with  regard  to  one  of.  the  co- 
ordinate axes.  The  terms  concave  and  convex  have  here  their 
ordinary  meaning. 


Fig.  72. — Convexity  and 
Concavity. 


§  60.  How  to  Find  whether  a  Curye  is  ConcaYe  or  Convex. 

Referring  to  Fig.  72,  along  the  concave  part  from  A  to  P, 
the  numerical  value  of  tana,  regularly  decreases  to  zero.  At  P 
the  highest  point  of  the  curve 
tan  a  =  0 ;  from  this  point  to  B 
the  tangent  to  the  angle  continu- 
ally decreases.  You  will  see  this 
better  if  you  take  numbers.  Let 
ai  =  450,a2  =  135°;  .-.  tanax  =  +1, 
and  tan  a2  =  -  1.  Hence  as  you 
pass  along  the  curve  from  A  to  P 
to  B,  the  numerical  value  of  the 
tangent  of  the  curve  ranges  from 

+  1,  to  0,  to  -  1. 
The  differential  coefficient,  or  rate  of  change  of  tan  a  with  respect 
to  x  for  the  concave  curve  APB  continually  decreases.      Hence 
d(ta,iia)/dx  is  negative,  or 

d(tana)      d2y 

— dx      =  ~dx2  =  negative  value  =  <  °-     •         •         (1) 

If  a  function,  y  =  f(x),  increases  with  increasing  values  of  x,  dy/dx 
is  positive ;  while  if  the  function,  y  =  /(#),  decreases  with  increas- 
ing values  of  x,  dy/dx  is  negative. 

Along  the  convex  part  of  the  curve  BQG,  tan  a  regularly  in- 
creases in  value.  Let  us  take  numbers.  Suppose  a2  =  135°, 
ag  =  45°,  then  tan  a2  =  -  1  and  tan  a3  =  +  1.  Hence  as  you  pass 
along  the  curve  from  B  to  Q,  tan  a  increases  in  value  from  -  1 
to  0.     At  the  point  Q,  tan  a  =  0,  and  from  Q  to  C,  tan  a  continually 


160  HIGHER  MATHEMATICS.  §  CI. 

increases  in  value  from  0  to  +  1.     The  differential  coefficient  of 
tana  with  respect  to  the  convex  curve  BQC  is,  therefore,  positive,  or 

^fa)  =  p-  =  positive  value   =  >  0.        .  .  (2) 

dx  dx2 

Hence  a  curve  is  concave  or  convex  upwards,  according  as  the  second 
differential  coefficient  is  positive  or  negative. 

I  have  assumed  that  the  curve  is  on  the  positive  side  of  the 
#-axis;  when  the  curve  lies  on  the  negative  side,  assume  the  z-axis 
to  be  displaced  parallel  with  itself  until  the  above  condition  is 
attained.  A  more  general  rule,  which  evades  the  above  limita- 
tion, is  proved  in  the  regular  text-books.  The  proof  is  of  little 
importance  for  our  purpose.  The  rule  is  to  the  effect  that  "a 
curve  is  concave  or  convex  upwards  according  as  the  product  of 
the  ordinate  of  the  curve  and  the  second  differential  coefficient,  i.e., 
according  as  yd2y/dx2  is  positive  or  negative  ". 

Examples.— (1)  Show  that  the  curves  y  =  log  a;  and  y  =  xlogx  are  re- 
spectively concave  and  convex  towards  the  avaxis.  Hint. 
dhjfdx2  —  -x~2  for  the  former;  and  +  a;-1  for  the  latter. 
The  former  is  therefore  concave,  the  latter  convex,  as  shown 
in  Fig.  73.  Note :  If  you  plot  y  =  log  x  on  a  larger  scale 
you  will  see  that  for  every  positive  value  of  x  there  is  one 
and  only  one  value  of  y ;  the  value  of  y  will  be  positive  or 
negative  according  as  x  is  greater  or  less  than  unity.     When 

^  x  =  l,  y=0 ;  when  x=0,  y=  -  oo  ;  when  x=  +  oo,  y=  +  oo. 

There  is  no  logarithmic  function  for  negative  values  of  x. 

(2)  Show  that  the  parabola,  yii=  4aa,  is  concave  upwards  below  the  aj-axis 
(where  y  is  negative)  and  convex  upwards  above  the  a;-axis. 

§  61.    How  to  Find  Turning  Points  or  Points  of  Inflexion. 

From  the  above  principles  it  is  clearly  necessary,  in  order  to 
locate  a  point  of  inflexion,  to  find  a  value  of  x,  for  which  tan  a 
assumes  a  maximum  or  a  minimum  value.  Bat 
dym  ;  d(tana)  _  d2y 
Una  =  dx^-~dx—-dx-2  =  0'  '  '  (3> 
Hence  the  rule  :  In  order  to  find  a  point  of  inflexion  we  must 
equate  the  second  differential  coefficient  of  the  function  to  zero ; 
find  the  value  of  x  which  satisfies  these  conditions ;  and  test  if  the 
second  differential  coefficient  does  really  change  sign  by  substitut- 
ing in  the  second  differential  coefficient  a  value  of  x  a  little  greater 
and  one  a  little  less  than  the  critical  value.  If  there  is  no  change 
of  sign  we  are  not  dealing  with  a  point  of  inflexion 


y         J- 

f 

—  *vf  — 

—  5?  ^"  ^Hy*^?*-" 

X 

§62. 


FUNCTIONS  WITH  SINGULAR  PROPERTIES. 


161 


Examples. — (1)  Show  that  the  curve  y= a  +  (x  -  6)3  has  a  point  of  inflexion 
at  the  point  y  =  a,  x  =  b.  Differentiating  twice  we  get  d^y/dx2  =  6(x  -  b). 
Equating  this  to  zero  we  get  x  =  b ;  by  substituting  x  =b  in  the  original  equa- 
tion, we  get  y  —  a.  When  x  =  b  -  1  the  second  differential  coefficient  is 
negative,  when  03=6  +  1  the  second  differential  coefficient  is  positive.  Hence 
there  is  an  inflexion  at  the  point  (b,  a).     See  Fig.  70,  page  158. 

(2)  For  the  special  case  of  the  harmonic  curve,  Fig.  52,  page  136,  y=8in  x ; 
.'.  cPy/dx2  =  -  sin  x  =  -  y,  that  is  to  say,  at  the  point  of  inflexion  the  ordinate 
y  changes  sign.  This  occurs  when  the  curve  crosses  the  a;-axis,  and  there  are 
an  infinite  number  of  points  of  inflexion  for  which  y  =  0. 

(3)  Show  that  the  probability  curve,  y  =  fee-*2*2,  has  a  point  of  inflexion 
for  x  =  ±  g/l/fc.     (Fig.  168,  page  513.) 

(4)  Show  that  Roche's  vapour  pressure  curve  p  =  a&0 '("*+"#)  has  a  point 
of  inflexion  when  0=m(log  6-2n)/2/i2;  and p  =  a¥l°& & - 2*)M<>g &.  gee  Ex.  (6), 
page  67  ;  and  Fig.  88,  page  172. 


§  62.    Six  Problems  in  Maxima  and  Minima. 

It  is  first  requisite,  in  solving  problems  in  maxima  and  minima, 
to  express  the  relation  between  the  variables  in  the  form  of  an 
algebraic  equation,  and  then  to  proceed 
as  directed  on  page  157.  In  the  ma- 
jority of  cases  occurring  in  practice, 
it  only  requires  a  little  common-sense 
reasoning  on  the  nature  of  the  problem, 
to  determine  whether  a  particular  value 
of  x  corresponds  with  a  maximum  or 
a  minimum.  The  very  nature  of  the 
problem  generally  tells  us  whether  we 
are  dealing  with  a  maximum  or  a  mini- 
mum, so  that  we  may  frequently  dis- 
pense with  the  labour  of  investigating  B 
the  sign  of  the  second  derivative. 

I.  Divide  a  line  into  any  two  parts  such  that  the  rectangle 
having  these  two  parts  as  adjoining  sides  may  have  the  greatest 
possible  area.  If  a  be  the  length  of  the  line,  x  the  length  of  one 
part,  a  -  x  will  be  the  length  of  the  other  part ;  and,  in  conse- 
quence, the  area  of  the  rectangle  will  be 


y  =  (a  -  x)x. 


Differentiate,  and 


dy 
dx 


=  a  -  2x. 


162 


HfGHER  MATHEMATICS. 


§62. 


that  is  to  say,  the  line  a  must  be 
and  the  greatest  possible  rectangle 


Equate  to  zero,  and,  x  =  ha 
divided  into  two  equal  parts, 
is  a  square. 

II.  Find  the  greatest  possible  rectangle  that  can  be  inscribed 
in  a  given  triangle.  In  Fig.  74,  let  b  denote  the  length  of  the  base 
of  the  triangle  ABC,  h  its  altitude,  x  the  altitude  of  the  inscribed 
rectangle.  We  must  first  find  the  relation  between  the  area  of  the 
rectangle  and  of  the  triangle.     By  similar  triangles,  page  603, 

AH:AK=  BC:DE;  h:h  -  x  =  b:  DE, 
but  the  area  of  the  rectangle  is  obviously  y  =  DE  x  KH,  and 


DE  =  %h  -  x),  KH 


V  =  j:(hx 


x2). 


^  =  h 
dx 


hv"      ~' '  '  '  *  ~  h" 

It  is  the  rule,  when  seeking  maxima  and  minima,  to  simplify 
the  process  by  omitting  the  constant  factors,  since,  whatever  makes 
the  variable  hx  -  x2  a  maximum  will  also  make  b(hx  -  x2)/h  a 
maximum.1  Now  differentiate  the  expression  obtained  above  for 
the  area  of  the  rectangle  neglecting  b/h,  and  equate  the  result  to 
zero,  in  this  way  we  obtain 

-  2x  =  0;  or  x  =7i. 
A 

That  is  to  say,  the  height  of  the  rectangle  must  be  half  the  altitude 
of  the  triangle. 

III.  To  out  a  sector  from  a  circular  sheet 
of  metal  so  that  the  remainder  can  be  formed 
into  a  conical-shaped  vessel  of  maximum 
capacity.  Let  ACB  (Fig.  75)  be  a  circular 
plate  of  radius,  r,  it  is  required  to  cut  out  a 
portion  AOB  such  that  the  conical  vessel 
formed  by  joining  OA  and  OB  together  may 
hold  the  greatest  possible  amount  of  fluid. 
Let  x  denote  the  angle  remaining  after  the 

sector  AOB  has  been  removed.      We  must  first  find  a  relation 

between  x  and  the  volume,  v,  of  the  cone.2 

The  length  of  the  arc  ACB  is  j^xttt,  (3),  page  603,  and  when 


1  This  is  easily  proved,  for  let  y  =  cf{x),  where  c  has  any  arbitrary  constant  value. 
For  a  maximum  or  minimum  value  dyfdx  =  cf'(x)  =  0,  and  this  can  only  occur  where 

/'(*)  =  0. 

2  Mensuration  formulae  (1),  (3),  (4),  (27),  §  191,  page  603  j  and  (1),  page  606,  will  be 
required  for  this  problem. 


§  62.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  163 

the  plate  is  folded  into  a  cone,  this  is  also  the  length  of  the  peri- 
meter of  the  circular  base  of  the  cone.  Let  B  denote  the  radius  of 
the  circular  base.  The  perimeter  of  the  base  is  therefore  equal  to 
2irR.     Hence, 

2ttR  =  -7rr;  or,  J8-~.    .         .         .         (1) 

If  h  is  the  height  of  the  vertical  oone, 

r2  =  B2  +  h2;ov,h  =  Jr2  -  B2.        .         .         (2) 
The  volume  of  the  cone  is  therefore 


-^-iOy--©'  •  _« 

Rejecting  the  constants,  v  will  be  a  maximum  when  x2  J^-rr2  -  x2, 
or  when  x4^2  -  x2)  is  a  maximum.     That  is,  when 

^-{x*(±7r2  -  x2)}  =  (16tt2  -  §x2)x*  =  0. 

If  x  =  0,  we  have  a  vertical  line  corresponding  with  a  cone  of  mini- 
mum volume.     Hence,  if  x  is  not  zero,  we  must  have 

167T2  -  6a?2  -  0;  or,  x  =  2  J*  x  180°  =  294°. 
Hence  the  angle  of  the  removed  sector  is  about  360°  -  294°  =  66°. 
The  application  to  funnels  is  obvious.     Of  course  the  sides  of  the 
chemists'  funnel  has  a  special  slope  for  other  reasons. 

IV.  At  what  height  should  a  light  be  placed  above  my  writing  table 
in  order  that  a  small  portion  of  the  surface  of  the  table,  at  a  given 
horizontal  distance  away  from  the  foot  of 
the  perpendicular  dropped  from  the  light 
on  to  the  table,  may  receive  the  greatest 
illumination  possible  ?  Let  S  (Fig.  76) 
be  the  source  of  illumination  whose  dis- 
tance, x,  from  the  table  is  to  be  deter- 
mined in  such  a  way  that  B  may  receive 
the  greatest  illumination.  Let  AB  =  a, 
and  a  the  angle  made  by  the  incident 
rays  SB  =  r  on  the  surface  B. 

It  is  known  that  the  intensity  of  illumination,  y,  varies  inversely 
as  the  square  of  the  distance  of  SB,  and  directly  as  the  sine  of  the 
angle  of  incidence.  Since,  by  Pythagoras'  theorem  (Euclid,  i.,  47), 
r2  =  a2  +  x2 ;  and  sin  a  =  x/r,  in  order  that  the  illumination  may 
be  a  maximum, 

_  sina__#  _  x  x 

V  ~  ~^~~r^"r2  Ja2  +  x2= (a2  +  x2)i 


164 


HIGHER  MATHEMATICS. 


62. 


By  differentiation,  we  get 

a2  -  2x2 


-=0;  ,•.*-  ajh 


must  be  a  maximum. 

dx    (a2  +  x2f 

The  interpretation  is  obvious.  The  height  of  the  light  must  be 
0*707  times  the  horizontal  distance  of  the  writing  table  from  the 
"  foot "  A.  Negative  and  imaginary  roots  have  no  meaning  in  this 
problem. 

V.  To  arrange  a  number  of  voltaic  cells  to  furnish  a  maximum 
current  against  a  known  external  resistance.  Let  the  electro- 
motive force  of  each  cell  be  E,  and  its  internal  resistance  r.  Let 
B  be  the  external  resistance,  n  the  total  number  of  cells.  Assume 
that  x  cells  are  arranged  in  series  and  n/x  in  parallel.  The  electro- 
motive force  of  the  battery  is  xE.  Its  internal  resistance  x2r/n, 
The  current  (7,  according  to  the  text- books  on  electricity,  is  given 
by  the  relation 

'  '  '  dx 


0- 


B  + 


(B  +  V)' 


Equate  to  zero,  and  simplify,  B  =  rx2/n,  remains.  This  means 
that  the  battery  must  be  so  arranged  that  its  internal  resistance 
shall  be  as  nearly  as  possible  equal  to  the  external  resistance. 

The  theory  of  maxima  and  minima  must  not  be  applied  blindly 
to  physical  problems.      It  is  generally  necessary  to   take   other 

things  into  consideration.  An  ar- 
rangement that  satisfies  one  set  of 
conditions  may  not  be  suitable  for 
another.  For  instance,  while  the 
above  arrangement  of  cells  will  give 
the  maximum  current,  it  is  by  no 
means  the  most  economical. 

VI.  To  find  the  conditions  which 
must  subsist  in  order  that  light  may 
travel  from  a  given  point  in  one 
medium  to  a  given  point  in  another 
medium  in  the  shortest  possible  time. 
Let  SP  (Fig.  77)  be  a  ray  of  light 
incident  at  P  on  the  surface  of 
separation  of  the  media  M  and  M ; 
let  PB  be  the  refracted  ray  in  the  same  plane  as  the  incident 


Fm.  77. 


§  62.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  165 

ray.  If  PN  is  normal  (perpendicular)  to  the  surface  of  incidence, 
then  the  angle  NPS  =  *,  is  the  angle  of  incidence ;  and  the  angle 
N'PB  =  r,  is  the  angle  of  refraction.  Drop  perpendiculars  from  S 
and  B  on  to  A  and  B,  so  that  SA  =  a,  BB  =  b.  Now  the  light  will 
travel  from  S  to  B,  according  to  Fermat's  principle,  in  the  shortest 
possible  time,  with  a  uniform  velocity  different  in  the  different  media 
M  and  M '.  The  ray  passes  through  the  surface  separating  the  two 
media  at  the  point  P,  let  AP  =  x,  BP  =  p  -  x.  Let  the  velocity 
of  propagation  of  the  ray  of  light  in  the  two  media  be  respectively 
VY  and  V2  units  per  second.  The  ray  therefore  travels  from  S  to  P 
in  PS/Vl  seconds,  and  from  P  to  B  in  BP/V>2  seconds,  and  the  total 
time,  t,  occupied  in  transit  from  S  to  B  is  the  sum 

<-tt+tt'    ;   •    •    (1) 

From  the  triangles  SAP  and  PBB,  as  indicated  in  (1),  page  603,  it 
follows  that  PS  =  J  a2  +  x2 ;  and  BP  -  Jb2  +  (p  -  xf.  Sub- 
stituting these  values  in  (1),  and  differentiating  in  the  usual  way, 
we  get 

dt  =  x  _  P-x  =o  (2) 

dx       Vx  Ja*  +  a*       V2  Jb*  +  (p  -  xf        '  W 

Consequently,  by  substituting  for  PS,  BP,  AP,  and  BP  as  above, 
we  get  from  the  preceding  equation  (2),  solved  for  VJV2, 
AP  x 

sini       PS__        J  a2  +  x2  V\ 


sinr 


*BP  p-x  V' 


BP  Jb2  +  (p  -  x)2 
This  result,  sometimes  called  Snell's  law  of  refraction,  shows  that 
the  sines  of  the  angles  of  incidence  and  refraction  must  be  pro- 
portional to  the  velocity  of  the  light  in  the  two  given  media  in 
order  that  the  light  may  pass  from  one  point  to  the  other  in  the 
shortest  possible  interval  of  time.  Experiment  justifies  Format's 
guess.  •  The  ratio  of  the  sines  of  the  two  angles,  therefore,  is 
constant  for  the  same  two  media.  The  constant  is  usually  de- 
noted by  the  symbol  /m,  and  called  the  index  of  refraction. 

Examples. — (1)  The  velocity  of  motion  of  a  wave,  of  length  A,  in  deep 
water  is  V  =  *J(\[a  +  a/A),  a  is  a  constant.  Required  the  length  of  the  wave 
when  the  velocity  is  a  minimum.     (N.  Z.  Univ.  Exam.  Papers.)     Ansr.  \  =  a. 

(2)  The  contact  difference  of  potential,  E,  between  two  metals  is  a 
function   of  the  temperature,  0,  such   that  E  =  a  +  be  +  cd2.      How  high 


166 


HIGHEK  MATHEMATICS. 


§62. 


Fig.  78. 


V 


must  the  temperature  of  one  of  the  metals  be  raised  in  order  that  the 
difference  of  potential  may  be  a  maximum  or  a  minimum,  a,  o,  c  are  con- 
stants.    Ansr.  6  =  -  b/2c. 

(3)  Show  that  the  greatest  rectangle  that  can  be  inscribed  in  the  circle 
x2  +  y2  =  r2  is  a  square.  Hint.  Draw  a  circle  of  radius  r,  Fig.  78.  Let  the 
sides  of  the  rectangle  be  2x  and  2y  respectively ;  .\  area  =  4xy,  x2  +  y2  =  r2. 
Solve  for  y,  and  substitute  in  the  former  equation.  Differ- 
entiate, etc.,  and  then  show  that  both  x  and  y  are  equal 
to  r  *J%,  etc. 

(4)  If  v0  be  the  volume  of  water  at  0°  C.,  v  the  volume 
at  0°  C,  then,  according  to  Hallstrom's  formula,  for  tem- 
peratures between  0°  and  30°, 

v = v0{l  -0-000057,5770  +  O'OOOOO7,56O102-  O-OOOOOO.O35O903). 
Show  that  the  volume  is  least  and  the  density  greatest 
when  0  =  3-92.     The  graph  is  shown  in  Fig.  79.     In  the  working  of  this  ex- 
ample, it  will  be  found  simplest  to  use  a,  b,  c  . . .  for  the  numerical  coefficients, 
differentiate,  etc.,  for  the  final  result,  restore  the  numerical  values 
of  a,  b,  c  . . . ,  and  simplify.     Probably  the  reader  has  already 
done  this. 

(5)  Later  on  I  shall  want  the   student   to  show  that  the 
expression  s/(q2  -  n2)2  +  4/%2  is  a  minimum  when  n2  =  q2  -  If2. 

(6)  An  electric  current  flowing  round  a  coil  of  radius  r  exerts 
Fig.  79.      a  force  F  on  a  small  magnet  whose  axis  is  at  some  point  on  a 

line  drawn  through  the  centre  and  perpendicular  to  the  plane  of 
the  coil.  If  x  is  the  distance  of  the  magnet  from  the  plane  of  the  coil,  F  = 
xj(r2  +  x2)5!2.     Show  that  the  force  is  a  maximum  when  x  =  £r. 

(7)  Draw  an  ellipse  whose  area  for  a  given  perimeter  shall  be  a  maximum. 
Hint.  Although  the  perimeter  of  an  ellipse  can  only  be  represented  with 
perfect  accuracy  by  an  infinite  series,  yet  for  all  practical  purposes  the 
perimeter  may  be  taken  to  be  tt(x  +  y)  where  x  and  y  are  the  semi-major  and 
semi-minor  axes.  The  area  of  the  ellipse  is  z  =  irxy.  Since  the  perimeter  is  to 
life  constant,  a  =  ir(x  +  y)  or  y  =  ajv  -  x.  Substitute  this  value  of  y  in  the 
former  expression  and  z  =  ax  -  irx2.  Hence,  x  =  a/2?r  when  z  is  a  maximum. 
Substitute  this  value  of  x  in  y  =  ajir  -  x,  and  y  =  af2ir,  that  is  to  say, 
x  =  y  =  a./27r,  or  of  all  ellipses  the  circle  has  the  greatest  area.  Boys'  leaden 
water-pipes  designed  not  to  burst  at  freezing  temperatures,  are  based  on  this 
principle.  The  cross  section  of  the  pipe  is  elliptical.  If  the  contained  water 
freezes,  the  resulting  expansion  makes  the  tube  tend  to  become  circular  in  cross 
section.  The  increased  capacity  allows  the  ice  to  have  more  room  without 
putting  a  strain  on  the  pipe. 

(8)  If  A,  B  be  two  sources  of  heat,,  find  the  position  of  a  point  O  on  the 
line  AB  =  a,  such  that  it  is  heated  the  least  possible.  Assume  that  the  in- 
tensity of  the  heat  rays  is  proportional  to  the  square  of  the  distance  from  the 
source  of  heat.  Let  AO  =  x,  BO  =  a  -  x.  The  intensity  of  each  source  of 
heat  at  unit  distance  away  is  a  and  &.  The  total  intensity  of  the  heat  which 
reaches  O  is  I  =  ax~2  +  $(a  -  x)~2.  Find  dlfdx.  I  is  a  minimum  when 
x  =  !Ja.al(i/a+  fj$). 


§62.         FUNCTIONS  WITH  SINGULAR  PROPERTIES.         167 

(9)  The  weight,  W  (lbs.  per  sec),  of  flue  gas  passing  up  a  chimney  at 
different  temperatures  T,  is  represented  by  W=  A(T  -  T0)  (1  +  a.T)~\  where 
A  is  a  constant,  T  the  absolute  temperature  of  the  hot  gases  passing  within 
the  chimney,  T0  the  temperature  (°0)  of  the  outside  air,  a  =  ^7  the  co- 
efficient of  expansion  of  the  gas.  Hence  show  that  the  greatest  amount  of 
gas  will  pass  up  the  chimney — the  "best  draught"  will  occur — when  the 
temperature  of  the  "  flue  gases "  is  nearly  333°  C.  and  the  temperature  of 
the  atmosphere  is  15°  C. 

(10)  If  VC  denotes  the  "input"  of  a  continuous  current  dynamo;  <r  the 
fixed  losses  due  to  iron,  friction,  excitation,  etc. ;  tC2,  the  variable  losses ;  0, 
the  current,  then  the  efficiency,  E,  is  given  by  E  =  1  -  (a  -  tC2)/VC.  Show 
that  the  efficiency  will  be  a  minimum  when  <r  =  tC2.     Hint.  Find  dEjdC,  etc. 

(11)  Show  that  xx  is  a  maximum  when  x  =  e.  Hint,  dyjdx  =  xx(l  - 
logjc)/aj2;  .*.  logo;  =  1,  etc. 

(12)  A  submarine  telegraph  cable  consists  of  a  circular  core  surrounded  by 
a  concentric  circular  covering.  The  speed  of  signalling  through  this  varies  as 
1 :  a;2log  x  ~ 1,  where  x  denotes  the  ratio  of  the  radius  of  the  core  to  that  of  the 
covering.  Show  that  the  fastest  signalling  can  be  made  when  this  ratio  is 
0-606  . . .  Hint.  1 :  \Z<T=  0-606. 

(18)  The  velocity  equations  for  chemical  reactions  in  which  the  normal 
course  is  disturbed  by  autocatalysis  are,  for  reactions  of  the  first  order, 
dx/dt  =  kx(a  -  x) ;  or,  dx\dt  =  k(b  +  x)  (a  -  x).  Hence  show  that  the 
velocity  of  the  reaction  will  be  greatest  when  x  =  \a  for  the  former  reaction, 
and  x  =  \  (a  -  b)  for  the  latter. 

(14)  A  privateer  has  to  pass  between  two  lights,  A  and  B,  on  opposite 
headlands.  The  intensity  of  each  light  is  known  and  also  the  distance  be- 
tween them.  At  what  point  must  the  privateer  cross  the  line  joining  the 
lights  so  as  to  be  illuminated  as  little  as  possible.  Given  the  intensity  of  the 
light  at  any  point  is  equal  to  its  illuminating  power  divided  by  the  square  of 
the  distance  of  the  point  from  the  source  of  light.  Let  px  and  p2  respectively 
denote  the  illuminating  power  of  each  source  of  light.  Let  a  denote  the 
distance  from  A  to  B.  Let  x  denote  the  distance  from  A  to  the  point  on 
AB  where  the  intensity  of  illumination  is  least ;  hence  the  ship  must  be 
illuminated  pjx2  +  pj(a  -  x)2.      This  function  will  be  a  minimum   when 

(15)  Assuming  that  the  cost  of  driving  a  steamer  through  the  water  varies 
as  the  cube  of  her  speed,  show  that  her  most  economical  speed  through  the 
water  against  a  current  running  V  miles  per  hour  is  |F.  Let  x  denote  the 
speed  of  the  ship  in  still  water,  x  -  V  will  then  denote  the  speed  against  the 
current.  But  the  distance  traversed  is  equal  to  the  velocity  multiplied  by 
the  time.  Hence,  the  time  taken  to  travel  s  miles  will  be  s/(x  -  V).  The 
cost  in  fuel  per  hour  is  ax*,  where  a  is  the  constant  of  proportion.  Hence : 
Total  cost  =  asaP/fa  -  V).  Hence,  aP/fa  -  V)  is  to  be  a  minimum.  Differen- 
tiate as  usual,  and  we  get  x  =  f  V.  The  captain  of  a  river  steamer  must  be 
always  applying  this  fact  subconsciously. 

(16)  The  stiffness  of  a  rectangular  beam  of  any  given  material  is  propor- 
tional to  its  breadth,  and  to  the  cube  of  its  depth,  find  the  stiffest  beam  that 


168 


HIGHER  MATHEMATICS. 


§63. 


can  be  cut  from  a  circular  tree  12  in.  in  diameter.  Let  x  denote  the  breadth 
of  the  beam,  and  y  its  depth ;  obviously,  122  ■—  x2  +  y2.  Hence,  the  depth  of 
the  beam  is  \/l22  -  sc2;  .*.  stiffness  is  proportional  to  sc(122  -  a:2)"5".  This  is  a 
maximum  when  x  =  6.    Hence  the  required  depth,  y,  must  be  6  s/s. 

(17)  Suppose  that  the  total  waste,  y,  due  to  heat,  depreciation,  etc.,  which 
occurs  in  an  electric  conductor  with  a  resistance  R  ohms  per  mile  with  an 
electric  current,  C,  in  amperes  is  C2R  +  (17)2/i2,  find  the  relation  between  G 
and  R  in  order  that  the  waste  may  be  a  minimum.  Ansr.  GR  =  17,  which  is 
known  as  Lord  Kelvin's  rule.  Hint.  Find  dy/dRt  assuming  that  C  is  con- 
stant. Given  the  approximation  formula,  resistance  of  conductor  of  cross 
sectional  area  a  is  i?  =  0-04/a ;  .-.  C  =  425a,  or,  for  a  minimum  cost,  the  current 
must  be  425  amperes  per  square  inch  of  cross  sectional  area  of  conductor. 

§  63.    Singular  Points. 

The  following  table  embodies  the  relations  we  have  so  far  de- 
duced between  the  shape  of  the  curve  y  =  f{x)  and  the  first  four 
differential  coefficients.  Some  relations  have  been  "  brought  for- 
ward "  from  a  later  chapter.  The  symbol  " .  .  "  means  that  the 
value  of  the  corresponding  derivative  does  not  affect  the  result : — 

Table  I. — Singular  Values  of  Functions. 


Property  of  Curve. 

dy 
dt 

d*y 
dP 

dhJ 

dt* 

d*y 

dt* 

Tangent  parallel  to  jc-axis 
Tangent  parallel  to  y-axis 

Maximum         .         .        .-J 

Minimum         .        .         A 

Point  of  inflexion     . 
Convex  downwards  . 
Concave  downwards 

0 

00 

0 
0 
0 
0 

not  zero 
not  zero 

0 

+ 

•     0 

0 

+ 

0 

0 

not  zero 

+ 

It  is  perhaps  necessary  to  add  a  few  more  remarks  so  as  to  give 
the  beginner  an  inkling  of  the  vastness  of  the  subject  we  are  about 
to  leave  behind.  P.  Frost's  An  Elementary  Treatise  on  Curve 
Tracing,  London,  1872,  is  the  text-book  on  this  subject.  Before 
you  proceed  to  the  actual  plotting,  look  out  for  symmetrical  axes ; 
maxima  and  minima  ordinates ;  points  of  inflexion ;  and  asymp- 
totes. Does  the  curve  cut  the  axes  at  any  points  ?  There  may  be 
other  singular  points  besides  points  of  inflexion,  and  maxima  and 
minima. 


§63. 


FUNCTIONS  WITH  SINGULAR  PROPERTIES. 


169 


Two  or  more  branches  of  the  same  curve  will  intersect  or  cross 
one  another,  as  shown  at  0,  Fig.  80,  when  the  first  differential 
coefficient  has  two  or  more  real  unequal  values,  and  y  has  at  least 
two  equal  values.  The  number  of  intersecting  branches  is  denoted 
by  the  number  of  real  roots  of  the  first  differential  coefficient.  The 
point  of  intersection  is  called  a  multiple  point.  If  two  branches 
of  the  curve  cross  each  other  the  point  is  called  a  node ;  Cayley 
calls  it  a  crunode. 


^^3v/ 

<Lk     5 

Fig.  80. 


Examples. — (1)  In  the  lemniscate  curve,  familiar  to  students  of  crystallo- 
graphy, y2  =  a2x2  -  x4 ;  y  =  +  x  J  a2  -  x2.  y  has  two  values,  of  opposite  sign, 
for  every  value  of  x  between  +  a ;  the  curve  is  therefore  symmetrical  with 
respect  to  the  a>axis.  When  x  =  +  a,  these  two  values  of  y  become  zero ; 
but  these  are  not  multiple  points  since  the  curve  does  not  extend  beyond 
these  limits,  and  therefore  cannot  satisfy  the  above  conditions.  When  x  =  0, 
the  two  values  of  y  become  zero,  and  since  there  are  two 
values  of  y,  one  on  each  side  of  the  point  x  =  0,  y  =  0,  this 
is  a  multiple  point.  Since  dy\dx  —  ±  (a2  -  2a;2)  (a2  -  x2)~h 
becomes  +  a  when  x  =  0,  it  follows  that  there  are  two  tan- 
gents to  the  curve  at  this  point,  such  that  tan  a  =  ±  a.  In 
order  to  plot  the  curve,  give  a  some  numerical  value,  say  5. 
The  graph  is  shown  in  Fig.  80.  The  node  is  at  O.  Notice 
that  if  the  numerical  value  of  x  is  greater  than  that  of  a  you  have  to 
extract  the  square  root  of  a  negative  quantity.  This  cannot  be  done  be- 
cause we  do  not  know  a  number  which  will  give  a  negative  quantity  when 
multiplied  by  itself.  Mathematicians  have  agreed  to  call  the  square  root 
of  a  negative  number  an  "imaginary  number  '  in  contrast  with  a  "real 
number". 

(2)  The  curve  y  =  b  ±  (x  -  a)  sjx  has  a  node  at  the  point  P(a,  b).  For 
every  value  of  x  there  are  two  unequal  values  of  y,  but  when  x  =  0,  the  two 
values  of  y  =  6,  and  when  x  =  a,  also,  y  =  &.  There  are  two 
real  values  of  y  on  each  side  of  the  point  P(a,  b) ;  this  can  be 
determined  by  substituting  a  +  h,  and  a  -  h  successively  in 
place  of  x.  dyjdx  =  ±  %(3x  -  a)x "  *.  For  x  =  a,  dy/dx 
=  ±  s/a'  Hence  the  tangents  to  the  curve  at  the  node  make 
angles  with  the  cc-axis  whose  tangents  are  +  sja.  The  point 
x  —  0,  y  =  b  is  not  a  multiple  point  because  when  x  is  nega- 
tive, y  is  imaginary.  This  shows  that  the  curve  does  not  go  to  the  left  of  the 
t/-axis.     The  singular  point  is  shown  in  Fig.  81. 


— *rh— 


Fig.  81. 


A  cusp  or  spinode  (Cayley)  is  a  point  where  two  branches  of 
a  curve  have  a  common  tangent  and  stop  at  that  point,  as  shown 
in  Figs.  82  and  83.  The  branches  terminate  at  the  point  of  con- 
tact and  do  not  pass  beyond.  Hence  the  values  of  y  on  one  side 
of  the  point  are  real ;  and  on  the  other  imaginary. 


170 


HIGHER  MATHEMATICS. 


§63. 


Examples.— (1)  In  the  cissoid  curve,  y  =  b  ±  sj(x2  -  a2f ;  y  is  imaginary 
for  all  values  of  x  between  ±a.  When  x  =  ±a,  y  has  one  value;  for  any 
point  to  the  right  of  x  =  +  a,  or  to  the  left  of  x  —  -  a,  y  has  two  values ; 
dyjdx  =  ±  Sx(x2  -  a2)*  vanishes  when  x  =  a.  The  two  branches  of  the  curve 
have  therefore  a  common  tangent  parallel  to  the  <c-axis  and  there  is  a  cusp. 
The  cusp  is  said  to  be  of  the  first  species,  or  a  ceratoid  cusp.  We  now  find 
that  there  are  two  real  and  equal  values  for  the  second  differential  coefficient, 


y 

AC 

0 

+x 

s 


+9 

/ 

-* 

f~\ 

Y 

+# 

L? 

\ 

V 

-// 

\ 

Fig.  82.— Single  Cusps 
(First  Species). 


Fig.  83.— Single  Cusp 
(Second  Species). 


Fig.  84. 


namely,  dzy/dx2  =  +  3(20^  -  a2)  (a?2  -  a2)~%.  The  upper  branch  is  convex  to- 
wards the  a:-axes;  and  the  lower  branch  is  concave  towards  the  x-axis,  as 
shown  in  Fig.  82. 

(2)  The  second  differential  coefficient  of  the  curve  (y  -  x2)2  =  a?  has  two 
different  values  of  the  same  sign.  The  cusp  is  then  said  to  be  a  cusp  of  the 
second  species,  or  a  rhamphoid  cusp.  The  lower  curve  also  has  a  maximum 
when  x  =  £f .     The  general  form  of  the  curve  is  shown  in  Fig.  83. 

The  cusps  which  have  just  been  described  are  called  single 
cusps  in  contradistinction  to  double  cusps  or  points  of  oscula- 
tion in  which  the  curves  extend  to  both  sides  of  the  point  of 
contact.  These  are  what  Cayley  calls  tacnodes.  The  differential 
coefficient  has  now  two  or  more  equal  roots  and  y  has  at  least  two 
equal  values.  The  different  branches  of  the  curve  have  a  common 
tangent. 

To  distinguish  cusps  from  points  of  osculation  :  compare  the 
ordinate  of  the  curve  for  that  point  with  the  ordinates  of  the  curve 
on  each  side.  For  a  cusp,  y  and  the  first  differential  coefficient 
have  only  one  real  value. 

Examples. — (1)  The  curve  y2  =  x4(x  +  5) ;  or  what  is  the  same  thing 
y  =  ±x2\f  x  +  5,  has  a  tacnode  at  the  origin,  as  shown  in  Fig.  84. 
dy/dx=  ±  §x(x  +  4)  (x  +  5)-^,  which  becomes  zero  when  a;  =  0. 
The  two  branches  of  the  curve  are  tangent  to  one  another  at 
this  point  when  x  =  -  5,  dyjdx  =  oo  and  y=0.  The  tangent 
cute  the  axis  at  right  angles  at  the  point  x  =  -  5,  and  when  x 
is  less  than  5,  y  is  imaginary.  In  Fig.  82,  each  cusp  at  the 
Fig.  85.  tacnode  is  of  the  first  species,  but  sometimes  the  cusps  are  of 
the  second  species,  as  shown  in  Fig.  85. 


§  63.        FUNCTIONS  WITH  SINGULA^  PKOrERTiES. 


171 


+7/ 

-X 

0 

+  X, 

-y 

Fig.  86. 


(2)  The  curve  y2  -  6xy  =  sc5  has  another  kind  of  tacnode  shown  in  Fig.  86. 
The  cusp  is  of  the  first  species  on  one  side  and  of  the  second 
species  on  the  other. 

After  the  student  has  investigated  the  fifth  chapter 
he  may  be  able  to  show  that  when  u  =  0 ;  (du/~dx)v  =  0 ; 
and  (&ufby)x  =  0,  then  if 

fl*u\  f~b2u\        /l*u\*      ( "  =  node' 

l  +  =  conjugate  point. 

A  conjugate  point  or  acnode  is  one  whose  coordinates  satisfy 
the  equation  of  the  curve  and  yet  is  itself  quite  detached  from  the 
curve.  On  each  side  of  a  conjugate  point,  real  values  of  one 
coordinate  give  a  pair  of  imaginary  values  of  the 
other.  On  plotting  the  graph  of  y2  =  x2(x  -  2),  for 
example,  it  will  be  found  that  there  is  a  point  at 
the  origin  whose  coordinates  satisfy  the  equation 
of  the  curve,  and  yet  the  curve  itself  does  not  pass 
through  the  origin  as  shown  in  Fig.  87.  There  are  FlG-  87- 
no  real  values  of  y  when  x  is  less  than  2  (except  when  x  =  0,  y  =  0) 
so  that  the  curve  does  not  go  to  the  left  of  x  =  2. 

One  branch  of  the  curve  y  =  x  log  x  suddenly  stops  at  the 
origin  (Fig.  73,  page  160).  This  is  said  to  be  a  point  d' arret  or 
a  terminal  point.  When  x  =  0,  y  =  0 ;  when  x  is  negative,  y 
has  no  real  value  because  negative  numbers  have  no  logarithms. 
Do  not  waste  any  time  trying  to  show  that  y  =  0,  when  x  =  0. 
There  is  a  difficulty  which  you  will  easily  master  later  on. 

Dela  Eoche  has  published  a  "proof "  that  the  vapour  pressure, 
p,  of  water  at  any  temperature,  6,  is  given  by  the  expression 


11/ 

-y  I  x 


p  =  abm  +  ne> 

where  a,   b,  m,  and   n   are  constants.      August,   Eegnault,   and 
Magnus  found  that  the  expression  represented  their  experimental 
results  fairly  well.     But  Eegnault  {Ann.  Chim.  Phys.,  [3],  11,  273, 
1844)  has  pointed  out  that  Eoche's  formula  can  only  be  regarded 
as  an  empirical  interpolation  formula  pure  and  simple.     The  pro- 
perties of  this  equation  do  not  agree  with  the  actual  phenomena 
See  Ex.  (4),  page  161.     The  curve  has  a  point  of  inflexion,  E,  Fig.  88 
when  0  is  equal  to  Jm(log  b  -  2n)?i  ~ 2.     The  curve  has  two  branches 
GAB,  and  DC.     The  portion  AB  alone  applies  to  the  observed  rela 
tions  between  p  and  6.     For  this  branch  there  is  a  terminal  point 


172 


HIGHER  MATHEMATICS. 


§64. 


G,  when  0  =  -  m/n,  provided  b  is  greater  than  unity.     This  curve 

is  also  asymptotic  to  a  line  p  =  abxln 
parallel  to  the  0-axis.  The  other  branch 
of  the  curve,  I  may  notice  en  passant, 
is  asymptotic  to  the  same  straight  line 
and  also  to  the  straight  line  9  =  -  m/n 
parallel  to  the  ^?-axis.  I  have  asked  a 
class  of  students  to  plot  the  above  equa- 
tion and  all  missed  the  point  of  inflexion 
at  E.  As  a  matter  of  fact  you  should 
try  to  get  as  much  information  as  you 
can  by  applying  the  above  principles 
before  actual  plotting  is  attempted.    You 

will  now  see  that  if  you  know  the  formula  of  a  curve,  the  calculus 

gives  you  a  method  of  finding  all  the  critical  points  without  going 

to  the  trouble  of  plotting. 


T 

6, 

V 

1 
1 

1 

D^ 

/ 
• 

B 

- 
/ 

A 

-e 


6       0 

Fig.  88. 


§  64.   pv-Curyes. 

We  have  already  had  something  to  say  about  van  der  Waals' 
equation  for  the  relation  between  the  pressure,  p,  the  volume,  v, 
and  the  temperature,  T°  abs.,  of  a  gas 


(p +  $)(*->)  =BT- 


a) 


Now  try  and  plot  this  equation  for  any  gas  from  the  published 
values  of  a,  b,  B.  For  example,  for  ethylene  a  =  0*00786, 
b  =  0-0024,  E  =  0*0037  ;  for  carbon  dioxide,  van  der  Waals 
gives 

/    +  0-00874N  ^  _  0.0023)  =  0-00369(273  +  6),  .        (2) 

where  6  denotes  °  C.  First  fix  the  values  for  0,  and  calculate  a  set 
of  corresponding  values  of  p  and  v,  thus,  for  0°  0.,  when 
v  =  0-1,  0-05,  0-025,  0-01,  0-0075,  0-005,  0-004,  0003, . . . ; 
p  =  9-4,  19-7,  30-3,  43-3,  37-9,  23-2,  45-8,  466-8, . . . 
Make  the  successive  increments  in  v  small  when  in  the  neighbour- 
hood of  a  singular  point.  Plot  these  numbers  on  squared  paper. 
Note  the  points  of  inflexion.  Now  do  the  same  thing  with 
0  =  32°  C,  and  6  =  91°  C.  The  set  of  curves  shown  in  the 
adjoining  diagram  (Fig.  89)  were  so  obtained.     In  this  way  you 


§  64.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  173 


120 


80 


will  get  a  better  insight  into  the  "  inwardness  "  of  van  der  Waals' 
equation  than  if  pages  of  160- 
descriptive  matter  were 
appended.  Notice  that 
the  a/v2  term  has  no  ap- 
preciable influence  on  the 
value  of  p  when  v  becomes 
very  great,  and  also  that 
the  difference  between  v 
and  v  -  b  is  negligibly 
small,  as  v  becomes  very 
large.  What  does  this 
signify?     When  the  gas 

is  rarefied,  it  will  follow  Boyle's  law  pv  =  constant, 
be  the  state  of  the  gas  when  v  =  0*0023  ? 

For  convenience,  solve  (1)  in  terms  of  p,  and  treat  BT  as  if  it 
were  one  constant, 


e 

4^ 

lf 

*i 

7z) 

t 

») 

=  /?r  — 

j\ 

V 

Sf 

t  > 

^Ko 

G^ 

[± 

V- 

.. 

"Ua 

34 

=322^3 

_S2 

0.    \'J  c 

;: 

002 

003 

0 

005  v 

Fig.  89. 


What  would 


BT       a  1     t^rn      a(v  -  b)\ 

V  =  r 5  i    or,  p  =  f    -til a — s — L   . 

r        V  -  b        V2  '       '  r        v  -  0\  V2       J 


(3) 


This  enables  us  to  see  that  p  will  become  zero  when  the  fraction 
a(v  -  b)/v2  becomes  equal  to  BT.  This  fraction  attains  the 
maximum  value  a/22b,  when  v  becomes  2b ;  and  obviously,  when 
v  =  b,  this  fraction  is  zero.  Hence,  p  will  become  zero  when  BT 
is  equal  to  a/22b,  and  v  =  2b.  The  curve  for  -  20°  C.  (Fig.  89) 
cuts  the  y -axis  in  two  real  points  D  and  E  when  BT  is  less  than 
a/22b.  If  BT  =  a/22b,  Ov  is  tangent  to  the  curve  at  G. 
Let  us  now  differentiate  (3)  with  respect  to  v, 
dp_       $T    ,  2a.,  __  dp ^JT}m2a(v-by\ 


BT 


')■ 


dv  (v  -  b)2  v2 '  '  dv  (v-  by 
If  T  be  great  enough,  dp/dv  will  always  be  negative,  that  is  to  say, 
the  curve,  or  rather  its  tangent,  will  slope  from  left  to  right  down- 
wards, like  the  hyperbola  for  91°  C.  (Fig.  89).  If  v  be  small 
enough  (v  -  b)2  also  becomes  very  small,  the  curve  will  retain  its 
negative  slope  because  dp/dv  will  be  negative ;  and  when  v  =  b, 
(v  -  b)2jvz  =  0.  When  dp/dv  becomes  zero,  the  tangent  to  the 
curve  is  horizontal.  This  means  that  we  may  have  maximum  or 
minimum  values  of  p.  If  J7  is  small  enough,  dp/dv  will  have  a 
positive  value  for  certain  values  of  v.  The  curve  then  slopes  from 
left  to  right  upwards,  as  at  AB  (Fig.  89). 

You  can  now  show  that  2a(v  -  b)2/v2  has  the  maximum  value 


174 


HIGHEK  MATHEMATICS. 


§64. 


23a/336,  when  v  =  36  ;  and  gradually  approaches  zero,  as  v  becomes 
very  great.  If,  therefore,  BT  is  greater  than  23a/336,  the  maximum 
value  of  2a(v  -  bf/v3,  then  v  will  increase  as  p  decreases.  When  BT 
is  less  than  23a/336,  p  will  decrease  for  small  and  large  values  of  vf 
but  it  will  increase  in  the  neighbourhood  of  v  =  36.  Consequently, 
p  has  a  maximum  or  a  minimum  value  for  any  value  of  v  which 
makes  2a(v  -  b)2/v*  =  BT.  This  curve  resembles  that  for  0°  C. 
(Fig.  89),  for  all  values  of  BT  between  a/22b  and  2%/236  ;  when 
BT  =  23a/336,  we  have  the  point  of  inflexion  K0  (Fig.  89). 

Let  us  now  see  what  we  can  learn  from  the  second  differential 
coefficient 


d*p 
dv2 


2BT 


6a 

*54 


d2p 


BT 


Sa(v  -  bf 


(5) 


(v  -  bf       v* '     '  dv*      (v  -  bfV~  v* 

The  curve  will  have  a  point  of  inflexion  when  the  fraction 
Sa(v  -  b)2/v*  =  BT.  By  the  methods  already  described  you  can 
show  that  Sa(v  -  6)3/^4  will  be  zero  when  v  =  b  ;  and  that  it  will 
attain  the  maximum  value  34a/446  when  v  =  46.  Every  value  of 
v  which  makes  (5)  zero  will  correspond  with  a  point  of  inflexion. 
BT  may  be  equal  to,  greater,  or  less  than  34a/446.  For  all  values 
of  BT  between  23a/336  and  3%/446,  there  will  be  two  points  of  in- 
flexion, as  shown  at  F  and  G  (Fig.  89).  When  BT  exceeds  the 
value  3%/446,  we  have  a  branch  of  the  rectangular  hyperbola  as 
shown  for  91°. 

If  we  take  the  experimental  curves  obtained  by  Andrews  for 
the  relation  between  the  pressure,  p,  and  volume,  v,  of  carbon 
dioxide  at  different  temperatures  T,  TQ,  Tv  T2,  ...  we  get  a  set 
of  curves  resembling  Fig.  90.     At  any  temperature  T  above  the 

critical  temperature,  the  relation  be- 
tween p  and  v  is  given  by  the  curve  pT. 
The  gas  will  not  liquefy.  Below  the 
critical  temperature,  say  Tv  the  volume 
decreases  as  the  pressure  increases,  as 
shown  by  the  curve  TlK1 ;  at  KY  the 
gas  begins  to  liquefy  and  the  pressure 
remains  constant  although  the  volume 
of  the  system  diminishes  from  KY  to 
Mv  At  M1  all  the  gas  will  have 
liquefied  and  the  curve  M^pY  will  repre- 
sent the  relation  between  the  pressure 
Similar  curves  T2K2M2P2,  TSK3MZP3, 


Pig.  90. 


and  volume  of  the  liquid. 


§  64.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  175 

...  are  obtained  at  the  different  temperatures  below  the  critical 
temperature  T0. 

The  lines  E0A,  KqP0  and  KJB  divide  the  plane  of  the  paper 
into  three  regions.  Every  point  to  the  left  of  AK0p0  represents  a 
homogeneous  liquid ;  every  point  to  the  right  of  BK0p0  represents  a 
homogeneous  gaseous  phase ;  while  every  point  in  the  region  AKQB 
represents  a  heterogeneous  liquid-gas  phase. 

By  gradually  increasing  the  pressure,  at  any  assigned  tempera- 
ture below  T0,  the  gas  will  begin  to  liquefy  at  some  point  along 
the  line  BK0 ;  this  is  called  the  dew  curYe — ligne  de  ros6e.  If  the 
pressure  on  a  liquid  whose  state  is  represented  by  a  point  in  the 
region  OAK^p,  be  gradually  diminished,  the  substance  will  begin 
to  assume  the  gaseous  state  at  some  point  along  the  line  K0A. 
This  is  called  the  boiling  curve — ligne  d' Ebullition.  At  K0  there 
is  a  tacnodal  point  or  double  cusp  of  the  mixed  species. 

A  remarkable  phenomenon  occurs  when  a  mixture  of  two  gases 
is  treated  in  a  similar  manner.  If  a  mixture  of  one  volume  of  air 
and  nine  volumes  of  carbon  dioxide  be  subjected  to  a  gradually 
increasing  pressure  at  about  2°  C,  the  gas  begins  to  liquefy  at  a 
pressure  of  72  atm. ;  and  on  increasing  the  pressure,  still  keeping 
the  temperature  constant,  the  liquid  again  passes  into  the  gaseous 
state  when  the  pressure  reaches  149  atm. ;  and  the  liquid  does  not 
reappear  again  however  great  the  pressure.  If  the  pressure  at 
which  the  liquid  appears  and  disappears  be  plotted  with  the  cor- 
responding temperature,  we  get  the  dew  curve 
BKC,  shown  in  Fig.  91.  For  the  same  ab- 
scissa Tv  there  are  two  ordinates,  pl  and  px't 
between  which  the  mixture  is  in  a  hetero- 
geneous condition.  At  temperatures  above 
T0,  no  condensation  will  occur  at  all ;  below 
Tx  only  normal  condensation  takes  place ;  at 
temperatures  between  TY  and  T0  both  normal  Fig.  91. 

and  retrograde  condensation  will  occur.  The  dotted  line  AG 
represents  the  boiling  curve;  above  AC  the  system  will  be  in  the 
liquid  state.  K  corresponds  with  the  critical  temperature  of  the 
mixture.     C  is  called  the  plait-point. 

The  phenomenon  only  occurs  with  mixtures  of  a  certain  com- 
position. Above  and  below  these  limits  the  dew  curves  are  quite 
normal.  This  is  shown  by  the  curves  DC  and  OC5  in  Fig.  92.  C 
is  the  plait-point ;  and  the  line  joining  the  plait-points  C2,  Cit  Cv  . . 


176  HIGHEK  MATHEMATICS.  §  65. 

for  different  mixtures  is  called  the  plait-point  curve.  The  dotted 
lines  in  the  same  diagram  represent  boiling  curves.  Note  the 
gradual  narrowing  of  the  border  curves  and  their  transit  into  or- 
dinary vapour  pressure  curves  DC  and  0G5 
at  the  two  extremes.  You  must  notice  that 
we  are  really  working  in  three  dimensions. 
The  variables  are  p,  v  and  T. 

The  plait-point   curve   appears  to  form 
a  double  cusp  of  the  second  species   at  a 
plait-point.     There  is  some  discussion  as  to 
Fig.  92.  whether,   say,   AGZKZB  really  forms  a  con- 

tinuous curve  line,  so  that  at  Gz  the  Hne  CGb  is  tangent  to  AC^KJB ; 
or  separate  lines  each  forming  a  spinode  or  cusp  with  the  line  CC5 
at  the  point  C3.  But  enough  has  been  said  upon  the  nature  of 
these  curves  to  carry  the  student  through  this  branch  of  mathema- 
tics in,  say,  J.  D.  van  der  Waals'  Bindre  Gemische,  Leipzig,  1900. 

Example. — Show  that  the  product  pv  for  van  der  Waals'  equation  fur- 
nishes a  minimum  when 


ab       _      \(      ab      \2  aW 

a-RbT      *sl{a-RbT    ~  a  -  BbT" 


Hint.  Multiply  the  first  of  equations  (3)  through  with  v ;  differentiate  to  get 
d(pv)/dv  =  0,  etc.  The  conclusion  is  in  harmony  with  M.  Amagat's  experi- 
ments (Ann.  Ghvm.  Phys.,  [5]  22,  353,  1881)  on  carbon  dioxide,  ethylene, 
nitrogen  and  methane.  For  hydrogen,  a,  in  (3),  is  negligibly  small,  hence 
show  that  pv  has  no  minimum. 

§  65.   Imaginary  Quantities. 

We  have  just  seen  that  no  number  is  known  which  has  a 
negative  value  when  multiplied  by  itself.  The  square  root  of  a 
negative  quantity  cannot,  therefore,  be  a  real  number.  In  spite  of 
this  fact,  the  square  roots  of  negative  quantities  frequently  occur  in 
mathematical  investigations.  Again,  logarithms  of  negative  num- 
bers, inverse  sines  of  quantities  greater  than  unity,  . . .,  cannot  have 
real  values.  These,  too,  sometimes  crop  up  in  our  work  and  we 
must  know  what  to  do  with  them. 

Let  J  -  a2  be  such  a  quantity.  If  -  a2  is  the  product  of  a2 
and  -  1,  +  n/-  a2  may  be  supposed  to  consist  of  two  parts,  viz., 
±  a  and  J  -  1.  Mathematicians  have  agreed  to  call  a  the  real 
part  of  J  -  a2  and  si  -  1,  the  imaginary  part.     Following  Gauss, 


$  65.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  177 

*J  -  1  is  written  t  (or  i).     It  is  assumed  that  J  -  1,  or  t  obeys  all 
the  rules  of  algebra.1     Thus, 

n/~1x  V^T=  -1;   */-"I-2*/Tl;   s/-~ax  n/5  =  J -ab;  i4  =  l. 

We  know  what  the  phrase  "  the  point  x,  y  "  means  If  one  or 
both  of  #  and  y  are  imaginary,  the  point  is  said  to  be  imaginary. 
An  imaginary  point  has  no  g'eometrical  or  physical  meaning  If 
an  equation  in  x  and  y  is  affected  with  one  or  more  imaginary  co- 
efficients, the  non-existent  graph  is  called  an  imaginary  curye ; 
while  a  similar  equation  in  x,  y  and  z  will  furnish  an  imaginary 
surface. 

Examples.— (1)  Show  &*  =  1 ;  i*»  +  *  =  »;  i*«  +  2  =  -  1 ;  t4"  +  3  =  -  t. 

(2)  Prove  that  a2  +  o2  =  (a  +  16)  (a  -  ib). 

(3)  The  quadratic  x2  +  bx  +  c  =  0,  has  imaginary  roots  only  when  62  -  4c 
is  less  than  zero  (5),  page  854.  If  o  and  )8  are  the  roots  of  this  equation,  show 
that  a  =  -  £6  +  $»  \/62~  4c  ;  and  $  =  -  \b  -  £*  \/&2  -  4c,  satisfy  the  equation. 

(4)  Show  (a  +  16)  (c  +  id)  m  (ac  -  bd)  +  (ad  +  bc)i. 

(5)  Show  by  multiplying  numerator  and  denominator  by  c  +  id  that 
a  +  ib      ac  -bd      be  +  ad 

c~^ld  ~  c2  +  d2  +  C2-}-^2  *' 

To  illustrate  the  periodic  nature  of  the  symbol  t,  let  us  suppose 
that  i  represents  the  symbol  of  an  operation  which  when  repeated 
twice  changes  the  sign  of  the  subject  of  the  operation,  and  when 
repeated  four  times  restores  the  subject  of  the  operation  to  its 
original  form  For  instance,  if  we  twice  operate  on  x  with  t,  we 
get  -  x,  or 

(  J  -  l)2x  =J-lxJ-lxx  =  -x;  and(\/-  Vfa  =  x, 
and  so  on  in  cycles  of  four.  If  the  imaginary  quantities  tx,  -  txt 
. . .  are  plotted  on  the  y-axis — axis  of  imaginaries — and  the  real 
quantities  x,  -  x,  . . .  on  the  rr-axis — axis  of  reals — the  operation 
of  i  on  a;  will  rotate  x  through  90°,  two  operations  will  rotate  x 
through  180°,  three  operations  will  rotate  x  through  270°,  and  four 
operations  will  carry  x  back  to  its  original  position. 


1  The  so-called  fundamental  laws  of  algebra  are  :  /.  The  law  of  association :  The 
number  of  things  in  any  group  is  independent  of  the  order.  //.  The  commutative  law : 
(a)  Addition.  The  number  of  things  in  any  number  of  groTips  is  independent  of  the 
order,  (b)  Multiplication.  The  product  of  two  numbers  is  independent  of  the 
order.  III.  The  distributive  law  :  (a)  Multiplication.  The  multiplier  may  be  distri- 
buted over  each  term  of  the  multiplicand,  e.g.,  m(a  +  b)  =  ma  +  mb.  (b)  Division. 
(a  •}-  b)jm  =  aim  +  b/m.  IV.  The  index  law:  (a)  Multiplication.  aman  =  ««  +  ». 
(ft)  Division.    am/an=a*»~". 

M 


178  HIGHER  MATHEMATICS.  §  60. 

We  shall  see  later  on  that  2t  sin  x  =  etx  -  e  - LX ;  hence,  if  x  =  tt, 
sin7r  =  0,  and  we  have 

el7r  —  e~lir  =  0;  or,  e4*"  =  e~lir, 
meaning  that  the  function  eix  has  the  same  value  whether  x  =  tt,  or 
a;  =  -  7r.     From  the  last  equation  we  get  the  remarkable  connec- 
tion between  the  two  great  incommensurables  it  and  e  discovered 
by  Euler : 

Example. — Show  x  =  xxl  =  xx  e2™  =  eio«x  +  2t7T.  This  means  that 
the  addition  of  2nr  to  the  logarithm  of  any  quantity  has  the  effect  of  multi- 
plying it  by  unity,  and  will  not  change  its  value.  Every  real  quantity  there- 
fore, has  one  real  logarithm  and  an  infinite  number  of  imaginary  logarithms 
differing  by  2inw,  where  n  is  an  integer. 

Do  not  confuse  irrational  with  imaginary  quantities.  Numbers 
like  \/2,  y  5,  . . .  which  cannot  be  obtained  in  the  form  of  a  whole 
number  or  finite  fraction  are  said  to  be  irrational  or  surd  num- 
bers. On  the  contrary,  JI,  \/%7,  . . .  are  rational  numbers.  Al- 
though we  cannot  get  the  absolutely  correct  value  of  an  irrational 
number,  we  can  get  as  close  an  approximation  as  ever  we  please ; 
but  we  cannot  even  say  that  the  imaginary  quantity  is  entitled 
to  be  called  a  quantity. 

§  66.    Curvature. 

The  curvature  at  any  point  of  a  plane  curve  is  the  rate  at  which 
the  curve  is  bending.  Of  two  curves  AG,  AD,  that  has  the  greater 
curvature  which  departs  the  more  rapidly  from  its  tangent  AB 
A  (Fig.  93).     In  passing  from  P  (Fig.  94)  to 

"b~  another  neighbouring  point  P1  along  any 


VC       arc  8s  of  the  plane  curve  AB,  the  tangent 
F      go  at  P  turns  through  the  angle  Ba,  where 

a  is  the  angle  made  by  the  intersection 
of  the  tangent  at  P  with  the  £-axis.  The  curvature  of  the  curve 
at  the  point  P  is  defined  as  the  limiting  value  of  the  ratio  8a/Ss 
when  P1  coincides  with  P.  When  the  points  P  and  Px  are  not 
infinitely  close  together,  this  ratio  may  be  called  the  mean  or 
average  curvature  of  the  curve  between  A  and  B.  We  might  now 
say  that 

H  =  -T-  =  Rate  of  bending  of  curve.  ,  .  (1) 

CLS 


§  66.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  179 

I.  The  curvature  of  the  circumference  of  all  circles   of  equal 

radius  is  the  same  at  all  points,  and  varies  inversely  as  the  radius. 

This  is  established  in  the  following 

way :  Let  AB  (Fig.  94)  be  a  part  of  a 

circle ;  Q,  the  centre ;  QP  =  QP1  = 

radius  =  B.    The  two  angles  marked 

8a  are  obviously  equal.      The  angle 

PQP1  is  measured  in  circular  measure, 

page  606,  by  the  ratio  of  the  arc  PPX 

to  the  radius,  i.e.,  the  angle  PQP-,  = 

1  Fig   94 

arc  PPJB;   or,   8a/Bs  =  1/R.      The 

curvature  of  a  circle  is  therefore  the  reciprocal  of  the  radius,  or,  in 

symbols, 

ds    B'        .....       ^; 

Example. — An  illustration  from  mechanics.  If  a  particle  moves  with  a 
variable  velocity  on  the  curve  AB  (Fig.  94)  so  that  at  the  time  t,  the  particle 
is  at  P,  the  particle  would,  by  Newton's  first  law  of  motion,  continue  to  move 
in  the  direction  of  the  tangent  PS,  if  it  were  not  acted  upon  by  a  central  force 
at  Q  which  compels  the  particle  to  keep  moving  on  the  curvilinear  path  PPXB. 
Let  Px  be  the  position  of  the  particle  at  the  end  of  a  short  interval  of  time  dt. 
The  direction  of  motion  of  the  particle  at  Px  may  similarly  be  represented  by 
the  tangent  P^.  Let  the  length  of  the  two  straight  lines  ap  and  apx  represent, 
in  direction  and  magnitude,  the  respective  velocities  of  the  particle  at  P  and 
at  Pr  Join  ppx.  The  angle  papx  is  evidently  equal  to  the  angle  5a.  Since 
ap  represents  in  direction  and  magnitude  the  velocity  of  the  particle  at  P, 
and  apv  the  velocity  of  the  particle  at  Px,  ppx  will  represent  the  increment 
in  the  velocity  of  the  particle  as  it  passes  from  P  to  Plt  for  the  parallelo- 
gram of  velocities  tells  us  that  apx  is  the  resultant  of  the  two  component 
velocities  ap  and  ppx,  in  direction  and  magnitude.  The  total  acceleration 
of  the  particle  in  passing  from  P  to  Px  is  therefore 

Total  acceleration  =  ^ocity  gained  =  m 
Time  occupied        dt 

Now  drop  a  perpendicular  from  the  point  p  to  meet  apx  at  ra.  The  in- 
finitely small  change  of  velocity  ppx  may  be  regarded  as  the  resultant  of  two 
changes  pm  and  pxm,  or  the  acceleration  ppx\dt  is  the  resultant  of  two  acceler- 
ations pm/dt  and  mpx/dt  represented  in  direction  and  magnitude  by  the  lines 
mp  and  pxm  respectively,  pm/dt  is  called  the  normal  acceleration.  pxm/dt,  the 
tangential  acceleration.  If  dt  be  made  small  enough,  the  direction  of  mp 
coincides  with  the  direction  of  the  normal  QP  to  the  tangent  of  the  curve 
at  the  point  P;  just  as  BPX  ultimately  coincides  with  SP  if  da  be  taken 
small  enough.  But  rap  =  op  sin  5a.  If  5a  is  small  enough,  we  may  write 
sin  5a  =  5a  (11),  page  602.  Let  V  denote  the  velocity  of  the  particle  at  the 
point  P,  then  rap  =  Vda.     From  (2),  5a  =  Ss/E  ;  and  5s/5£  =  V,  hence, 

M  * 


180  HIGHER  MATHEMATICS.  §  66. 

Normal  acceleration  =  -±-  =—.  —  =--. 
dt       R     dt       R 

That  is  to  say,  when  the  particle  moves  on  the  curve,  the  acceleration  in  the 

direction  of  the  normal  is  directly  proportional  to  the  square  of  the  velocity, 

and  inversely  as  the  radius  of  curvature.     Similarly  the 

dV 
Tangential  acceleration  =  -— -. 

Cut 

If  the  particle  moves  in  a  straight  line,  R  =  5a,  and  the  normal  acceleration 
is  zero. 

Just  as  a  straight  line  touching  a  curve,  may  be  regarded  as 
a  line  drawn  through  two  points  of  the  curve  infinitely  close  to 
each  other  (definition  of  tangent),  so  a  circle  in  contact  with  a 

curve  may  be  considered  to  pass 
through  three  consecutive  points 
of  the  curve  infinitely  near  each 
other.  Such  a  circle  is  called  an 
"osculatory  circle"  or  a  "circle 
of  curvature".  The  osculatory 
circle  of  a  curve  has  the  same 
Eig.  95.  curvature  as  the  curve  itself  at 

the  point  of  contact.  The  curvature  of  different  parts  of  a  curve 
may  be  compared  by  drawing  osculatory  circles  through  these 
points.  If  r  be  the  radius  of  an  osculatory  circle  at  P  (Fig.  95) 
and  rx  that  at  Pv  then 

Curvature  at  P   :  Curvature  at  P-,  =  -  :  — -  .     .     (3) 

1      r     rl  v  ' 

In  other  words,  the  curvature  at  any  two  points  on  a  curve  varies 
inversely  as  the  radius  of  the  osculatory  circles  at  these  points. 
The  radius  of  the  osculatory  circle  at  different  points  of  a  curve  is 
called  the  "radius  of  curvature"  at  that  point.  The  centre  of 
the  osculatory  circle  is  the  "centre  of  curvature  ". 

II.  To  find  the  radius  of  curvature  of  a  curve.  Let  the  co- 
ordinates of  the  centre  of  the  circle  be  a  and  b,  B  the  radius,  then 
the  equation  of  the  circle  is,  page  98, 

(x  -  a)2  +  (y  -  b)2  =  B2.  .         .         .         (4) 
Differentiating  this  equation  twice  ;  and,  dividing  by  2,  we  get 

(*-«)  +  (y-*)!  =  0;  and,  1  +  (J,  -  *)g+(|)2=  0.        (5) 

Let  u  =  dy/dx  and  v  =  d2y/dx2,  for  the  sake  of  ease  in  manipulation, 
(5)  then  becomes 


§  66.        FUNCTIONS  WITH  SINGULAR  PROPERTIES.  181 

1    ■   -'»                          1  +  w2  /Av 

and,  X-a  = U     .  .  (b) 


V  V 

by  substituting  for  y  -  b  in  the  first  of  equations  (5).  Now  u,  v,  x 
and  y  at  any  point  of  the  curve  are  the  same  for  both  the  curve 
and  the  osculating  circle  at  that  point,  and  therefore  a,  b x  and  B 
can  be  determined  from  x,  y,  u,  v.  By  substituting  equation  (6) 
in  (4),  we  get 

ii     • .  (7) 

r-  j(i  +  u*y y> 

The  standard  equation  for  the  radius  of  curvature  at  the  point 
(x,  y)  is 

1=4  = ^;or,JUPf  (8) 

B      ds     c         /<fy\2U  3^  V 

I1  +  W  J  ^ 

When  the  curve  is  but  slightly  inclined  to  the  #-axis,  dy/dx 

is  practically  zero,  and  the  radius  of  curvature  is  given  by  the 

expression 

B=^y (9) 

dx*  N 

a  result  frequently  used  in  physical  calculations  involving  capil- 
larity, superficial  tension,  theory  of  lenses,  etc. 

III.  The  direction  of  curvature  has  been  discussed  in  §  60. 
It  was  there  shown  that  a  curve  is  concave  or  convex  upwards  at 
a  point  (x,  y)  according  as  d2y/dx2  >  or  <  0. 

Examples. — (1)  Find  the  radius  of  curvature  at  any  point  (x,  y)  on  the 

ellipse 

xl    t.-i       .  <ty__&x.  d?y_  _V_     n_     (aY  +  bWfi 
aa+62  "  dx~     a?y'  dx2~     a*y*  a464 

At  the  point  x  =  a,  y  =  0,  R  =  b2/a.     Hint.  The  steps  for  dhf\dx*  are  : 

gty       6*  6=    y-x.dyjdx_     o2    aV  +  W        62    a262 

da2  ~  afy'6  ~     a*  '         y3        ~     a? '       a?y3      ~     a*'  y*  ' 
(2)  The  radius  of  curvature  on  the  curve  xy  =  a,  at  any  point  (x,  y)>  is 
(a;2  +  y2)%l2a. 

1  The  determination  of  a  and  b  is  of  little  use  in  practical  work.  They  give 
equations  to  the  evolute  of  the  curve  under  consideration.  The  evolute  is  the  curve 
drawn  through  the  centres  of  the  osculatory  circles  at  every  part  of  the  curve,  the 
curve  itself  is  called  the  involute.  Example  :  the  osculatory  circle  has  the  equation 
(x  -  a)2  +  {y  -  bf  —  R.  a  and  b  may  be  determined  from  equations  (4),  (7)  and  (8). 
The  evolate  of  the  parabola  y-  =  mx  is  27my'2  =  8(2&  -  7ii)'K 


182 


HIGHER  MATHEMATICS. 


§67, 


The  equation 


§  67.  Envelopes. 


m 
y  =  — h  ax. 
u       a 


represents  a  straight  line  cutting  the  ?/-axis  at  m/a,  and  making  an 
angle  tan  ~  la  with  the  sc-axis.  If  a  varies  by  slight  increments,  the 
equation  represents  a  series  of  straight  lines  so  near  together  that 
their  increments  may  be  considered  to  lie  upon  a  continuous  curve. 
a  is  said  to  be  the  variable  parameter  of  the  family,  since  the 
different  members  of  the  family  are  obtained  by  assigning  arbitrary 
values  for  a.     Let  the  equations 


m 
Vl=~  +  ax 


#2  = 


2/3  = 


m 


a  +  la 

m 
a  +  2la 


+  (a  +  la)x 
+  (a  +  2Sa)x 


a) 

(2) 
(3) 


be  three  successive  members  of  the  family.     As  a  general  rule  two 

distinct  curves  in  the  same  family 
will  have  a  point  of  intersection.  Let 
P  (Fig.  96)  be  the  point  of  intersection 
of  curves  (1)  and  (2) ;  P1  the  point  of 
intersection  of  curves  (2)  and  (3), 
then,  since  P2  and  P2  are  both  situ- 
ated on  the  curve  (2),  PPY  is  part  of 
the  locus  of  a  curve  whose  arc  PPX 
coincides  with  an  equal  part  of  the 
curve  (2).  It  can  be  proved,  in  fact, 
that  the   curve   PPl  . . .  touches  the 

whole  family  of  curves  represented  by  the  original  equation.     Such 

a  curve  is  said  to  be  an  envelope  of  the  family. 

To  find  the  equation  of  the  envelope,  bring  all  the  terms  of  the 

original  equation  to  one  side, 


Fig.  96.— Envelope. 


y 


m 
a 


ax 


0. 


Then  differentiate  with  respect  to  the  variable  parameter,  and  put 
mda 


w-xda  =  0;  .-.-2 


x  =  0. 


Eliminate  a  between  these  equations, 


§  67.         FUNCTIONS  WITH  SINGULAR  PROPERTIES.  183 


J 


m .  x  -  x  a/—  =  0,  or  y  -  2  Jm  .x  =  Q. 


r 


=  Amx. 


Examples. — (1)  Find  the  envelope  of  the  family  of  circles 


(x  -  af 
where  a  is  the  variable  parameter, 
to  0,  and  we  get  aj-a=0;  then 
eliminating  a,  we  get  y  =  +  r,  which 
is  the  required  envelope.  The  en- 
velope y=*  ±r  represents  two  straight 
lines  parallel  to  the  aj-axis,  AB,  and 
at  a  distance  +  r  and  -  r  from  it. 
Shown  Fig.  97. 

(2)  Show  that  the  envelope  of  the 
family  of  curves  xja  +  y/fi  =  1,  where 
a  and  0  are  variable  parameters  sub- 
ject to  the  condition  that  a)8  =  4m2, 
is    the  hyperbola  xy  =  to2.      Hint. 


+  V  =  r\ 
Differentiate  with  respect  to  a,  equate 

enveiope 


Fig.  97. 


envelop?  F 

-Double  Envelope. 


Differentiate  each  of  the  given  equations  with  respect  to  the  given  para- 
meters, and  we  get  xda/a*  +  ydpftP  =  0  from  the  first,  and  &da  +  adp  =  0, 
from  the  second.  Eliminate  da  and  dp.  Hence  x/a  =  yj0  =  $  ;  .'.  a  =  2x; 
/8  m  2y.     Substitute  in  oj8  sa  4m2,  etc. 

If  a  given  system  of  rays  be  incident  upon  a  bright  curve,  the 
envelope  of  the  reflected  rays  is  called  a  caustic  by  reflection. 


CHAPTER  IV. 

THE  INTEGBAL  CALCULUS. 

"  Mathematics  may  be  defined  as  the  economy  of  counting.  There  is 
no  problem  in  the  whole  of  mathematics  which  cannot  be  solved 
by  direct  counting.  But  with  the  present  implements  of  mathe- 
matics many  operations  of  counting  can  be  performed  in  a  few 
minutes,  which,  without  mathematics,  would  take  a  lifetime." — 
E.  Mach. 

§  68.  The  Purpose  of  Integration. 

In  the  first  chapter,  methods  were  described  for  finding  the  mo- 
mentary rate  of  progress  of  any  uniform  or  continuous  change  in 
terms  of  a  limiting  ratio,  the  so-called  "differential  coefficient" 
between  two  variable  magnitudes.  The  fundamental  relation 
between  the  variables  must  be  accurately  known  before  one  can 
form  a  quantitative  conception  of  the  process  taking  place  at  any 
moment  of  time.  When  this  relation  or  law  is  expressed  in  the 
form  of  a  mathematical  equation,  the  "methods  of  differentiation" 
enable  us  to  determine  the  character  of  the  continuous  physical 
change  at  any  instant  of  time.  These  methods  have  been 
described. 

Another  problem  is  even  more  frequently  presented  to  the 
investigator.  Knowing  the  momentary  character  of  any  natural 
process,  it  is  asked :  "  What  is  the  fundamental  relation  between 
the  variables?"  "What  law  governs  the  whole  course  of  the 
physical  change?" 

In  order  to  fix  this  idea,  let  us  study  an  example.  The  con- 
version of  cane  sugar — CJ12H22On — into  invert  sugar — C6H1206 — 
in  the  presence  of  dilute  acids,  takes  place  in  accordance  with  the 
reaction : 

^12-^22^11  +  -E-2O  =  2C6H1206. 
Let  x  denote  the  amount  of  invert  sugar  formed  in  the  time  t; 
the  amount  of  sugar  remaining  in  the  solution  will  then  be  1  -  x, 

184 


§  68.  THE  INTEGRAL  CALCULUS.  185 

provided  the  solution  originally  contained  one  gram  of  cane  sugar. 
The  amount  of  invert  sugar  formed  in  the  time  dt,  will  be  dx.  From 
the  law  of  mass  action,  "  the  velocity  of  the  chemical  reaction 
at  any  moment  is  proportional  to  the  amount  of  cane  sugar  actually 
present  in  the  solution  ".     That  is  to  say, 

|-*a-«),  ....     a) 

where  k  is  the  "constant  of  proportion,"  page  23.  The  meaning 
of  h  is  obtained  by  putting  a;  =  0.  Thus,  dxjdt  =  k,  or,  k  denotes 
the  rate  of  transformation  of  unit  mass  of  sugar,  or 

'-£ <2) 

where  V  denotes  the  velocity  of  the  reaction.  This  relation  is 
strictly  true  only  when  we  make  the  interval  of  time  so  short 
that  the  velocity  has  not  had  time  to  vary  during  the  process. 
But  the  velocity  is  not  really  constant  during  any  finite  interval 
of  time,  because  the  amount  of  cane  sugar  remaining  to  be  acted 
upon  by  the  dilute  acid  is  continually  decreasing.  For  the  sake 
of  simplicity,  let  k  =  x\p  and  assume  that  the  action  takes  place  in 
a  series  of  successive  stages,  so  that  dx  and  dt  have  finite  values, 
say  Sx  and  St  respectively.     Then, 

y       Amount  of  cane  sugar  transformed  _  Sx 

Internal  of  time  St'  '  '  ^   ' 

Let  St  be  one  second  of  time.  Let  ^  of  the  cane  sugar  present 
be  transformed  into  invert  sugar  in  each  interval  of  time,  at  the 
same  uniform  rate  that  it  possessed  at  the  beginning  of  the  interval. 
At  the  commencement  of  the  first  interval,  when  the  reaction 
has  just  started,  the  velocity  will  be  at  the  rate  of  0100  grams  of 
invert  sugar  per  second.  This  rate  will  be  maintained  until  the 
commencement  of  the  second  interval,  when  the  velocity  suddenly 
slackens  down,  because  only  0*900  grams  of  cane  sugar  are  then 
present  in  the  solution. 

During  the  second  interval,  the  rate  of  formation  of  invert 
sugar  will  be  ^  of  the  0*900  grams  actually  present  at  the  be- 
ginning. Or,  0*090  grams  of  invert  sugar  are  formed  during  the 
second  interval. 

At  the  beginning  of  the  third  interval,  the  velocity  of  the  re- 
action is  again  suddenly  retarded,  and  this  is  repeated  every  second 
for  say  five  seconds. 


186  HIGHER  MATHEMATICS.  §  68. 

Now  let  Sxv  $x2, . . .  denote  the  amounts  of  invert  sugar  formed 
in  the  solution  during  each  second,  U.  Assume,  for  the  sake  of 
simplicity,  that  one  gram  of  cane  sugar  yields  one  gram  of  invert 

sugar. 


Cane 

i  sugar  transformed. 

During  the  1st  second, 

,  Sx1  =  0-100 

„    2nd 

>> 

Sx2  =  0-090 

„    3rd 

M 

Sx3  =  0-081 

„         „    4th 

»> 

8o34  ~  0073 

it        n    5th 

>J 

Sx5  =  0-066 

Total,         0-410 

This  means  that  if  the  chemical  reaction  proceeds  during  each 
successive  interval  with  a  uniform  velocity  equal  to  that  which  it 
possessed  at  the  commencement  of  that  interval,  then,  0410  gram 
of  invert  sugar  would  be  formed  at  the  end  of  five  seconds.  As  a 
matter  of  fact,  0*3935  gram  is  formed. 

But  0-410  gram  is  evidently  too  great,  because  the  retardation 
is  a  uniform,  not  a  jerky  process.  We  have  resolved  it  into  a 
series  of  elementary  stages  and  pretended  that  the  rate  of  forma- 
tion of  invert  sugar  remained  uniform  during  each  elementary 
stage.  We  have  ignored  the  retardation  which  takes  place  from 
moment  to  moment.  If  we  shorten  the  interval  and  determine 
the  amounts  of  invert  sugar  formed  during  intervals  of  say  half  a 
second,  we  shall  have  ten  instead  of  five  separate  stages  to  sum 
up,  thus  : 

Cane  sugar  transformed. 
During  the  1st  half  second,  5xl  =  0-0500 


2nd 

ii 

Sx2 

=  0-0475 

3rd 

„ 

Sx3 

=  0-0451 

4th 

ii 

8z4 

=  00429 

5th 

,, 

Sx5 

=  0-0407 

6th 

,, 

Sx6 

=  0-0387 

7th 

>i 

8oj7 

=  0-0367 

8th 

n 

8z8 

=  0-0349 

9th 

11 

5x9 

=  0-0332 

10th 

II 

5x10 

=  0-0315 

Total,        0-4012 

The  quantity  of  invert  sugar  calculated  on  the  supposition 
that  the  velocity  is  retarded  every  half  second  instead  of  every 
second,  corresponds  more  closely  with  the  actual  change.  The 
smaller  we  make  the  interval  of  time  the  more  accurate  the  result. 
Finally,  by  making  U  infinitely  small,  although  we  should  have 


§68. 


THE  INTEGRAL  CALCULUS. 


187 


an  infinite  number  of  equations  to  add  up,  the  actual  summation 
would  give  a  perfectly  accurate  result.  To  add  up  an  infinite 
number  of  equations  is,  of  course,  an  arithmetical  impossibility, 
but,  by  the  "methods  of  integration"  we  can  actually  perform 
this  operation. 

X  =  Sum  of  all  the  terras    V  .  dt,   between  ^=0,  and   t  m  5  ; 
.'.  X  =  V  .dt  +   V  .dt  +  V.dt  +.  .  .to  infinity. 
This  is  more  conveniently  written, 

5  f5 

X  =  2,  (V  .dt)  ;   or,  better  still,  X  =  I    V.  dt. 

Jo 
The  signs  "  2  "  and  u[ "  are  abbreviations  for  "  the  sum  of  all 

the  terms  containing .  . .  " ;  the  subscripts  and  superscripts  denote 


the  limits  between  which  the  time  has  been  reckoned.  The  second 
member  of  the  last  equation  is  called,  on  Bernoulli's  suggestion, 
an  integral.  "jf(x).dx"  is  read  "the  integral  of  f(x).dx". 
When  the  limits  between  which  the  integration  (evidently  another 
word  for  "  summation ")  is  to  be  performed,  are  stated,  the 
integral  is  said  to  be  definite ;  when  the  limits  are  omitted,  the 
integral  is  said  to  be  indefinite.  The  superscript  to  the  symbol 
u  J "  is  called  the  upper  or  superior  limit ;  the  subscript,  the 
lower  or  inferior  limit.  For  example,  JJjp .  dv  denotes  the  sum 
of  an  infinite  number  of  terms  p .  dv,  when  v  is  taken  between  the 
limits  v2  and  vv  In  order  that  the  "  limit  "  of  integration  may  not 
be  confounded  with  the  "limiting  value"  of  a  function,  some 
writers  call  the  former  the  "  end-values  of  the  integral ". 

To  prevent  any  misunderstanding,  I  will  now  give  a  graphic 


188  HIGHER  MATHEMATICS.  §68. 

representation  of  the  above  process.  Take  Ot  and  Ov  as  co- 
ordinate axes  (Figs.  98  and  99).  Mark  off,  along  the  abscissa  axis, 
intervals  1,  2,  3, . . .  ,  corresponding  to  the  intervals  of  time  St. 
Let  the  ordinate  axis  represent  the  velocities  of  the  reaction 
during  these  different  intervals  of  time.  Let  the  curve  vbdfh . . . 
represent  the  actual  velocity  of  the  transformation  on  the  supposi- 
tion that  the  rate  of  formation  of  invert  sugar  is  a  uniform  and 
continuous  process  of  retardation.  This  is  the  real  nature  of  the 
change.  But  we  have  pretended  that  the  velocity  remains  con- 
stant during  a  short  but  finite  interval  of  time  say  St  =  1  second. 
The  amount  of  cane  sugar  inverted  during  the  first  second  is, 


(y    0-j  to  i>a  2>o  2-5  so  3\S  *~o  *■£  $•&  second* 
Fig.  99. 

therefore,  represented  by  the  area  valO  (Fig.  98) ;  during  the 
second  interval  by  the  area  bc21,  and  so  on. 

At  the  end  of  the  first  interval  the  velocity  at  a  is  supposed 
to  suddenly  fall  to  b,  whereas,  in  reality,  the  decrease  should  be 
represented  by  the  gradual  slope  of  the  curve  vb. 

The  error  resulting  from  the  inexact  nature  of  this  "  simplifying 
assumption  "  is  graphically  represented  by  the  blackened  area  vab  ; 
for  succeeding  intervals  the  error  is  similarly  represented  by  bed, 
def, ...  In  Fig.  99,  by  halving  the  interval,  we  have  considerably 
reduced  the  magnitude  of  the  error.  This  is  shown  by  the  dimin- 
ished area  of  the  blackened  portions  for  the  first  and  succeeding 
seconds  of  time.  The  smaller  we  make  the  interval,  the  less  the 
error,  until,  at  the  limit,  when  the  interval  is  made  infinitely 
small,  the  result  is  absolutely  correct.     The  amount  of  invert  sugar 


§  68.  THE  INTEGRAL  CALCULUS.  189 

formed  during  the  first  five  seconds  is  then  represented  by  the  area 
vbdf...§0. 

The  above  reasoning  will  repay  careful  study  ;  once  mastered, 
the  "  methods  of  integration  "  are,  in  general,  mere  routine  work. 

The  operation1  denoted  by  the  symbol  "  J"  is  called  integra- 
tion. When  this  sign  is  placed  before  a  differential  function,  say 
dx,  it  means  that  the  function  is  to  be  integrated  with  respect  to 
dx.  Integration  is  essentially  a  method  for  obtaining  the  sum.  of 
an  infinite  number  of  infinitely  small  quantities.  This  does  not 
mean,  as  some  writers  have  it,  "  if  enough  nothings  be  taken  their 
sum  is  something".  The  integral  itself  is  not  exactly  what  we 
usually  understand  by  the  term  "  sum,"  but  it  is  rather  "  the  limit 
of  a  sum  when  the  number  of  terms  is  infinitely  great ". 

Not  only  can  the  amount  of  substance  formed  in  a  chemical 
reaction  during  any  given  interval  of  time  be  expressed  in  this 
manner,  but  all  sorts  of  varying  magnitudes  can  be  subject  to  a 
similar  operation.  The  distance  passed  over  by  a  train  travelling 
with  a  known  velocity,  can  be  represented  in  terms  of  a  definite 
integral.  The  quantity  of  heat  necessary  to  raise  the  temperature, 
0,  of  a  given  mass,  m,  of  a  substance  from  0>1  to  0°2,  is  given  by 
the  integral  f^ma- .  dO,  where  o-  denotes  the  specific  heat  of  the 
substance.  The  work  done  by  a  variable  force,  F,  when  a  body 
changes  its  position  from  s0  to  sx  is  j'tl0F .  ds.  This  is  called  a  space 
integral.  The  impulse  of  a  variable  force  F,  acting  during  the 
interval  of  time  t2  -  tv  is  given  by  the  time  integral  ftF  .dt.  By 
Newton's  second  law,  "  the  change  of  momentum  of  any  mass,  m, 
is  equal  to  the  impulse  it  receives  ".  Momentum  is  defined  as  the 
product  of  the  mass  into  the  velocity.  If,  when  t  is  tv  v  =  v1; 
and,  when  t  is  t2,  v  =  v2,  Newton's  law  may  be  written 


I  m.dv  =      F .dt. 


The  quantity  of  heat  developed  in  a  conductor  during  the 
passage  of  an  electric  current  of  intensity  0,  for  a  short  interval 
of  time  dt  is  given  by  the  expression  kO .dt  (Joule's  law),  where  k 
is  a  constant  depending  on  the  nature  of  the  circuit.  If  the  current 
remains  constant  during  any  short  interval  of  time,  the  amount  of 


1  The  symbol  "  J  "  is  supposed  to  be  the  first  letter  of  the  word  "sum  ".  "  Omn," 
from  omnia,  meaning  '/all,"  was  once  used  in  place  of  "J".  The  first  letter  of  the 
differential  dx  is  the  initial  letter  of  the  word  "  difference". 


190  HIGHER  MATHEMATICS.  §  68. 

heat  generated  by  the  current  during  the  interval  of  time  t2  -  tv 
is  given  by  the  integral  jffiC.dt.  The  quantity  of  gas,  q,  con- 
sumed in  a  building  during  any  interval  of  time  t2  -  tv  may  be 
represented  as  a  definite  integral, 


=  [\.dt, 


where  v  denotes  the  velocity  of  efflux  of  the  gas  from  the  burners. 
The  value  of  q  can  be  read  off  on  the  dial  of  the  gas  meter  at  any 
time.  The  gas  meter  performs  the  integration  automatically. 
The  cyclometer  of  a  bicycle  can  be  made  to  integrate, 


=  j  \dt. 


Differentiation  and  integration  are  reciprocal  operations  in  the 
same  sense  that  multiplication  is  the  inverse  of  division,  addition 
of  subtraction.     Thus, 

a  x  b  +  b  =  a;  a  +  b  -  b  =  a;  J  a2  =  a ; 
dja .dx  =  a.dx;  \dx  =  x  ; 

Bx2dx  is  the  differential  of  #3,  so  is  x3  the  integral  of  3x2dx.  The 
differentiation  of  an  integral,  or  the  integration  of  a  differential 
always  gives  the  original  function.  The  signs  of  differentiation 
and  of  integration  mutually  cancel  each  other.  The  integral, 
\f(x)dx,  is  sometimes  called  an  anti-differential.  Integration 
reverses  the  operation  of  differentiation  and  restores  the  differ- 
entiated function  to  its  original  value,  but  with  certain  limitations 
to  be  indicated  later  on. 

While  the  majority  of  mathematical  functions  can  be  differenti- 
ated without  any  particular  difficulty,  the  reverse  operation  of 
integration  is  not  always  so  easy,  in  some  cases  it  cannot  be  done 
at  all.  If,  however,  the  function  from  which  the  differential  has 
been  derived,  is  known,  the  integration  can  always  be  performed. 
Knowing  that  d  (log  x)  =  x  ~  l  dx,  it  follows  at  once  that 
jx~1dx  =  log  x.  The  differential  of  xn  is  nxH~1dx,  hence 
}nxn~ldx  =  xn.  In  order  that  the  differential  of  xn  may  assume 
the  form  of  x~\  we  must  have  n  -  1  =  -  1,  or  n  =  0.  In  that 
case  xn  becomes  x°  =  1.  This  has  no  differential.  The  algebraic 
function  xn  cannot  therefore  give  rise  to  a  differential  of  the  form 
x~xdx.  Nor  can  any  other  known  function  except  logic  give  rise 
to-x~ldx.  If  logarithms  had  not  been  invented  we  could  not  have 
integrated  fx~ldx.     The  integration  of  algebraic  functions  may 


§  68.  THE  INTEGRAL  CALCULUS.  191 

also  give  rise*  to  transcendental  functions.  Thus,  (1  -  x)  ~  hdx 
becomes  sin-1a;;  and  (1  +  x2)~1dx  becomes  tan-1#.  Still 
further,  the  integration  of  many  expressions  can  only  be  effected 
when  new  functions  corresponding  with  these  forms  have  been 
invented.  The  integrals  jex2 .dx,  and  j(xs  +  l)-$dx,  for  example, 
have  not  yet  been  evaluated,  because  we  do  not  know  any  function 
which  will  give  either  of  these  forms  when  differentiated. 

The  nature  of  mathematical  reasoning  may  now  be  denned 
with  greater  precision  than  was  possible  in  §  1.  There,  stress 
was  laid  upon  the  search  for  constant  relations  between  observed 
facts.  But  the  best  results  in  science  have  been  won  by  antici- 
pating Nature  by  means  of  the  so-called  working  hypothesis.  The 
investigator  first  endeavours  to  reproduce  his  ideas  in  the  form  of 
a  differential  equation  representing  the  momentary  state  of  the 
phenomenon.  Thus  Wilhelmy's  law  (1850)  is  nothing  more  than 
the  mathematician's  way  of  stating  an  old,  previously  unverified, 
speculation  of  Berthollet  (1779) ;  while  Guldberg  and  Waage's  law 
(1864-69)  is  still  another  way  of  expressing  the  same  thing. 

To  test  the  consequences  of  Berthollet's  hypothesis,  it  is  clearly 
necessary  to  find  the  amount  of  chemical  action  taking  place  during 
intervals  of  time  accessible  to  experimental  measurement.  It  is 
obvious  that  Wilhelmy's  equation  in  its  present  form  will  not  do, 
but  by  "methods  of  integration  "  it  is  easy  to  show  that  if 

&x  .  .       1    .         1 

m  =  &(1  -  X),  then,  k  =  y  log  j— j, 

where  x  denotes  the  amount  of  substance  transformed  during  the 
time  t.  x  is  measurable, '£  is  measurable.  We  are  now  in  a  posi- 
tion to  compare  the  fundamental  assumption  with  observed  facts. 
If  Berthollet's  guess  is  a  good  one,  k,  above,  must  have  a  con- 
stant value.  But  this  is  work  for  the  laboratory,  not  the  study, 
as  indicated  in  connection  with  Newton's  law  of  cooling,  §  20. 

Integration,  therefore,  bridges  the  gap  between  theory  and  fact 
by  reproducing  the  hypothesis  in  a  form  suitable  for  experimental 
verification,  and,  at  the  same  time,  furnishes  a  direct  answer  to  the 
two  questions  raised  at  the  beginning  of  this  section.  The  idea  was 
represented  in  my  Chemical  Statics  and  Dynamics  (1904),  thus: — 
Hypothesis  — >  Differential  Equation  — >  Integration  — >  Observation. 

We  shall  return  to  the  above  physical  process  after  we  have  gone 
through  a  drilling  in  the  methods  to  be  employed  for  the  integration 
of  expressions  in  which  the  variables  are  so  related  that  all  the  x's 
and  dx's  can  be  collected  to  one  side  of  the  equation,  all  the  y's  and 


192  HIGHER  MATHEMATICS.  §  70. 

dy's  to  the  other.  In  a  later-chapter  we  shall  have  to  study  the  in- 
tegration of  equations  representing  more  complex  natural  processes. 
If  the  mathematical  expression  of  our  ideas  leads  to  equations 
which  cannot  be  integrated,  the  working  hypothesis  will  either 
have  to  be  verified  some  other  way,1  or  else  relegated  to  the  great 
repository  of  unverified  speculations. 

§  69.  Table  of  Standard  Integrals. 

Every  differentiation  in  the  differential  calculus,  corresponds 
with  an  integration  in  the  integral  calculus.  Sets  of  corresponding 
functions  are  called  "  Tables  of  Integrals  ".  Table  II.,  page  193, 
contains  the  more  important ;  handy  for  reference,  better  still  for 
memorizing. 

§  70.    The  Simpler  Methods  of  Integration. 

I.  Integration  of  the  product  of  a  constant  term  and  a  differ- 
ential. On  page  38,  it  was  pointed  out  that  "  the  differential  of 
the  product  of  a  variable  and  a  constant,  is  equal  to  the  constant 
multiplied  by  the  differential  of  the  variable  ".  It  follows  directly 
that  the  integral  of  the  product  of  a  constant  and  a  differential,  is 
equal  to  the  constant  multiplied  by  the  integral  of  the  differential. 
E.g.,  if  a  is  constant, 

fa  .  dx  —  ajdx  =  ax ;  /log  a .  dx  =  log  ajdx  =  x  .  log  a. 
On  the  other  hand,  the  value  of  an  integral  is  altered  if  a  term 
containing  one  of  the  variables  is  placed  outside  the  integral  sign. 
For  instance,  the  reader  will  see  very  shortly  that  while 

jx2dx  =  \xz ;  xjxdx  =  \xz. 

II.  A  constant  term  must  be  added  to  every  integral.  It  has 
been  shown  that  a  constant  term  always  disappears  from  an 
expression  during  differentiation,  thus, 

d(x  +  C)  =  dx. 
This   is    equivalent  to   stating   that   there  is  an  infinite   number 
of  expressions,  differing  only  in  the  value  of  the  constant  term, 
which,   when   differentiated,    produce   the   same   differential.      In 

1  Say,  by  slipping  in  another  " simplifying  assumption".  Clair aut  expressed  his 
ideas  of  the  moon's  motion  in  the  form  of  a  set  of  complicated  differential  equations, 
but  left  them  in  this  incomplete  stage  with  the  invitation,  "Now  integrate  them  who 


§70. 


THE  INTEGRAL  CALCULUS. 


193 


Table  II. — Standard  Integrals. 


Function. 


Differential  Calculus. 


Integral  Calculus. 


u  =  Xn. 
u  =  ax. 
u  =  ex. 
u  =  log^. 
u  =  sin  x. 
u  =  cos  a;. 
u  =  tan  x. 
u  =  cot  a;. 
u  =  sec  x. 
u  =  cosec  x. 
u  =  sin  -  1jc. 
u  =  cos  ~  ^x. 
u  =  tan-1cc. 
u  =  cot-1®. 
u  =  sec-1a;. 
u  =  cosec  ~  lx. 
u  =  vers  ~  1£c. 
u  =  covers  ~  *x. 


du 

dx 

du 

dx~  =  aXl°Z°a' 

du 

du  _  1 

dx  ~  x' 

du 

dx  =  cosx- 

du 

dx  =  ~8inX' 

du 

dx 

du 

dx 

du  _      since 

die  ~~      cos2jc 

du  _      cos  a; 

~dx  ~      sin2^' 

du  1 


=      sec2^. 
=  -  cosec2a;. 


dx 
du 

V(l  "  *2)" 
1 

dx  ~ 
du 

x/(l  "  *2) 

1         <| 

dx 
du 

1+/2" 

dx 
du 

1  +  a;2    J 

1 

dx 
du 

*V(x2  -  1) 
1 

dx 
du 

XsJiX*-  1) 
1 

dx 
du 

>J(2x  -  a2) 
1 

dx         J{2x  -  x2) 


x^dx 
a*dx 


\exdx 
rdx 
J'* 


x 

cos  axdx 

sin  axdx 

sec2axdx      •■ 

cosetfax.dx. 


/ 
/' 
/' 
/ 

fsin  x , 
]c^xdx        " 


xn  +  l 
n  +  1 

a* 
logea 

e*.     . 

logeOJ. 

sinqa? 

a    ' 
cos  ace 


a 
tan  ace 
a     ' 
cot  ax 


cos2cc 
rcosa; 
sin2a; 

dx 


/cosec 
.  J  sin2a; 

/; 


s/(a*-x*)    = 


/•    da; 
J  a2  +  a:2 

[__dx__ 

r    dx    r= 


(1) 
(2) 

(3) 

(4) 

(5) 
(6) 
(7) 
(8) 

(9) 
(10) 
(11) 
(12) 
(13) 
(14) 
(15) 

a*16* 
vers-1a:.   .      (17) 

-  covers ~lx.  (18) 


sec  a;. 
-  cosec  a;. 

,35 


l— 


1.  ix 
-tan-1-- 
a  a 

1  x 

a  a 

1       _,* 

— sec  x-« 
a  a 


•cosec 


stating  the  result  of  any  integration,  therefore,  we  must  provide 
for  any  possible  constant  term,  by  adding  on  an  undetermined, 
11  empirical,"  or  u  arbitrary  "  constant,  called  the  constant  of 
integration,  and  usually  represented  by  the  letter  G.     Thus, 

jdu  =  u  +  0. 
If  we  are  given 

dy  =  dx, 

N 


194  HIGHER  MATHEMATICS.  §  70. 

or,  if  we  put  C=G2-  Gv  we  get  \dy  +  C1  =  jdx  +  C2 ;  y+C1  =  x  +  G2; 

y  =  x  +  C. 
The  geometrical  signification  of  this  constant  is  analogous  to 
that  of  "  b"  in  the  tangent  form  of  the  equation  of  the  straight 
line,  formula  (8),  page  94  ;  thus,  the  equation 

y  =  mx  +  bf 
represents  an  infinite  number  of  straight  lines,  each  one  of  which 
has  a  slope  m  to  the  #-axis  and  cuts  the  y-a>xis  at  some  point  b. 
An  infinite  number  of  values  may  be  assigned  to  b.      Similarly, 
an  infinite  number  of  values  may  be  assigned  to  G  in  J  . . .  dx  +  G. 

Example. — Find  a  curve  with  the  slope,  at  any  point  (x,  y),  of  2x  to  the 
oj-axis.  Since  dyjdx  is  a  measure  of  the  slope  of  the  curve  at  the  point  (a;,  y), 
dyjdx  =  2x  ;  .*.  y  =  x2  +  C.  If  G  =  0,  we  have  the  curve  y  =  x2 ;  if  C  =  1, 
another  curve,  y  =  x2  +  1 ;  if  G  =  S,  y  =  x2  +  3  . ..  In  the  given  problem  we 
do  not  know  enough  to  be  able  to  say  what  particular  value  G  ought  to  possess. 

According  to  (5),  (6),  (7),  (8),  Table  II. ,  which  are  based  upon 
(1),  (2),  (3),  (4),  page  48, 

■  =  sin  ~  lx  =  -  cos  ~lx  ;   I     ,  =  tan  ~  lx  =  -  cot  ~~  1x, 


y;.J- 


J  Jl-x2  '  J  Jl  +  , 

etc.  This  means  that  sin  -  xx,  cos  ~lx;  or  tan  ~  xx,  cot  ~  lx, . . .  only 
differ  by  a  constant  term.  This  agrees  with  the  trigonometrical 
properties  of  these  functions  illustrated  on  page  48.  The  following 
remarks  are  worth  thinking  over  : 

"  Fourier's  theorem  is  a  most  valuable  tool  of  science,  practical  and  theo- 
retical, but  it  necessitates  adaptation  to  any  particular  case  by  the  provision  of 
exact  data,  the  use,  that  is,  of  definite  figures  which  mathematicians  humorously 
call  '  constants,'  because  they  vary  with  every  change  of  condition.  A  simple 
formula  is  n  +  n  =  2n,  so  also  n  x  n  =  n2.  In  the  concrete,  these  come  to  the 
familiar  statement  that  2  and  2  equals  4.  So  in  the  abstract,  40  +  40  =  80. 
but  in  the  concrete  two  40  ft.  ladders  will  in  no  way  correspond  to  one  80  ft, 
ladder.  They  would  require  something  else  to  join  them  end  to  end  and  to 
strengthen  them.  That  something  would  correspond  to  a  '  constant '  in  the 
formula.  But  even  then  we  could  not  climb  80  ft.  into  the  air  unless  there 
was  something  to  secure  the  joined  ladder.  We  could  not  descend  80  ft.  into 
the  earth  unless  there  was  an  opening,  nor  could  we  cross  an  80  ft.  gap.  For 
each  of  these  uses  we  need  something  which  is  a  •  constant '  for  the  special 
case.  It  is  in  this  way  that  all  mathematical  demonstrations  and  assertions 
need  to  be  examined.  They  mislead  people  by  their  very  definiteness  and 
apparent  exactness. . . ." — J.  T.  Spragde. 

III.  Integration  of  a  sum  and  of  a  difference.     Since 
d(x  +  y  +  z  +...)  =  dx  +  dy  +  dz  + ... , 


§  70.  THE  INTEGRAL  CALCULUS.  195 

it  follows  that 

j(dx  +  dy  +  dz  +...)  =*  jdx  +  jdy  +  jdz  + . . . , 
=  x  +  y  +  z  + . . . , 
plus   the   arbitrary  constant   of  integration.     It   is  customary  to 
append  the  integration  constant  to  the  final  result,  not  to  the  inter- 
mediate stages  of  the  integration.     Similarly, 

j(dx  -  dy  -  dz  - . . . )  =  jdx  -  jdy  -  jdz  - . . . , 

=  x-y-z-...+  C. 

In  words,  the  integral  of  the  sum  or  difference  of  any  number  of 

differentials  is  equal  to  the  sum  of  their  respective  integrals. 

Examples. — (1)  Remembering  that  log  xy  =  log  x  +  log  y,  show  that 

/{log  (a  +  bx)  (1  +  2x)}dx  =  j  log  (a  +  bx)dx  +  j  log  (1  +  2x)dx  +  C. 

(2)  Show  J  log^-^dx  =  Aog(a  +  bx)dx  -  (log  (1  +  2x)dx  +  G. 

IV.  Integration  of  xndx.    Since  the  differential  calculus,  page  37, 
teaches  us  that 

d(xn  + x)  =  (n  +  l)xndx ;  xn .  dx  =  d(- — =-)  ; 

we  infer  that 


-J 


?n+l 


x"dx  =  n~Tl+ °'  *  *  '  (1) 
Hence,  to  integrate  any  expression  of  the  form  axn .  dx,  it  is 
necessary  to  increase  the  index  of  the  variable  by  unity,  multiply 
by  any  constant  term  that  may  be  present,  and  divide  the  product 
by  the  new  index.  An  apparent  exception  occurs  when  n  =  -  1, 
for  then 


f 


x~1+1     1 


But  we  can  get  at  the  integration  by  remembering  that 

dQog  x)  =  —  =  x  ~ 1 .  dx  ;  .-.    —  =  log  x  +  G.    .         (2) 

If,  therefore,  the  numerator  of  a  fraction  can  be  obtained  by  the 
differentiation  of  its  denominator,  the  integral  is  the  natural  log- 
arithm of  the  denominator. 

I  want  the  beginner  to  notice  that  instead  of  writing  log  x  +  (7, 
we  may  put  log  x  +  log  a  =  log  ax,  for  log  a  is  an  arbitrary  con- 
stant as  well  as  C.     Hence  log  a  =  G. 

Examples. — (1)  One  of  the  commonest  equations  in  physical  chemistry  is, 
dx  =  k(a  -  x)dt.     Rearranging  terms,  we  obtain 


196  HIGHER  MATHEMATICS.  §  70. 

J         J  a  -  x  J    a  -  x 

Hence  kt  =  -  log  (a  -  x) ;  but  log  1  =  0,  ,\  kt  =  log  1  -  log(a  -  x) ;  or, 

(2)  Wilhelmy's  equation,  dy/dt  =  -  ay,  already  discussed  in  connection 
with  the  "  compound  interest  law,"  page  63,  may  be  written 

^=-adt;.:    (dJi  =  -at. 

y  J  y 

Remembering  that  log  e  =  1 ;  log  y  =  log  b  -  at  log  e  m  log  e  -  <**  +  log  b, 
where  log  b  is  the  integration  constant,  hence,  log  be  -  «<  =  log  y ;  and,  y  =  be~  "i. 
A  meaning  for  the  constants  will  be  deduced  in  the  next  section. 

(3)  Show  j±x  - 5  dx  =  4fa?  - 5 .  dx  =  -  x  - 4  +  G.  Here  n  of  (1)  =  -  5, 
and  n+l  =  -5+l=-4. 

(4)  Show  fax* .  dx  =  \a&  +  C. 

(5)  Show  flax  -  £  .  dx  =  5axr  +  C. 

(6)  Show  j2bx .  dxj{a  -  bx2)  =  -  log{a  -  6a;2)  +  C 

(7)  By  a  similar  method  to  that  employed  for  evaluating  jxndx ;  \x  -  ldx ; 
6how  that 

faxdx  =  ^-^  +  C  ;  [e*dx  =  e*  +  C',  (e-  «*dx  =  -  -e  -  "*  +  G.         (3) 

(8)  Prove  that  -  /^=_i-  .  _i_  +  O,         .        .        .        .        (4) 
11  J  xn    n  -  1    xn~x        *  \' 

by  differentiating  the  right-hand  side.     Keep  your  result  for  use  later  on. 

(9)  Evaluate  Jsin4a; .  cos  x  .  dx.  Note  that  cos  x  .  dx  =  d(sin  x),  and  that 
sin4a?  is  the  mathematician's  way l  of  writing  (sin  a;)4.     Ansr.  £  sin5ic  +  C. 

.*.  Jsin4aj .  cos  x  .  dx  =  j£  sin4a; .  d(sin  x)  =  \  sinBaj .  +  G. 

(10)  What  is  wrong  with  this  problem :  »•  Evaluate  the  integral  \x'A "  ? 
Hint.  The  symbol  "  J  "  has  no  meaning  apart  from  the  accompanying  "  dx". 
For  brevity,  we  call  "  JM  the  symbol  of  integration,  but  the  integral  must  be 
written  or  understood  to  mean  j . . .  dx. 

(11)  If  y  =  a+  M  +  cP  ;  show  that  jydt  =  at+  %bt2  +  $ct*+C.  (Heilborn, 
Zeit.  phys.  Chem.,  7,  367,  1891). 

V.  Integration  of  the  product  of  a  polynomial  and  its  differential. 
Since,  page  39, 

d(axm  +  b)n  =  n(axm  +  b)n  ~  lamxm  ~  xdx, 
where   amxm  ~  Ydx  has  been  obtained   by  differentiating  the   ex- 
pression within  the  brackets, 

.-.  n\{axm  +  b)n-1amxm~1dx  =  (axm  +  b)n  +  C.    .         (5) 
In  words,  integrate  the  product  of  a  polynomial  with  its  differ- 

1  But  we  must  not  write  sin  -  *x  for  (sin  x)  -  l,  nor  (sin  x)-1  for  sin  -  lx.  Sin  -  *f 
cos  -  K  tan  - 1, . . .  have  the  special  meaning  pointed  out  in  §  17. 


§  70.  THE  INTEGRAL^CALCULUS.  197 

ential,  increase  the  index  of  the  polynomial  by  unity  and  divide  the 
result  by  the  new  exponent. 

Examples.— (1)  Show  J(3aa^  +  l)29aa2 .  dx  =  ${Sax?  +  l)3  +  G. 
(2)  Show  j(x  +  1)  ~Ux  =  3{x  +  l)i  +  C. 

VI,  Integration  of  expressions  of  the  type  : 

(a  +  bx  +  ex2  + . . .  )mxdx,  ...  (6) 
where  m  is  a  positive  integer.  Multiply  out  and  integrate  each 
term  separately. 

Examples.— (1)  J(l  +  xfxHx  =  j(x3  +  2x4  +  x5)dx  =  (\  +  \x  +  ^x*  +  C. 
(2)  Show  that  ](a  +  x\)}x\dx  =  (fa2  +  ax\  +  %x)x%  +  G 

Here  are  a  few  simple  though  useful  "  tips  "  for  special  notice  : 
(i)  Any  constant  term  or  a  number  may  be  added  to  the  nume- 
rator of  a  fraction  provided  the  differential  sign  is  placed  before 
it.  The  object  of  this  is  usually  to  show  that  the  numerator  of 
the  given  integral  has  been  obtained  by  the  differentiation  of  the 
denominator.  If  successful  the  integral  reduces  to  the  logarithm 
of  the  denominator.     E.g., 

0f   xdx  (72(1  -  x2)         .      n  ''■■.  1  „ 

H.  Danneel  (Zet'J.  phys.  Chem.,  33,  415,  1900)  used  an  integral  like 
this  in  studying  the  "  free  energy  of  chemical  reactions  ". 

(ii)  Note  the  addition  of  log  1  makes  no  difference  to  the  value  of 
an  expression,  because  log  1  =  0  ;  similarly,  multiplication  by  logee 
makes  no  difference  to  the  value  of  any  term,  because  logee  =  1. 

(iii)  Jsin  nx  .  dx  may  be  made  to  depend  on  the  known  integral 
Jsin  nx  .  d(nx)  by  multiplying  and  dividing  by  n.     E.g., 

I  cos  nx  .  dx  =  - 1  cos  nx .  d(nx)  =  -sin  nx  +  C. 
J  n)  v     '       n 

(iv)  It  makes  no  difference  to  the  value  of  any  term  if  the  same 
quantity  be  added  and  then  subtracted  from  it ;  or  if  the  term  be 
multiplied  with  and  then  divided  by  the  same  quantity.     E.g., 

x.dx  _f(* +  »)-!,,_     f/1     1  1     \,_    x    lfdq  +  ae) 

2x  ' 


(x^    fa  +  x)  -  j    _  f/i  _  l        l    \d     x  _  l  Ki  + 
Jl  +  2*    J     1  +  2x    ax    )\2    2  •  1  +  2*7^-2     lJ~TT 


Examples. — (1)  Show  by  (16),  page  44,  and  (2)  above, 

f  dx  _f    d(logx)     _  fdjlogx  -1)      __  t%1 

J  (x*logx  -  x*)i-J  (logs  -  l)i~J(loga;-l)i-2(logaJ  "  1)i+0' 

(2)  The  following  equation  occurs  in  the  theory  of  electrons  (Encyc.  Brit., 


198  HIGHER  MATHEMATICS.  §  71 

26,  61,  1902) :   dxjdt  =  (ua/D)  sinpt ;  hence  show  x  =  -  (ualpD)  cos  pt  +  C 
where  u,  a,  p  and  JD  are  constants.     Use  (iv)  above. 

(3)  Show  that  jx(l  +  2x)  -  ldx  =  \x  -  \  log  (1  +  2x)  +  C.     Use  (iv)  and  (i). 

The  favourite  methods  for  integration  are  by  processes  known 
as  "  the  substitution  of  a  new  variable,"  "  integration  by  parts  "  and 
by  "  resolution  into  partial  fractions  ".  The  student  is  advised  to 
pay  particular  attention  to  these  operations.  Before  proceeding 
to  the  description  of  these  methods,  I  will  return  once  more  to  the 
integration  constant. 

§  71.    How  to  find  a  Value  for  the  Integration  Constant. 

It  is  perhaps  unnecessary  to  remind  the  reader  that  integration 
constants  must  not  be  confused  with  the  constants  belonging  to  the 
original  equation.  For  instance,  in  the  law  of  descent  of  a  falling 
body 

^=  g  ;  jW=  g^dt ;  or,  V  =  gt  +  C.  .         .         (1) 

Here  g  is  a  constant  representing  the  increase  of  velocity  due  to 
the  earth's  attraction,  C  is  the  constant  of  integration.  There  are 
two  methods  in  general  use  for  finding  the  value  of  the  integration 
constant. 

Fikst  Method. — Eeturning  to  the  falling  body,  and  to  its 
equation  of  motion, 

V  =  gt  +  G. 
On  attempting  to  apply  this  equation  to  an  actual  experiment,  we 
should  find  that,  at  the  moment  we  began  to  calculate  the  velocity, 
the  body  might  be  moving  upwards  or  downwards,  or  starting 
from  a  position  of  rest.  All  these  possibilities  are  included  in  the 
integration  constant  G.  Let  VQ  denote  the  initial  velocity  of  the 
body.     The  computation  begins  when  t  =  0,  hence 

V0  =  gx  0  +  G;  or,  C=  70. 

If  the  body  starts  to  fall  from  a  position  of  rest,  V0  =  C  =  0,  and 

jdv  =  gt ;  or,  V  =  gt. 

This  suggests  a  method  for  evaluating  the  constant  whenever 
the  nature  of  the  problem  permits  us  to  deduce  the  value  of  the 
function  for  particular  values  of  the  variable.  If  possible,  there- 
fore, substitute  particular  values  of  the  variables  in  the  equation 
containing  the  integration  constant  and  solve  the  resulting  ex- 
pression for  G. 


§71.  THE  INTEGRAL  CALCULUS.  199 

Examples. — (1)  Find  the  value  of  C  in  the  equation 

'  =  Sl0^+C>  ....         (2) 

which  is  the  integral  of  a  standard  "  velocity  equation  "  of  physical  chemistry. 
t  represents  the  time  required  for  the  formation  of  an  amount  of  substance  x. 
When  the  reaction  is  just  beginning,  x  =  0  and  t  =  0.  Substitute  these 
values  of  x  and  t  in  (2). 

^Iogi+C  =  0;or,  0  =  -  ^  log -. 

Substitute  this  value  of  C  in  the  given  equation  and  we  get 
1/ 


k^^-x-^aj^k^T^x' 
10232  -  0-1685T  -  O'OOIOIT2         ,     .       fl  ftje       . 

(2)  if  -^-fir    = 2T2 '  and  *  =  6'25,  when 

T  =  3100,  show  that  the  integration  constant  is  -  2-0603.  Hint,  log  k  = 
5116/T  +  0-08425  log  T  +  0-000505T  +  C.  Use  natural  logs.  Substitute  the 
above  values  of  k  and  T  in  this  equation.  We  get  1*8326  =  1-65  +  0-6774  + 
1-5655  +  C ;  etc. 

(3)  If  the  temperature  of  a  substance  be  raised  dT(°  abs.)  it  is  commonly 
said  that  it  has  gained  the  entropy  d<p  =  dT/T.  Show  that  the  entropy,  <j>, 
of  one  gram  of  water  at  T°  is  log  T  -  log  273  if  the  entropy  at  0°  G.  be  taken 
as  zero.     Hint.  When  <f>  =  0,  T  =  273,  etc. 

(4)  In  Soret's  experiments  "  On  the  Density  of  Ozone "  {Ann.  Chim. 
Phys.  [4],  7,  113,  1866 ;  13,  257,  1868)  a  vessel  A  containing  v0  volumes  of 
ozone  mixed  with  oxygen  was  placed  in  communication  with  another  vessel 
B  containing  oxygen,  but  no  ozone.  The  volume,  dv,  of  ozone  which  diffused 
from  A  to  B  during  the  given  interval  of  time,  dt,  is  proportional  to  the 
difference  in  the  quantity  of  ozone  present  in  the  two  vessels,  and  to  the 
duration  of  the  interval  dt.  If  v  volumes  of  ozone  have  passed  from  A  to  B 
at  the  time  t,  the  vessel  A,  at  the  time  t,  will  have  v0  -  v  volumes  of  ozone 
in  it,  and  the  vessel  B  will  have  v  volumes.  The  difference  in  the  amount  of 
ozone  in  the  two  vessels  is  therefore  v0  -  2v.  By  Graham's  law,  the  rate  of 
diffusion  of  ozone  from  A  to  B  is  inversely  proportional  to  the  square  root  of 
the  density,  p,  of  the  ozone.    Hence,  by  the  rules  of  variation,  page  22, 

dv  =  -^={v0  -  2v)dt ;  or,  -£  =  -^{v0  -  2v), 

where  a  is  a  constant  whose  numerical  value  depends  upon  the  nature  of  the 
vessels  used  in  the  experiment,  etc.     Now,  remembering  that  v0  is  a  constant, 
[     dv       '   1  f  d(2v)     m      1  fd(v0  -  2v)  _  _  log  K  -  2v) 
Jv0  -  2v      2}  v0  -  2v  2)    v0  -  2v  2 

But  when  t  =  0,  v  =  0,  .-.  C  =  £  log  vQ.     Consequently, 

v0  2a,  v       1/  _^\ 

10^i^-^  =  7/;o^  =  2(i-e  *) 

For  the  same  gas,  the  same  apparatus,  and  the  same  interval  of  time,  p,  t,  and 
o  will  all  be  constant,  and  therefore, 

—  =  Constant. 
vn 


200  HIGHER  MATHEMATICS.  §  72. 

With  different  gases,  under  the  same  conditions,  any  difference  in  the  value 
of  vnjv0  must  be  due  to  the  different  densities  of  the  gases.  The  mean  of  a 
series  of  experiments  with  chlorine  (density,  35-5),  carbon  dioxide  (density, 
22),  and  ozone  (density,  a;),  gave  the  following  for  the  value  of  this  ratio : — 

C02,  0-29;  Ozone,  0-271 ;  Cl2,  0-227. 
Compare  chlorine  with  ozone.     Let  x  denote  the  density  of  ozone.     Then,  by 
Graham's  law, 

7? (°s) :  ?(01a)  =  ^/S^E  :  six', .-.  (0-271)2  :  (0-227)2  =  35-5  :  »;  .-.  x  =  24-9, 
which  agrees  with  the  triatomic  molecule,  03. 

Second  Method. — Another  way  is  to  find  the  values  of  x 
corresponding  to  two  different  values  of  t.  Substitute  the  two 
sets  of  results  in  the  given  equation.  The  constant  can  then  be 
made  to  disappear  by  subtraction.  The  result  of  this  method  is  to 
eliminate,  not  evaluate  the  constant. 

Examples. — (1)  In  the  above  equation,  (2),  assume  that  when  t  =  tv 
x  =  xv  and  when  t  =  t2,  x  =  x2 ;  where  xlt  x2l  ^  and  t2  are  numerical  measure- 
ments.    Substitute  these  results  in  (2). 

1.  1  n     .        1  .  1  „ 

By  subtraction  and  rearrangement  of  terms 

7  1       ,      a  -  x, 

Jc  -  _.  log z*. 

t2  -  tx     °  a  -  x2 

(2)  If  the  specific  heat,  a;  of  a  substance  at  0°  is  given  by  the  expression 
<r  =  a  +  bd,  and  the  quantity  of  heat,  dQ,  required  to  raise  the  temperature 
of  unit  mass  of  the  substance  is  dQ  =  add,  show  that  the  amount  of  heat 
required  to  heat  the  substance  from  0°  to  02°  is 

Q  =]* {a  +  be)de  =  a(02  -  ej  +  bb{e2*  -  e*). 

Numerous  examples  of  both  methods  will  occur  in  the  course 
of  this  work.  Some  have  already  been  given  in  the  discussion  on 
the  "  Compound  Interest  Law  in  Nature,"  page  56. 

§  72.  Integration  by  the  Substitution  of  a  New  Variable. 

When  a  function  can  neither  be  integrated  by  reference  to 
Table  II.,  nor  by  the  methods  of  §  71,  a  suitable  change  of  variable 
may  cause  the  function  to  assume  a  less  refractory  form.  The  new 
variable  is,  of  course,  a  known  function  of  the  old.  This  method 
of  integration  is,  perhaps,  best  explained  by  the  study  of  a  few 
typical  examples. 

I.  Evaluate  j(a  +  x)ndx.     Put  a  +  x  =  y;  therefore,  dx  =  dy. 


§  72.  THE  INTEGRAL  CALCULUS.  201 

Substitute  y  and  dy  in  place  of  their  equivalent  values  a  —  x  and 
dx  in  the  given  integral.  We  thus  obtain  an  integral  with  a  new 
variable  y,  in  place  of  x,  namely,  \{a  +  x)ndx  =  jyndy.  From  (1), 
page  195,  jyndy  =  yn  +  1/(n  +  1)  +  G.  Restore  the  original  values 
of  y  and  dy,  and  we  get 

\(*  +  x")dx  =  t±fp  +  C.  .        .        (1) 

When  the  student  has  become  familiar  with  integration  he  will 
find  no  particular  difficulty  in  doing  these  examples  mentally. 

Examples. — (1)  Integrate  j(a  -  bx)ndx.     Ansr.  -  (a  -  bx)n  +  1/b(n  + 1)  +  C. 

(2)  Integrate  J  (a2  +  s2)  ~  lPxdx.     Ansr.  V(«2  +  »2)  +  G. 

(3)  Show 

f     dx  1  1  f    dx  1  1 

J  (a  +  x)»~     n-1'  (a  +  x)n-1  +  C;  J  {a-x)»~n-l '  (a-aj)"-1+C* 

(4)  Show  that  /^^  =«  =  log(logs)  +  0. 

II.  Integrate  (1  -  ax)mxndx,  where  m  or  n  is  a  positive  integer, 
and  the  x  within  the  brackets  has  unit  index.  Put  y  =  1  -  ax, 
therefore,  x  =  (1  -  y)/a ;  and  dx  =  -  dy/a.  Substitute  these 
values  of  x  and  dx  in  the  original  equation,  and  we  get 

J(l  -  ax)mxndx  -  —^  j(l  -  y)n{-  ym)dy, 

which  has  the  same  form  as  (6),  page  197.  The  rest  of  the  work 
is  obvious — expand  (1  -  y)n  by  the  binomial  theorem,  page  36; 
multiply  through  with  -  ymdy ;  and  integrate  as  indicated  in 
III.,  page  194. 

Example. — Show  jx(a  +  x)$dx  =  ^{ix  -  3a) (a  +  %)$  +  G.  Hint.  Put 
a  +  x  =  y,  etc. 

III. — Trigonometrical  functions  can  often  be  integrated  by  these 
methods.     For  example,  required  the  value  of  Jtan  xdx. 

f  fsin  x 

I  tan  xdx  =   I dx.      ...         (2) 

J  J  cos  X 

Let  cos  x  =  u,  -  sin  xdx  =  du.  Since  -  jdu/u  —  -  log  u,  and 
log  1  =  0,  therefore, 

Jtan  xdx  =  log 1  =  log  sec  x  +  G. 
°cos  x        ° 

Or,  remembering  that  -  d(cos  x)  =  sin  x .  dx,  we  can  go  straight 


202  HIGHER  MATHEMATICS.  §  72. 

on  without  any  substitution  at  all, 

fsin  x  fd(cos  x) 

— —ax  =  -    -r -  =  -  log  cos  x,  etc. 

J  cos  x  J    cos  x  & 

Examples. — (1)  Show  that  Jsin  x  .  cos  x  .  dx  =  £sin2<c  +  C.    Put  sin  x=u. 

(2)  Show  Jcot  xdx  =  log  sin  x  +  C.     Hint,  cot  a;  =  cos  as/sin  x,  etc. 

(3)  Show  Jsin  x  .  dx/cos2x  =  sec  a?  +  C.     Hint.  Put  cos  x  =  w,  or  go  the 
hort  cut  as  in  (2)  above. 

(4)  Show  that  Jcos  x  .  dxjsin2x  =  -  cosec  x  +  C. 

(5)  Show  that  je - *2xdx  =  -\\e~*2.  2xdx  =  +  \\e - *2d{x2)  =  - \e ~ *2  +  C. 

Some  expressions  require  a  little  "  humouring  ".  Facility  in 
this  art  can  only  be  acquired  by  practice.  A  glance  over  the 
collection  of  formulae  on  pages  611  to  612  will  often  give  a  clue. 
In  this  way,  we  find  that  sin  x  =  2  sin  \x  .  cos  \x.     Hence  integrate 

dx  f  sec  \x  .  dx 

1 2  sin  \x .  cos  \x  or  J   2  sin  \x 
Divide  the  numerator  and    denominator  by  cos-|#,   then,    since 
l/cos2-|#  =  sec2^#  ;  and  d(tan  x)  =  sec2# .  dx,  page  49,  (3), 

Jdx       fsec2  J  a; .  d(\x)     Cd(ta,n±x)     ;  x     „ 

-• — =  — f — dr~  -1-1 — r— =  logtano  +  C. 
smx    J       tanjic  J   tannic  &        2 

The  substitutions  may  be  dim  cult  to  one  not  familiar  with  trig- 
onometry. 

Examples. — (1)  Remembering  that  cos  x  =  sin  (Jtt  +  x),  (9),  page  611, 
show  that  fdxjcos  x  =  log  tan  {^v  +  %x)  +  C.  Hint.  Proceed  as  in  the  illus- 
tration just  worked  out  in  the  text. 

(2)  Integrate/" Hint,  see  (19),page611;  cosxdx  =  d{sinx);  etc. 

w  J  sin  x .  cos  x 

fcos2x  +  sin2cc ,        /"cos  x ,        /"sin  x ,  _ 

.*.  /  — : aa>=  /  -! — aa;+  / dx  =  log  tan  a;  +  C. 

'  'J    sin  x  cos  x  J  sin  as     ^  J  cos  a;  8  T 

(3)  Integrate  J(a2  -  cc2)  "  %dx.      Put  y  =  je/a;  .:x  =  ay,    .•.  dx  =  ady; 


f  ^       ;     r 
Jsina  '  **  JS 


V(a2  -K2)=flvT 


/; 


da;        =  f     dy 


=  sm~ly  =  sin-1-+  C. 


a:2     J  Jl  -  y2  ~  a 


//a?  -  x2  (a2  -  x2)?  1 

&— dx=  ~  ~12&-  +  c'    Hint  Put  x  =  y- 

IV.  Expressions  involving  the  square  root  of  a  quadratic  binomial 
can  very  often  be  readily  solved  by  the  aid  of  a  lucky  trigono- 
metrical substitution.  The  form  of  the  inverse  trigonometrical 
functions  (Table  II.)  will  sometimes  serve  as  a  guide  in  the  right 
choice.     If  the  binomial  has  the  forms  : 


§  72.  THE  INTEGRAL  CALCULUS.  203 

Jx2  -f-  1 ;  or,   six1  +  a2 ;  try  x  =  tan  0  ;  or,  a  tan  0  ;  or,  cot  6.  (3) 

v/l  -  x2 ;  or,   a/a'2  -  x2 ;  try  a;  =  sin  0  ;  or,  a  sin  0  ;  or,  cos  6.  (4) 

Va;2  -  1 ;  or,   Jx2 -  a'2;  try  a?  =  sec  0 ;  or,  a  sec  0 ;  or,  cosec  0,  (5) 

•  »/aa-  (a  +  6)'2 ;  try  re  +  b  =  a  sin  0.     .         .  (6) 

Examples. — (1)  Find  the  value  of  j  *J{a?  -  x2)dx.  In  accordance  with  the 
above  rule  (4),  put  x  =  a  sin  0,  .*.  dx  =  a  cos  0  .  d0.  Consequently,  by  substi- 
tution, 

JV(a2  -  x2)dcc  =  a2Jcos20d0  -  |a2J(l  +  cos  29)dd  =  £a2(0  +  £  sin  2^)  I 

since  2cos20  =  1  +  cos 20,  (31),  page  612.     But  x  =  a  sin  0,  0  =  sin-  'a/a,  and 

.-.  £  sin  20  =  sin  0  .  cos  0  =  sin  0  J (1  -  sin20)  =  si  a2  -  x2 .  a;/**2. 

.-.  JV(a2  -  a;2)dic  =  £a2sin  ~^x\a  +  \x  sld2  -  x2  +  C. 

(2)  The  integration  of  fee2  si  a2  -  x2 .  dx  arises  in  the  study  of  molecular  dyn- 
amics (Helmholtz'sForteswwgrew  Uber  theoretische  Physik,  2, 176, 1902).  Rule  (4). 
Put  x  =  a  sin  0;  .-.  x2  =  a%in20  ;  dx  =  a  cos  0 .  dd ;  ^/(a2  -  a?2)  =  a^l  -  sin20). 
Remembering  that  cos  0  =  ^(1  -  sin20),  and  sin  2x  =  2  sin  a? .  cos  as,  (19)  and 
(29),  pages  611  and  612,  we  get  the  expression  ' 

jx2  sla^^x2  .  dx  =  a4Jsin20  .  cos20  .  dd  =  |a4Jsin220  .  d(20) ; 
which  will  be  integrated  very  shortly. 

(3)  Show  f     _J*  2  =  y\^rj  +  G'     Put«=cos0.    Rule  (4).     If 

the  beginner  has  forgotten  his  "  trig."  he  had  better  verify  these  steps  from 
the  collection  of  trigonometrical  formulae  in  Appendix  I.,  page  611.  The 
substitutions  are  here  very  ingenious,  but  difficult  to  work  out  de  novo. 

f  sin  Ode  f       dd  1  f    dO        _[         20J6\. 

=  'J  (1  -  cmQJjZT^fi  ~  "Jl  -  cos0  "      2jsin2^0         J  cosec2  A^ ' 

2         sin  £0       \  2  sin2£0       \  1  -  cos  0       \  1  -  x 


,,  2      1)  =  log  (a;  +  six2  +  1)  +  C.     Put  ic=tan  0.     Rule  (3). 
?  +  sec0;    .•.  =  x  +  *y(2:2  +  1) ;  see  Ex.  (1),  pre- 

log  (x  +  s/x2  -  1)  +  C.    Put  a?  -  sec  0.    Rule  (5). 


Given  tan(|?r  +  $0)  =  tan  0  +  sec0;    ,\  -  x  +  x/(a;2  +  1) ;  see  Ex.  (1),  pre- 
ceding set. 

dx 


(5)  Show/"- 


x/(*2  -  1) 

It  may  be  here  remarked  that  whenever  there  is  a  function  of 
the  second  degree  included  under  a  root  sign,  such,  for  instance,  as 
six2  +  px  +  q,  the  substitution  of 

z  =  x  +  six2  +  px  +  q,  ...  (7) 
will  enable  the  integration  to  be  performed.  For  the  sake  of  ease, 
let  us  take  the  integral  discussed  in  Ex.  (4),  above,  for  illustrative 
purposes.     Obviously,  on  reference  to  (7),  p  =  0,  q  =  1.     Hence, 


204  HIGHER  MATHEMATICS.  §  72. 

put  Z  =  X  +  sjx2  +  1  ; 

.'.  z2  -  2zx  +  x2  =  x2  +  1  ;  x  =  J(^2  -  1)^  - 1 ; 
Va;2  +  1  =  *  -W-  i(^  +  1)^-1;  <fo  =  i(s2  +  l>-2^. 

f      d#  Cdz      _  ,  . 

'""  J  ^a;2  +  1  =  J7  =  loS*  =  lo%(x  +  «&'+  x)  +  C- 

F.  2%e  integration  of  expressions  containing  fractional  powers 
of  x  and  of  xm(a  +  bxn)p.dx.  Here  m,  n,  or  p  may  be  fractional. 
In  this  case  the  expression  can  be  made  rational  by  substituting 
x  =  zr,  or  a  +  bx  =  zr,  where  r  is  the  least  common  multiple — 
L.C.M. — of  the  denominators  of  the  several  fractions. 

Examples.— (1)  Evaluate  fxP(l  +  x2)idx.  Here  the  L.C.M.  of  the 
denominators  of  the  fractional  parts  is  2.  Put  1  +  x2  =  z2 ;  then,  x2  =  z2  - 1 ; 
.•.  z=  vl  +  x2;  x.dx  —  z.dz.  Substitute  these  values  as  required  in  the 
original  expression,  and 

Ja^l  +  xrfdx  =  j{z2  -  l)2z2dz  =  \{z*  -  2z4  +  z2)dz  =  f#'  -  f^  +  \z* ; 
.-.  Jz5(l  +  xrfdx  =  ^(1  +  z2)*{15(l  +  a;2)2  +  42(1  +  x2)  +  35}  +  G. 

(2)  Evaluate  J»  -  4(1  +  x2)  -  *da;.  Here  again  r  =  2.  Put  1  +  x2  =  z2x2 ; 
.-.o;-2  =  *2-l;  .-.  aj-4  =  (02-l)2;  .-.  x  =  (*2-l)  ri  ;  &c=  -  (s2-l)  "te; 
(1  +  a?2)  ~  *  =  1/sa;  =  N/(22  -  l)jz.  Consequently,  we  get  the  expression 
Jaj~4(l  +  x2)  ~$dx  =  -  j{z2  -  l)dz  =  -  $z*  +  z ;  and  hence, 

dx =  (2s2  -!)(!+  x2)$ 

x*(l  +  x2)$  Sx*  + 

(3)  Evaluate  J(l  +  x%)  -  lx%dx.  Here,  the  L.C.M.  is  6.  Hence,  put 
x  =  z6.  The  final  result  is  fa;*  -  £ x%  +  $x%  -  f x%  +  6  tan  -  ]a;£  +  C.  Hint. 
To  integrate  (i  +  z2)  —  1z8dz;  first  divide  z8  by  1  +  z2,  and  multiply  through 
With  dz. 

(4)  Show  ft**-*  dx = J|a;T^ - l?ajH  +  O.  The  least  common  multiple 
is  12.     Hence,  put  a;  =  z12,  etc. 

I  have  no  doubt  that  the  reader  is  now  in  a  position  to  under- 
stand why  the  study  of  differentiation  must  precede  integration. 
"  Common  integration,"  said  A.  de  Morgan,  "  is  only  the  memory  of 
differentiation,  the  different  artifices  by  which  integration  is  effected 
are  changes,  not  from  the  known  to  the  unknown,  but  from  forms 
in  which  memory  will  not  serve  us  to  those  in  which  it  will" 
{Trans.  Cambridge  Phil.  Soc,  8,  188,  1844).  The  purpose  of  the 
substitution  of  a  new  variable  is  to  transform  the  given  integral 
into  another  integral  which  has  been  obtained  by  the  differentia- 


/; 


§  73.  THE  INTEGRAL  CALCULUS.  205 

tion  of  a  known  function.  The  integration  of  any  function 
therefore  ultimately  resolves  itself  into  the  direct  or  indirect 
comparison  of  the  given  integral  with  a  tabulated  list  of  the  results 
of  the  differentiation  of  known  function!  The  reader  will  find  it 
an  advantage  to  keep  such  a  list  of  known  integrals  at  hand. 
A  set  of  standard  types  is  given  in  Table  II. ,  page  193,  but  this 
list  should  be  extended  by  the  student  himself ;  or  A  Short  Table 
of  Integrals  by  B.  0.  Pierce,  Boston,  1898,  can  be  purchased. 

When  an  expression  cannot  be  rationalised  or  transformed  into 
a  known  integral  by  the  foregoing  methods,  we  proceed  to  the 
so-called  "  methods  of  reduction"  which  will  be  discussed  in  the 
three  succeeding  sections.  These  may  also  furnish  alternative 
methods  for  transforming  some  of  the  integrals  which  have  just 
been  discussed. 

§  73.   Integration  by  Parts. 

The  differentiation  of  the  product  uv,  furnishes 
d(uv)  =  vdu  +  udv. 
By  integrating  both  sides  of  this  expression  we  obtain 

uv  =  jvdu  +  judv. 
Hence,  by  a  transposition  of  terms,  we  have 

judv  =  uv  -  jvdu  +  G.  .         .        (A) 

that  is  to  say,  the  integral  of  udv  can  be  obtained  provided  vdu  can 
be  integrated.  This  procedure  is  called  integration  by  parts. 
The  geometrical  interpretation  will  be  apparent  after  A  has  been 
deduced  from  Fig.  7,  page  41.  Since  equation  A  is  used  for  re- 
ducing involved  integrals  to  simpler  forms,  it  may  be  called  a 
reduction  formulae.  More  complex  reduction  formulae  will 
come  later. 

Examples. — (1)  Evaluate  jx  log  xdx.    Put 

u  —  log  x,  I  dv  =  x .  dx ; 
du  =  dx/x,  I    v  =  $x2. 
Substitute  in  A,  and  we  obtain 

ju.dv  =  jx  log x .  dx  =  uv  -  jv .  du, 

=  %x2 log x  -  j\x.dx  =  \x> log x  -  |aj2, 
=  ^2(logaj-i)  +  a 
(2)  Show  that  jx  cos  x  .   dx  =  x  sin  x  +  cos  x  +  C.    Put     . 
u  =  x,    I  dv  =  cos  x.dx; 
du  —  dx,  I    v  =  sin  x. 
From  A,  jx  cos  x .  dx  =  x  sin  x  -  Jsin  x.dx;  etc. 


206  HIGHER  MATHEMATICS.  §  74. 

(3)  Evaluate  J  „J(a2  -  x2)dx,  by  "  integration  by  parts  ".     Put 

u  =  ,J{a2  -  x2),  \dv  =  dx; 

du=  -  x.dxfjia2  -  x2),  \    v  =  x. 

r  , r     x2dx 

.-.  J  V(«2  -  a;2)^  =  x  si  a?  -  x2  +  J  ^(a2  _  ^ 

/-a s-      r(«2  -  (a2  -  a;2)Wa; 

/a2dx  C 

J{a2-x2)-\>J(a2-^dx' 

Transpose  the  last  term  to  the  left-hand  side : 

2j\Ja2  -  x2  dx  =  xs/a2  -  x2  +  a2sin-1x/a  (page  193), 

.-.  |V(a2  -  x2)dx  =  $0?  sin -*xla  +  lxsj{a2  -  x2)  +  C. 

(4)  Show  that  jxexdx  =  (x  -  l)ex  +  C.    Take  u  =  x ;  <2v  =  e*doj. 

(5)  Show  JzVcte  =  (x2  -  2x  +  2)ex  +  C.  Take  dv  =  exda:  and  use  the 
result  of  the  preceding  example  for  vdu. 

(6)  Show,  integrating  by  parts,  that  J  log  x .  dx  =  a?(log  x  -  1)  +  0. 

(7)  Show  that  the  result  of  integrating  jx  ~  ldx  by  parts  is  \x  ~  xdx  itself. 

The  selection  of  the  proper  values  of  u  and  v  is  to  be  determined 
by  trial.  A  little  practice  will  enable  one  to  make  the  right  selec- 
tion instinctively.  The  rule  is  that  the  integral  jv.du  must  be 
more  easily  integrated  than  the  given  expression.  In  dealing  with 
Ex.  (4),  for  instance,  if  we  had  taken  u  =  <f,  dv  =  xdx  ,  jv  .du 
would  have  assumed  the  form  ^jx2exdx,  which  is  a  more  complex 
integral  than  the  one  to  be  reduced. 

§  75.    Successive  Integration  by  Parts. 

A  complex  integral  can  often  be  reduced  to  one  of  the  standard 
forms  by  the  "  method  of  integration  by  parts  ".  By  a  repeated 
application  of  this  method,  complicated  expressions  may  often  be 
integrated,  or  else,  if  the  expression  cannot  be  integrated,  the 
non-integrable  part  may  be  reduced  to  its  simplest  form.  This 
procedure  is  sometimes  called  integration  by  successive  reduc- 
tion.    See  Ex.  (5),  above. 

Examples. — (1)  Evaluate  Jx2cos  nxdx.    Put 

u  =  a?2,         I  dv  =  {cos  nx .  d(nx)}fn ; 
du  =  2x  .'dx,  |    v  =  (sin  nx)Jn. 
Hence,  on  integration  by  parts, 


jx2 


nxdx  = 

a;2sin  nx 

2 
n 

/  x  sin  nx  . 

dx. 

n 

u 

-*.  1 

\dv 

=  sin  nx  .  dx; 

du 

=  dx,  \ 

V 

=  - 

(cos  nx)Jn 

Now  put 


S  74.  THE  INTEGRAL  CALCULUS.  207 

Hence, 

f     .  7  ajcosna;       f-  cos nx.dx  xooanx      sinnaj 

/  x  sm  nx .  dx  = /  = H 5 — -.       (2) 

J  n  J  n  n  nz    ■       v  ' 

Now  substitute  (2)  in  (1)  and  we  get, 


h 


a>2sin  nx      2x  cos  nx      2  sin  nx       _ 
x2cosnx.dx= 1 s • ; h  C. 

VI.  VI*  Vt.o 


(2)  In  t'he  last  example,  we  made  the  integral  Jsc2cos  nx .  dx  depend  on 
that  of  x  sin  nx .  dx,  and  this,  in  turn,  on  that  of  -  cos  nx .  d{nx),  thus 
reducing  the  given  integral  to  a  known  standard  form.  The  integral 
(a;4cosa;  dx  is  a  little  more  complex.     Put 

u  =  a;4,        I  dv  =  cos  xdx ; 
du  =  4x3dx,  I    v  =  sin  x. 
.'.  Jaj4cos  x  .  dx  =  a;4sin  x  -  ^a^sin  x .  dx. 
In  a  similar  way, 

4ja;3sin x.dx  =  4<c3cos x  -  3  .  4f<c2cos x  .  dx. 
Similarly, 

3 .  4j<c2cos  x  .  dx  =  3 . 4 .  ar^sin  x  +  2  3 .  4ja;  sin  x .  dx% 
and  finally, 

2.3. 4  fa;  sin  xdx  =  2 . 3 .  4a;  cos  x  +  1 . 2 . 3 . 4  sin  x. 
All  these  values  must  be  collected  together,  as  in  the  first  example.  In 
this  way,  the  integral  is  reduced,  by  successive  steps,  to  one  of  simpler  form. 
The  integral  Ja;4cos  x  .  dx  was  made  to  depend  on  that  of  ar'sin  x  .  dx,  this,  in 
turn,  on  that  of  a;2cos  x  .  dx,  and  so  on  until  we  finally  got  Jcos  x  .  dx,  a  well- 
known  standard  form. 

(3)  It  is  an  advantage  to  have  two  separate  sheets  of  paper  in  working 
through  these  examples  ;  on  one  work  as  in  the  preceding  examples  and  on 
the  other  enter  the  results  as  in  the  next  example.     Show  that 
\x*e?dx  =  x^e*  -  Sjx^dx ; 

=  xV  -  3(a;V  -  2jxexdx) ; 

=  x*ex  -  Sx2ex  +  2 . 3(a^  -  jexdx ; 

=  (a;3  -  3a;2  +  6x  -  6)ex  +  G. 

It  is  also  interesting  to  notice  that  we  sometimes  obtain  different 
results  with  different  substitutions.     For  instance,  we  get  either 

according  as  we  take  u  =  x  *,  etc.,  or,  u  =  ex.  In  the  last  series, 
(4),  the  numerator  and  denominator  of  each  term  has  been  multi- 
plied by  x*.  Differences  of  this  kind  can  generally  be  traced  to 
differences  in  the  range  of  the  variable,  or  to  differences  in  the 
value  of  the  integration  constants.  Another  example  occurs  during 
the  integration  of  (1  -  x)  ~  2dx.     In  one  case, 


208  HIGHEK  MATHEMATICS.  §  75. 

J(l  -  x)-Ux  =  -  J(l  -  aj)-2d(l  -  a?)  =  (1  -  x)~l  +  Ot;  (5) 
but  if  we  substitute  x  =  s  ~ J  before  integration,  then 

J(l-^)-2^=- J(^-l)-2^  =  (^_l)-i  =  o;(l-a;)-1  +  02.  (6) 
The  two  solutions  only  differ  by  the  constant  "  —  1 ".  By  adding 
-  1  to  (5),   (6)  is   obtained.      Gx  is   not  therefore  equal  to  C2, 

§  75.    Reduction  Formulae  (/or  reference). 

We  found  it  convenient,  on  page  205,  to  refer  certain  integrals 
to  a  ''standard  formula"  A;  and  on  page  206  reduced  complex 
integrals  to  known  integrals  by  a  repeated  application  of  the  same 
formula,  namely,  Ex.  (1),  page  206,  etc.  Such  a  formula  is  called  a 
reduction  formula.  The  following  are  standard  reduction  formulas 
convenient  for  the  integration  of  binomial  integrals  of  the  type  : — 

\xm(a  +  bxn)pdx (1) 

These  expressions  can  always  be  integrated  if  (m  +  l)/n  be  a 
positive  integer.  Four  cases  present  themselves  according  as  m, 
or  p  are  positive  or  negative. 

I. — m  is  positive. 

The  integral  fxm(a  +  bxn)p .  dx,  may  be  made  to  depend  on  that 
of  jxm  ~  n(a  +  bxn)p  +  1.dx,  through  the  following  reduction  formula. 
The  integral  (1)  is  equal  to 

xm-n  +  1(a  +  bxn)p  +  1  _  a{m 
b(m  +  np  +  1)  b(m 

when  m  is  a  positive  integer.  This  formula  may  be  applied  suc- 
cessively until  the  factor  outside  the  brackets,  under  the  integral 
sign,  is  less  than  n.  Then  proceed  as  on  page  204.  B  can  always  be 
integrated  if  {m  -  n  +  V)jn  is  a  positive  integer.    See  Ex.  (7),  below. 

II. — m  is  negative. 

In  B,  m  must  be  the  positive,  otherwise  the  index  will  increase, 
instead  of  diminish,  by  a  repeated  application  of  the  formula. 
When  m  is  negative,  it  can  be  shown  that  the  integral  (1)  is  equal 
to 

xm  +  l(a  +  bxn)p  +  l      b(np  +  m  +  n  +  1)  f      ..         ,   n^nl      ,,_ 

S rv ( — r^T\ "  la  +  (a  +  bxnYdx,  (C) 

aim  +  1)  aim  +1)  )  K  i      •>  \y\ 


+  np\1]i)\xm~^a+bxnydx'  <B) 


§75.  THE  INTEGRAL  CALCULUS.  209 

where  m  is  negative.  This  formula  diminishes  m  by  the  number 
of  units  in  w.  If  np  +  m  +  n  +  1  =  0,  the  part  to  be  integrated 
will  disappear,  and  the  integration  will  be  complete.  C  can  always 
be  integrated  if  (m  +  n  +  l)/n  is  a  positive  integer.  See  Ex.  (7), 
below. 

III. — p  is  positive. 

Another  useful  formula  diminishes  the  exponent  of  the  bracketed 
term,  so  that  the  integral  (1)  is  equal  to 

*~+>+l*r  +     ™p     Ua  +  bx~y  -  >dx>  .   (D) 

m  +  np  +  1        m  +  np  +  1 J     v  '  '  v   ' 

where  p  is  positive.  By  a  repeated  application  of  this  formula  the 
exponent  of  the  binomial,  if  positive,  may  be  reduced  to  a  positive 
or  negative  fraction  less  than  unity. 

IV. — p  is  negative. 
If  p  is  negative,  the  integral  (1)  is  equal  to 
xm  +  Ha  +  bxn)p  + 1      (np  +  m  +  n  +  1)  f     ,         7     «  A  , -         ' 
"         an(p  +  1)         +         an(P  +  l)        j*>  +  ****■  <*> 

Formulae  B,  C,  D,  E  have  been  deduced  from  (1),  page  208,  by 
the  method  of  integration  by  parts.  Perhaps  the  student  can  do 
this  for  himself.  The  reader  will  notice  that  formula  B  decreases 
(algebraically)  the  exponent  of  the  monomial  factor  from  m  to 
m  -  n  +  1,  while  C  increases  the  exponent  of  the  same  factor  from 
m  to  m  +  1.  Formula  D  decreases  the  exponent  of  the  binomial 
factor  from  p  top  -  1,  while  E  increases  the  exponent  of  the  binomial 
factor  from  p  to  p  +  1.  B  and  D  fail  when  np  +  m  +  1  =  0  ;  C 
fails  when  m  +  1  =  0  ;  E  fails  when^p  +  1  =  0.  When  B,  C,  and  D 
fail  use  7,  page  204  ;  if  E  fails  p  =  -  1  and  the  preceding  methods 
apply. 

Examples. — Evaluate  the  following  integrals : — 

(1)  J"  sj{a  +  x2)dx.  Hints.  Use  D.  Put  m  =  0,  b  =z  1,  n  =  2,  p  =  £.  Ansr. 
$[xj(a  +  x2)  +  a  log  {x  +  J  {a  +  x2)}]  +  C.     See  bottom  of  page  206. 

(2)  \xidxlsl(di  -  x2).  Hints.  Put  m  =  4,  b  =  -  1,  n  =  2,  p  =  -  £.  Use 
B  twice.     Ansr.  \{  ta'sin  -  yx\a  -  x(2x2  +  3a2)  J  (a2  -  x2)}  +  C. 

(3)  jjil  -  x')x*dx.     Hint.     Use  B.     Ansr.  -  \{x2  +  2)  ^/(l  -  x2)  +  C. 

(4)  )  sj{a  +  bx2)  ~3dx.     Ansr.  x{a  +  bx2)  -  *P/a  +  G     Use  E. 

dx  C 

,  i.e.     x-s(-a2  +  x2)~ldx.     Hint.  Use  C.    m  =  -3,  6=1, 


(5) 


J  X:isJ& 


Jx2  —  a2      1 
n=2,  p=  -b.     Ansr.  ^^     +^sec 

O 


210  HIGHER  MATHEMATICS.  §  75. 

(6)  Renyard  (Ann.  Chim.  Phys.,  [4],  19,  272,  1870),  in  working  out  a 
theory  of  electro-dynamic  action,  integrated  j(a?  +  x2)~  Idx.  Hence  ra=0, 
n  =  2,  b  =  1,  a  =  a2,  p  =  -  |.     Use  E.     Ansr.     x(a2  +  x2)  -ija2  +  C. 

(7)  To  show  that  it  is  possible  to  integrate  the  expression 

\{an~l  -  xn-^-lxfa-Vdx (2) 

when  n  =  . . .  £ ,  £  -  1,  f,  £ ,  . . . ;  and  when  n  =  . . .  f ,  % ,  £,  0,  2,  f , . . .,  substi- 
tute z  =  a  +  bx™  in  (1).     We  get 

m  +  1  m  +  1      t 

n-16       M    J"(s  -  a)    n        s"<kr (3) 

If  (w  +  l)jn  be  a  positive  integer,  the  expression  in  brackets  can  be  expanded 
by  the  binomial  theorem,  and  integrated  in  the  usual  manner.  By  compar- 
ing (2)  with  (3),  it  is  easy  to  see  that  (2)  can  be  integrated  when 

m  +  1  _  j(n  -  1)  +  1  1  1 

n  n  -  1  n  -  1  "*"  2'      *        *        '        w 

is  a  positive  integer.  From  B  and  C,  the  integral  (2)  depends  upon  the 
integral 

jxm~n(a  +  bxn)Pdx;  or,  jxm  +  n(a  +  bxn)rdx.        .        .        (5) 

By  the  substitution  of  m  -  n,  and  m  +  n  respectively  for  m  in  (m  +  l)/n,  and 
comparison  with  (2),  we  find  that  (2)  can  be  integrated  when 

m-n+1  _  ^{n  -  1)  -  (n  -  1)  +  1  _      1      _  1 

n  n  -  1  ~  n  -  1      2'  *        *        ^ 

is  a  positive  integer ;  or  else  when 

m  +  n  +  1    -  £fo  -  1)  +  (n  -  1)  +  1    ,       1  3 

n  n  -  1  n  -  1  +  2'  *        *        [  ' 

is  a  positive  integer.  But  (7)  can  be  reduced  to  either  (4)  or  (6)  by  subtracting 
unity,  and  since  integration  by  parts  can  be  performed  a  finite  number  of 
times,  we  have  the  condition  that 

5^T-5-      ;....      (8) 

be  a  positive  or  negative  integer,  or  zero,  in  order  that  (2)  may  be  integrated  ; 
in  other  words,  we  must  have 

^^--^  =  0,1,2,3,...;  =-  1,-2,-3,...    .        .        (9) 

Similarly,  by  substituting  x  =  z~1,  in  (1),  we  obtain  -z~nP-~m~2(azn  +  b)Pdz. 
As  with  (3)  and  (4),  this  can  be  integrated  when 

m  +  1      -np-m-2  +  1  m+1 

-Hr-= n =  ~-n—-P        '        '       (10> 

is  a  positive  integer.  From  D  and  E,  and  with  the  method  used  in  deducing 
(8),  we  can  extend  this  to  cases  where 

or,  where  -  (n  —  l)-1  is  a  positive  or  negative  integer.  Equating  this  to 
1,  2,  3, ... ;  and  to  -  1,  -  2,  -  3, ...  we  get,  with  (9),  the  desired  values  of  n. 
Notice  that  we  have  not  proved  that  these  are  the  only  values  of  n  which  will 
allow  (2)  to  be  integrated. 


§  75.  THE  INTEGRAL  CALCULUS.  211 

•  The  remainder  of  this  section  may  be  omitted  until  required. 
If  n  be  a  positive  integer,  we  can  integrate  JsinM#  .  dx  by  putting, 
u  =  sinn  ~1x  v  =  -  cos  x. 

du  =  (n  -  1)  sinn  ~  2x  cos  x.dx  \  dv  =  sin  x . dx. 
.•.  Jsinn#  .dx  =  -  sin*  _  *x .  cos  x  +  (w  -  1)  Jsinn  _  2# .  cos2# .  dx ; 
=«  -  sinM  ~  Yx  .  cos  x  +  (n  -  1)  Jsinn  ~  2x(l  -  sin2#)d# ; 
=  -  sinn  _  lx .  cos  x  +  (n  -  1)  Jsinn  _  2# .  dx  -  (n  -  1)  Jsinn#  .  dx. 
Transpose  the  last  term  to  the  first  member  ;  combine,  and  divide 
by  n.     The  result  is 

J                        sin"  -  lx  .  cos  x    n  -  1  f  .        „       ,  „. 

sinn#.cto  =  - + sm" - 2x  .  dx.       (12) 

Integrating  Jcosna: .  dx  by  parts,  by  putting  u  =  cosn  ~lx\  dv  =  cos 
<c .  d#,  we  get 

J                    sin  x .  cosn  ~  lx    n  -  1  f  „      ,  „  „ 

cosn# .  d#  = + cos"  -  2x .  dx.         (13) 

Eemembering  that  cos|7r  =  cos  90°  =  0 ;  and  that  sin  0°  =  0,  we 

can  proceed  further 

(V  n  -  1  ft* 

sinn#  .  dx  = sinM  -  2x .  dx. 

Jo  n     Jo 

Now  treat  n  -  2  the  same  as  if  it  were  a  single  integer  n. 

fi*  n  -  3^ 

,\  I     sinn - 2x . dx  = o       8inn-*x  .dx. 

Jo  n  -  2J0 

Combine  the  last  two  equations,  and  repeat  the  reduction.     Thus, 

we  get  finally 

P*^*^     (»-l)(»-8)-8.1(l*  (n-l)(n-8)...8.1, 

J0  «n«to-  «(»-2)...*.a  J,  da:=    n(n-a)...4.a  -a-  <F> 

when  n  is  even 

Pffsin^-^-1)(^3)--'2f^  8in«fa-(n"1)(n"8)"'2        iA 
j^  sm^-      w(w_2)...3    J0    slna^-      w(»-2)...3     '      (U) 

when  w  is  odd.     If  we  take  the  cosine  integral  (3)  above,  and  work 
in  the  same  way,  we  get 

V  (w-l)(w-3) .  ..3.1    tt 

iBeos^fa-  ^_2).,.4,2        -g;     •        •       (H) 

if  n  is  even,  and 

J0  cos^-  W(W_  2)...  5. 3  '  *  '  W 
if  n  is  odd.  Test  this  by  actual  integration  and  by  substituting 
n  =  1,  2,  3,  . . .  Note  the  resemblance  between  H  with  F,  and  I 


j: 


212  HIGHER  MATHEMATICS.  §  76. 

with  G.  The  last  four  reduction  formulae  are  rather  important  in 
physical  work.  They  can  be  employed  to  reduce  joo3nxdx  or 
\sinnxdx  to  an  index  unity,  or  |-7r. 

If  n  is  greater  than  unity,  we  can  show  that 

(V  n  -  lfi* 

smmx  .  cosnxdx  =  — ; —  I     sinm#  .  cosn  _  2xdx ;       .     (J) 
Jo  m  +  n}0  v  ' 

by  integration  by  parts,  using  u  =  sin771-1^,  dv  =  cosw#.  ^(cossc). 
If  m  is  greater  than  unity,  it  also  follows  that 

(V  m  -  lf^ 

I    sinma; .  cosnxdx  =  — ; —  I     sinm  _  2x .  cosnxdx.       .    (K) 

Jo  m  +  n)0  v    ' 

fin  1        fhn  1 

Examples. — (1)  Show  /      sin x  .  cos xdx  =  x ;    /      8in2a; .  cos xdx  =  ~. 

[iir  5  fin  2 

(2)  I      sin6a;da;  =  qott;  /      sin30.  de  =  5. 

Air  1       ru  w 

(3)  /      sin  x  .  cos2aKfoc  =  s  ;    I      sin2a; .  cos2cc<&c  =  =-~. 

§  76.   Integration  by  Resolution  into  Partial  Fractions. 

Fractions  containing  higher  powers  of  x  in  the  numerator  than 
in  the  denominator  may  be  reduced  to  a  whole  number  and  a 
fractional  part.     Thus,  by  division, 


Cx5 .  dx     (7  0  x     \ , 

)x^n:=){x  -x+x^i)dx- 


The  integral  part  may  be  differentiated  by  the  usual  methods, 
but  the  fractional  part  must  often  be  resolved  into  the  sum  of  a 
number  of  fractions  with  simpler  denominators,  before  integration 
can  be  performed. 

We  know  that  f-  may  be  represented  as  the  sum  of  two  other 
fractions,  namely  ^  and  ^,  such  that  -|  =  i  +  f-  Each  of  these 
parts  is  called  a  partial  fraction.  If  the  numerator  is  a  com- 
pound quantity  and  the  denominator  simple,  the  partial  fractions 
may  be  deduced,  at  once,  by  assigning  to  each  numerator  its  own 
denominator  and  reducing  the  result  to  its  lowest  terms.     E.g., 

x2  +  x  +  1   x2    ,z    r""i   1_    1 

x3         ~  x3    xz    x3~x    x2     xz" 
When  the  denominator  is  a  compound  quantity,  say  x2  -  x,  it 
is  obvious,  from  the  way  in  which  the  addition  of  fractions  is  per- 
formed, that  the  denominator  is  some  multiple  of  the  denominator 
of  the  partial  fractions  and  contains  no  other  factors.     We  there- 


§  76.  THE  INTEGRAL  CALCULUS.  213 

fore  expect  the  denominators  of  the  partial  fractions  to  be  factors 
of  the  given  denominator.  Of  course,  this  latter  may  have  been 
reduced  after  the  addition  of  the  partial  fractions,  but,  in  practice, 
we  proceed  as  if  it  had  not  been  so  treated. 

To  reduce  a  fraction  to  its  partial  fractions,  the  denominator 
must  first  be  split  into  its  factors,  thus  :  x2  -  x  is  the  product  of 
the  two  factors  :  x,  and  x  -  1.  Then  assume  each  factor  to  be 
the  denominator  of  a  partial  fraction,  and  assign  a  certain  indeter- 
minate quantity  to  each  numerator.  These  quantities  may,  or 
may  not,  be  independent  of  x.  The  procedure  will  be  evident 
from  the  following  examples.  There  are  four  cases  to  be  con- 
sidered. 

Case  i.  —  The  denominator  can  be  resolved  into  real  unequal 
factors  of  the  type  : 

(a  -  x)  (b  -  xy  .        .        •        (1) 

By  resolution  into  partial  fractions,  (1)  becomes 

1 A_        B        A(b  -  x)  +  B{a  -  x) 

(a  -  x)  (b  -  x)  ~  a  -  x  +  b  -  x  ~       (a  -  x)  (b  -  x)      * 

1  Ab  +  Ba  -  Ax  -  Bx 

•  ^= 

"  (a  -  x)  (b  -  x)  (a  -  x)  (b  -  x) 

We  now  assume — and  it  can  be  proved  if  necessary — that  the 
numerators  on  the  two  sides  of  this  last  equation  are  identical,1 


1  An  identical  equation  is  one  in  which  the  two  sides  of  the  equation  are  either 
identical,  or  can  be  made  identical  by  reducing  them  to  their  simplest  terms.     E.g., 

ax2  +  bx  +  c  =  ax2  +  bx  +  c ;  (a  -  x)/(a  -  x)2  =  lj(a  -  x), 
or,  in  general  terms, 

a  +  bx  +  ex2  +. . .  =  a'  +  b'x  +  c'x2  +... 
An  identical  equation  is  satisfied  by  each  or  any  value  that  may  be  assigned  to  the 
variable  it  contains.  The  coefficients  of  like  powers  of  x,  in  the  two  numbers,  are  also 
equal  to  each  other.  Hence,  if  x  =  0,  a  =  a'.  We  can  remove,  therefore,  a  and  a' 
from  the  general  equation.  After  the  removal  of  a  and  a',  divide  by  x  and  put  x  =  0, 
hence  b  =  b' ;  similarly,  c=&,  etc.  For  fuller  details,  see  any  elementary  text-book 
on  algebra.  The  symbol  "■  =  "  is  frequently  used  in  place  of  *'="  when  it  is  desired 
to  emphasize  the  tact  that  we  are  dealing  with  identities,  not  equations  of  condition. 
While  an  identical  equation  is  satisfied  by  any  value  we  may  choose  to  assign  to  the 
variable  it  contains,  an  equation  of condition  is  only  satisfied  by  particular  values  of  the 
variable.  As  long  as  this  distinction  is  borne  in  mind,  we  may  follow  customary  usage 
and  write  "  =  "  when  "  =  "  is  intended.  For  M  =  "we  may  read,  "  may  be  trans- 
formed into.,  .whatever  value  the  variable  x  may  assume"  ;  while  for  "  =,"  we 
must  read,  "  is  equal  to .  . .  when  the  variable  x  satisfies  some  special  condition  or 
assumes  some  particular  value  ". 


214  HIGHER  MATHEMATICS.  §  76. 

Ab  +  Ba  -  Ax  -  Bx  =  1. 

Pick  out  the  coefficients  of  like  powers  of  x,  so  as  to  build  up 
a  series  of  equations  from  which  A  and  B  can  be  determined.  For 
example, 

Ab  +  Ba  =  1 ;  x(A  +  B)  =  0 ;  .-.  A  +  B  =  0 ;  .'.  A  =  -  B  ; 

r'A  =  b^a?  •••£  =  -  b^~a 
Substitute  these  values  of  A  and  B  in  (1).- 

I =  _2_.^ L_._i_.  (2) 

(a  -  x)  (b  -  x)      b  -  a   a  -  x      b  -  a   b  -  x 

An  alteknative  method,  much  quicker  than  the  above,  is 
indicated  in  the  following  example :  Find  the  partial  fractions  of 
the  function  in  example  (3)  below. 

1 •        A  B  G 

(a  -  x)  (b  -  x)  (c  -  x)  ~  a  -  x      b  -  x      c  -  x '  " 
Consequently, 

(b  -  x)  (c  -  x)A  +  (a  -  x)  (c  -  x)B  +  (a  -  x)  (b-x)C  =  1. 
This  identical  equation  is  true  for  all  values  of  x,  it  is,  therefore,  true 

1 
(b  -  a)  (c  -  a)' 

1 
(c  -  b)  {a  -  b) ' 

1 
(a  -  c)  ( b  -  cj 
Examples. — (1)  In  studying  bimolecular  reactions  we  meet  with 

J(a-x)(b-x)  ~  J(b-a)(a-x)      J(b-a){b-x)  ~  F^l'  °ga~^x  +  C' 
(2)  J.  J.  Thomson's  formula  for  the  rate  of  production  of  ions  by  the 
Rontgen  rays  is  dx/dt  =  q-ax2.     Remembering  that  a  -  x2  =  ( si  a  -  x)  ( si  a  +  x) ; 
show  that  if  we  put  q\a  =  &2,  for  the  sake  of  brevity,  then 

1        b  +  x 

(a  -x)  (b  -  x)  (c  -  x)'    Keep  y°Ur  answer  for  U8e  later  on' 
f      dx  1    ,      a  +  bx      „ 

(4)  Show  that  J  jc^  =  ^ab  lo%  a~^bx  +  G' 

(5)  If  the  velocity  of  the  reaction  between  bromic  and  hydrobromic  acids 
is  represented  by  the  equation :  dx/dt  =  k(na  +  x)  (a  -  x),  then  show  that 

,na  +  x      „ 


when  x  =  a,  . 

\  {b  -  a)(c  -  a)A  =  1 ; 

.-.  A  = 

when  x  =  b,  . 

\  (c  -  b)  (a  -  b)B  =  1 ; 

.-.  5  = 

when  x  =  c,  . 

\  {a  -  c)  {b  -  c)C  =  1; 

.'.  Om 

(n  +  l)at         a  -  x 


§76.  THE  INTEGRAL  CALCULUS.  215 

,„,    T,  dx      , ,  ■..  ,    . ,        „    ,  ,         2-3026       ,        a  +  x 

(6)  H  to  m  k(a  +  x)  (na  -  x) ;  show  that  k  =  (n  +  l)at  .  log1(&T-- 

(7)  S.  Arrhenius,  in  studying  the  hydrolysis  of  ethyl  acetate,  employed  the 
integral. 

f  1  +  mx  -  nx2  ,  /T"  1  +  nab  +  \m  -  n(a  +  b))x~~\ , 

J  {a-x)(b-x)d*  •  ■■■  "  j  L  "  "  + (<.-«)(»-.)  > 

Substitute  jp  =  1  +  nab  ;  q  =  m  -  n(a  +  b),  then,  by  the  method  of  partial 
fractions,  show  that 


P  +  9.X         ,       P  +  aq.      ■  .      p+bq 

(a-x)(b-x)dx  =  TTTl0^a  "  »)  -¥^6  loo(6  "  »>  +  C* 


/ 

/da;  1  x 

■  ia  _  x\  =  ~  lo8  a  _  x  +  G  are  very  common  in  chemi- 
cal dynamics — autocatalysis. 

(9)  H.  Danneel  (Zeit.  phys.  Chem.,  33,  415,  1900)  has  the  integral 


kt 


f  xdx  1  .      a2  -  x?  ^ 


if  x  =  »lf  when  f  =  ^ ;  and  x  =  a52>  when  t  =  J* 

(10)  R.  B.  Warder's  equation  for  the  velocity  of  the  reaction  between 
chloroacetic  acid  and  ethyl  alcohol  is 

§  =  ak{l  -  (1  +  b)y}  {1  -  (1  -  b)y}.     .-.  log*  ~_  |*  "  ^  =  2o6W. 

Case  ii. — T/ie  denominator  can  be  resolved  into  real  factors 
some  of  which  are  equal.     Type : 

1 
(a  -  xf{b  -  x) 
The  preceding  method  cannot  be  used  here  because,  if  we  put 
1  A      .      S  0         A  +  B         C 


(a  -  x)2(b  -  x)  a  -  x  a  -  x  b  -  x  a  -  x  b  -  x 
A  +  B  must  be  regarded  as  a  single  constant.  Reduce  as  before 
and  pick  out  coefficients  of  like  powers  of  x.  We  thus  get  three 
independent  equations  containing  two  unknowns.  The  values  of 
A,  B  and  C  cannot,  therefore,  be  determined  by  this  method.  To 
overcome  the  difficulty,  assume  that 

1  A  B  G 

(a  -  x)2(b  -  x)  ~  {a  -  x)2      a  -  x      b  -  x 
Multiply  out  and  proceed  as  before,  the  final  result  is  that 


A-A-.t*--'^,*--     X 


b  -  a  '  (b-a) 


Examples. — (1)  H.  Goldschmidt  represents  the  velocity  of  the  chemical 
reaction  between  hydrochloric  acid  and  ethyl  alcohol,  by  the  relation 
dxjdt  =  k{a  -  x)  {b  -  x)2.     Hence, 


216  HIGHER  MATHEMATICS.  §  76. 

k  f     _  f  dx 1       f  [(a  -  b)dx  _   f  dx  C  dx   \ 

J         J(a-x)(b-x)*-  (a-  6)2\J    {b  -  xf       Jb-x+  Ja-xf' 
Integrate.    To  find  a  value  for  0,  put  x  =*  0  when  t  =  0.     The  final  result  is 

<2>  ShoyfJx*(a  +  bx)  =  a*l°Z—a—  ~  m  +  C' 

rat  at,        f  ^  *,      g  +  1      !        !  /, 

(3>  Show  J  (x  -  l)»(aj  +  1)  =  I loS  x^Tl  ~  2  '  5^~1  +  °'      An    exPressl0n 

used  by  W.  Meyerhofer,  Zeit.  phys.  Chem.,  2,  585,  1838. 

...   __         f  xdx  1        f         a(b  -  x)      x(a  -  b)} 

<*>  sll0WJ(o-*)(6-*)2  ~  5!^Tp{«»»g j^r^  +  ^j. 

for  values  of  x  from  a;  =  a?  to  x  =  0.     (H.  Kiihl,  Zeit.  phys.  Chem.,  M,  385, 
1903.) 

(5)  P.  Henry  (Zeit.  phys.  Chem.,  10,  96,  1892)  in  studying  the  phenomenon 
of  autocatalysis  employed  the  expression 

dj=h(a-  x)  (\/4:K(a  -  x)  +  K2  -  K). 
To  integrate,  put  4:K(a  -x)  +  K?=s2  ;.-.a-x={zi-  K?)j±K ;  dx=  -  z .  dzfiK. 
z.dz  kdt  1  s-iT    11  &$ 

;  •'•r^1°g  rr^-o    .— »•  =  -  o"  +  G 


(s-Z)  (^-Z2)-       2    '  '-iKx"*z  +  K~2     z-K~~2 


Now  put  P  =  sJIKia  -  s)  +  Z2 ;  Q  =  ^Za  +  K\  and  show  that  if  x  m  0 
when£  =  0, 


M 


Q-p      _L  i  1ng(p+Jg)(g-g)    , 
(p-Z)(Q-z)+2Z10g(P-Z)  (g+z)-** 


For  a  more  complex  example  see  T.  S.  Price,  Journ.  Chem.  Soc,  79,  314,  1901. 
(6)  J.  W.  Walker  and  W.  Judson's  equation  for  the  velocity  of  the  chemical 
reaction  between  hydrobromic  and  bromic  acids,  is 

dx  1  f      1  1 ) 

The  reader  is  probably  aware  of  the  fact  that  he  can  always 
prove  whether  his  integration  is  correct  or  not,  by  differentiating 
his  answer.     If  he  gets  the  original  integral  the  result  is  correct. 

Case  iii.  —  The  denominator  can  be  resolved  into  imaginary 
factors  all  unequal.     Type  : 

1 
(a2  +  x2)  {b  +  x)' 

Since  imaginary  roots  always  occur  in  pairs  (page  353),  the 
product  of  each  pair  of  imaginary  factors  will  give  a  product  of  the 
form,  x2  +  a2.  Instead  of  assigning  a  separate  partial  fraction  to 
each  imaginary  factor,    we   assume,  for  each  pair   of   imaginary 


§  76.  THE  INTEGRAL  CALCULUS.  217 

factors,  a  partial  fraction  of  the  form  : 

Ax  +  B 

a2  +  x2' 

Hence  we  must  write 

1  Ax  +  B        G 


{a2  +  x2)  (b  +  x)     a2  +  x2  T5  +  x% 
Now  get  (13),  page  193,  fixed  upon  your  mind. 

ExAmp^S.-(1)/(^1)^  +  1)=/(^2  +  ^+^)^.       Here 

A  =  $;B=-l;  C  =  £;  Z>=0.      Ansr   J  log  (a;2  +  l)  (s-1)  ~2-£  (as-1)  -*+0. 

(2)  Show  Jr^=^ tan-1* +^  log  ^~+C- 

(3)  H.  Danneel  (^ei^.  phys.  Chem.y  33,  415,  1900)  used  a  similar  expres- 
sion in  his  study  of  the  "  Free  Energy  of  Chemical  Reactions  ".     Thus,  he  has 

x2dx       ,  ,  1/  x„  x,  \      1  (x0  —  a)  (x,+a) 

-j -x  =  Mt.    .-.  2/^ -*,)  =  -(  tan ~1-^- tan"1-1    +r  log  F ]        . 

a4-*1  V2      u     a\  a  a)     2a     B  (x.2  +  a)  (x1-a) 

in  an  experiment  where  x  =  xi  when  t  =  t^\  and  x=x2  when  £=£2. 

/da; 
>    _  bx)2i3(c  -  a?l   bas  to  be  8olved  wnen  the  rate 

of  dissolution  of  a  spheroidal  solid  is  under  discussion.     Put  a  -  bx  =  s3 ; 

a-6c=w3;  .*.  x=(a-z3)lb;  dx=  -Bz2dz/b.      Substitute  these  results  in  the 

given  integral,  and  we  get 

f   dz     _     /" dz 1  f  dz       1  j"  (z  +  2n)dz 

JriF^z"  ~    J  (n  -  z)  (n2+nz  +  z2)  =n2J  n-z  +  n2]  n2+nz  +  z2 ' 

by  resolution  into  partial  fractions.    Let  me  make  a   digression.     Obviously, 
we  may  write 

f(y  +  2b)dy    Iff     2y  +  b  36        \ 

Ja  +  by  +  y*-2j\a  +  by  +  y*  +  a  +  by  +  yyay' 
The  numerator  of  the  first  fraction  on  the  right  is  the  differential  coefficient 
of  the  denominator ;  and  hence,  its  integral  is  £  log  (a+by  +  y2) ;  the  integral 
of  the  second  term  of  the  right  member  is  got  by  the  addition  and  subtraction 
of  %b2  in  the  denominator.     Hence, 

f       dV  f  dy 2        fnn-i    2y  +  b 

J  a+by  +  y2-  J  (a  -  $2)  +  (y  +  $b)2~  .J^^b2  s/la'-b2 

Returning  to  the  original  problem,  we  see  at  once  that 


0f    dz        if         ,J<n?  +  nz  +  z2\  2z+n     „ 

SJW^=n2{l°Z         n-z        j+^/3tan'1^:+a 

Now  restore  the  original  values,  z  =  (a  -  bx)\,  and  n  =  (a  -  bc)$. 

In  most  of  the  examples  which  I. have  selected  to  illustrate  my 
text,  the  denominator  of  the  integral  has  been  split  up  into  factors  so 


218  HIGHER  MATHEMATICS.  §  77. 

as  not  to  divert  the  student's  attention  from  the  point  at  issue.  If 
the  student  feels  weak  on  this  subject  a  couple  of  hours'  drilling 
with  W.  T.  Knight's  booklet  on  Algebraic  Factors,  London,  1888, 
will  probably  put  things  right. 

Case  iY. — The  denominator  can   be  resolved  into   imaginary 
factors,  some  of  which  are  equal  to  one  another.     Type: 

1 
(a2  +  x2)\b  +  x)' 
Combining  the  preceding  results, 

1  Ax  +  B      Cx  +  D       E 


(a2  +  x2)2(b  +  x)~(a2  +  x2)2^  a2  +  x2  ^b  +  x 
In  this  expression,  there  are  just  sufficient  equations  to  determine 
the  complex  system  of  partial  fractions,  by  equating  the  coefficients 
of  like  powers  of  x.  The  integration  of  many  of  the  resulting 
expressions  usually  requires  the  aid  of  one  of  the  reduction 
formula  (§  76). 


Example. 


[(a?  +  x-l)dx     fxdx        f     dx 
Frove  J      (a2  +  l)2     =J  xUTJ  -J(x*  +  l)2' 


Integrate.    Ansr.  J  log  {x2  +1)  -  %xj(l  +  x2)  +£tan_1a;  +  C.    Use  formula  E, 
page  209,  for  evaluating  the  last  term. 

Consequently,  if  the  denominator  of  any  fractional  differential 
can  be  resolved  into  factors,  the  differential  can  be  integrated  by 
one  or  other  of  these  processes.  The  remainder  of  this  chapter 
will  be  mainly  taken  up  with  practical  illustrations  of  integration 
processes.  A  number  of  geometrical  applications  will  also  be  given 
because  the  accompanying  figures  are  so  useful  in  helping  one  to 
form  a  mental  picture  of  the  operation  in  hand. 

§  77.    The  Yelocity  of  Chemical  Reactions. 

The  time  occupied  by  a  chemical  reaction  is  dependent,  among 
other  things,  on  the  nature  and  concentration  of  the  reacting  sub- 
stances, the  presence  of  impurities,  and  on  the  temperature.  With 
some  reactions  these  several  factors  can  be  so  controlled,  that 
measurements  of  the  velocity  of  the  reaction  agree  with  theoretical 
results.  But  a  great  number  of  chemical  reactions  have  hitherto 
defied  all  attempts  to  reduce  them  to  order.  For  instance,  the 
mutual  action  of  hydriodic  acid  and  bromic  acid  ;  of  hydrogen  and 
oxygon ;  of  carbon  and  oxygen  ;  and  the  oxidation  of  phosphorus. 


§  77.  THE  INTEGRAL  CALCULUS.  219 

The  magnitude  of  the  disturbing  effects  of  secondary  and  catalytic 
actions  obscures  the  mechanism  of  such  reactions.  In  these  cases 
more  extended  investigations  are  required  to  make  clear  what 
actually  takes  place  in  the  reacting  system. 

Chemical  reactions  are  classified  into  uni-  or  mono-molecular, 
bi-molecular,  ter-  or  tri-molecular,  and  quadri-molecular  reactions 
according  to  the  number  of  molecules  which  are  supposed  to  take 
part  in  the  reaction.  Uni-,  bi-,  ter-, . . .  molecular  reactions  are 
often  called  reactions  of  the  first,  second,  third, . . .  order. 

I. — Beactions  of  the  first  order.  Let  a  be  the  concentration  of 
the  reacting  molecules  at  the  beginning  of  the  action  when  the 
time  t  =  0.  The  concentration,  after  the  lapse  of  an  interval  of 
time  t,  is,  therefore,  a  -  x,  where  x  denotes  the  amount  of  sub- 
stance transformed  during  that  time.  Let  dx  denote  the  amount 
of  substance  formed  in  the  time  dt.  The  velocity  of  the  reaction, 
at  any  moment,  is  proportional  to  the  concentration  of  the  reacting 
substance — Wilhelmy's  law — hence  we  have 

gj-  k(a  -  »);  or,  k  =  j  .log^^;  •        •        W 

or,  what  is  the  same  thing,  x  =  a(l  -  e  ~  **),  where  A;  is  a  constant 
depending  on  the  nature  of  the  reacting  system.  Reactions  which 
proceed  according  to  this  equation  are  said  to  be  reactions  of  the 
first  order. 

II. — Beactions  of  the  second  order.  Let  a  and  b  respectively 
denote  the  concentration  of  two  different  substances,  say,  in  such 
a  reacting  system  as  occurs  when  acetic  acid  acts  on  alcohol,  or 
bromine  on  fumaric  acid,  then,  according  to  the  law  of  mass 
action,  the  velocity  of  the  reaction  at  any  moment  is  proportional 
to  the  product  of  concentration  of  the  reacting  substances.  For 
every  gram  molecule  of  acetic  acid  transformed,  the  same  amount 
of  alcohol  must  also  disappear.  When  the  system  contains  a  -  x 
gram  molecules  of  acetic  acid  it  must  also  contain  b  -  x  gram 
molecules  of  alcohol.     Hence 

dx       t  /  ^  ,i         x  ,       1       1     ,      (a  -  x)b 

<*r- *(* - *) (*  -  *y>  •••  * =T'a^blo%(b^iJja--    (2) 

Reactions  which  progress  according  to  this  equation  are  called 
reactions  of  the  second  order.  If  the  two  reacting  molecules  are 
the  same,  then  a  =  b.  From  (2),  therefore,  we  get  log  1  x  £  =  0  x  oo. 
Such  indeterminate  fractions  will  be  discussed  later  on.     But  if  we 


220  HIGHER  MATHEMATICS.  §  77. 

start  from  the  beginning,  we  get,  by  the  integration  of 

§_*(a_a),;i_i._£L_    .     .     (3) 

In  the  hydrolysis  of  cane  sugar, 

^12^22^11  +  H20  =  2C6H1206, 
let  a  denote  the  amount  of  cane  sugar,  b  the  amount  of  water 
present  at  the  beginning  of  the  action.  The  velocity  of  the  re- 
action can  therefore  be  represented  by  the  equation  (3),  when  x 
denotes  the  amount  of  sugar  which  actusrily  undergoes  transforma- 
tion. If  the  sugar  be  dissolved  in  a  large  excess  of  water,  the 
concentration  of  the  water,  b,  is  practically  constant  during  the 
whole  process,  because  b  is  very  large  in  comparison  with  x,  and 
a  small  change  in  the  value  of  x  will  have  no  appreciable  effect 
upon  the  value  of  b  ;  b  -  x  may,  therefore,  be  assumed  constant. 
.-.  k'  =  k(b  -  x),  where  k'  and  k  are  constant.  Hence  equation  (1) 
should  represent  the  course  of  this  reaction.  Wilhelmy's  measure- 
ments of  the  rate  of  this  reaction  shows  that  the  above  supposition 
corresponds  closely  with  the  truth.  The  hydrolysis  of  cane  sugar 
in  presence  of  a  large  excess  of  water  is,  therefore,  said  to  be  a 
reaction  of  the  first  order,  although  it  is  really  bimolecular. 

Example. — Proceed  as  on  page  59  with  the  following  pairs  of  values  of 
x and  t : — 

*-     15,  30,  45,  60,  75,... 

x  =  0-046,         0-088,         0-130,  0-168,        0-206,... 

Substitute  these  numbers  in  (1) ;  show  that  k'  is  constant.  Make  the  proper 
changes  for  use  with  common  logs.     Put  a  =  1. 

III. — Beactions  of  the  third  order.     In  this  case  three  molecules 

take  part  in  the  reaction.     Let  a,  b,  c,  denote  the  concentration  of 

the  reacting  molecules  of  each  species  at  the  beginning  of  the 

reaction,  then, 

dx 

-£-=  k(a  -  x)  (b  -  x)  (c  -  x).  .        .         (4) 

Integrate  this  expression  and  put  x  =  0  when  t  =  0  in  order  to 
find  the  value  of  G.  The  final  equation  can  then  be  written  in  the 
form, 


[V     a    y-y     b    y-y    c    \~b\ 

'"\v%  -  x)      \b  -  x)      \c  -  x)       j 
t(a  -  b)  (b  -  c)  (c  -  a) 
where  a,  b,  c,  are  all  different.     If  we  make  a  =  b  =  c,  in  equation 


loc 

7.  y\tv   —    -u}/         \u    —    jo/         \u   —   -jo/         )  /c\ 

t(a  -  b)  (b  -  c)  (c  -  a) 


§  77.  THE  INTEGRAL  CALCULUS.  221 

(4)  and  integrate  the  resulting  expression 

^-k(a    xY-h-H-^—    11-  *<2fl-*>-  (6) 

dt-tc[a-x)   ,  «- 2t^a _ xy    a.2j  - 2ta^a _ xy.  {») 

By  rearranging  the  terms  of  equation  (6)  so  that, 

-i1  ~  jsktd-  •     •     •     (7) 

we  see  that  the  reaction  can  only  come  to  an  end  (x  =  a)  after  the 
elapse  of  an  infinite  time,  t  =  oo.  If  o  =  b  when  a  is  not  equal 
to  b, 

-       1  1       f  (a  -  6)s       .      a(b  -  s)l 

*  *  r  (a  -  6)»  (6(6  -  a?)  +  log  b{a  -  x)j '      '        {  } 

Among  reactions  of  the  third  order  we  have  the  polymerization 
of  cyanic  acid,  the  reduction  of  ferric  by  stannous  chloride,  the 
oxidation  of  sulphur  dioxide,  and  the  action  of  benzaldehyde  upon 
sodium  hydroxide.  For  full  particulars  J.  W  Mellor,  Chemical 
Statics  and  Dynamics,  might  be  consulted. 

IV. — Beactions  of  the  fourth  order.  These  are  comparatively 
rare.  The  reaction  between  hydrobromic  and  bromic  acids  is, 
under  certain  conditions,  of  the  fourth  order.  So  is  the  reaction 
between  chromic  and  phosphorous  acids ;  the  action  of  bromine 
upon  benzene ;  and  the  decomposition  of  potassium  chlorate. 
The  general  equation  for  an  w-molecular  reaction,  or  a  reaction 
of  the  nth.  order  is 

dx     w         %       7      1        1     f        1  11 

3r-*frrfFi  h  =  l-n—i\{a-xY-i-^)'-      ^ 

The  intermediate  steps  of  the  integration  are,  Ex.  (3)  and  Ex.  (4), 
page  196.  The  integration  constant  is  evaluated  by  remembering 
that  when  x  =  0,  t  =  0.     We  thus  obtain 

(n-l)(a-x)n-i  =  kt  +  G>  G=+(n-l)an-1> 

V. — To  find  the  order  of  a  chemical  reaction.  Let  Clt  G2  re- 
spectively denote  the  concentration  of  the  reacting  substance  in 
the  solution,  at  the  end  of  certain  times  tx  and  t2.  From  (9),  if 
0=  Gv  when  t  =  tv  etc., 

-a? " k0";  •'•  ^Mnpi-api}-^-^     (io) 

where  n  denotes  the  number  of  molecules  taking  part  in  the  re- 


-1 


222  HIGHER  MATHEMATICS.  §  77. 

action.     It  is  required  to  find  a  value  for  n.     From  (10) 

°2dC      , .  .       log  L  -  log  t9  ,-.  1 N 

-^  =  kt;  or,  n=l  +  -2_J _*— 2.      .         (11) 

<?iG  log  u2  -  log  Uj 

Why  the  negative  sign  ?  The  answer  is  that  (10)  denotes  the  rate 
of  formation  of  the  products  of  the  reaction,  (11)  the  rate  at  which 
the  original  substance  disappears.     G  —  a-x,  .-.  dC=  -dx. 

Numerical  Illustration. — W.  Judson  and  J.  W.  Walker  (Journ.  Chem. 
Soc,  73,  410,  1898)  found  that  while  the  time  required  for  the  decomposition 
of  a  mixture  of  bromic  and  hydrobromic  acids  of  concentration  77,  was 
15  minutes  ;  the  time  required  for  the  transformation  of  a  similar  mixture 
of  substances  in  a  solution  of  concentration  51-33,  was  50  minutes.  Substi- 
tuting these  values  in  (11), 

■       log  3-333      n    ■ 

"  =  1+l?gT5-  =  3'97- 

The  nearest  integer,  4,  represents  the  order  of  the  reaction.  Use  the  Table  of 
Natural  Logarithms,  page  627. 

The  intervals  of  time  required  for  the  transformation  of  equal 
fractional  parts  m  of  a  substance  contained  in  two  solutions  of 
different  concentration  C±  and  G2,  may  be  obtained  by  graphic 
interpolation  from  the  curves  whose  abscissae  are  tx  and  t2  and 
whose  ordinates  are  GY  and  G2  respectively. 

Another  convenient  formula  for  the  order  of  a  reaction,  is 

n=^JLZ^lJL.    .     .     .    (i2) 

log  Gx  -  log  G2 

The  reader  will  probably  be  able  to  deduce  this  formula  for  himself. 
The  mathematical  treatment  of  velocity  equations  here  outlined 
is  in  no  way  difficult,  although,  perhaps,  some  practice  is  still  re- 
quisite in  the  manipulation  of  laboratory  results.  The  following 
selection  illustrates  what  may  be  expected  in  practical  work  if 
the  reaction  is  not  affected  by  disturbing  influences. 

Examples. — (1)  M.  Bodenstein  (Zeit.  phys.  Chem.,  29,  315,  1899)  has  the 

equation 

dx  , 

-jj  =  k(a  -  x)  (b  -  xp 

for  the  rate  of  formation  of  hydrogen  sulphide  from  its  elements.  To  inte- 
grate this  expression  put  b-x=z2,  .•.  dx  =  -  2zdx,  and  therefore 


/; 


dz  ht  2  ,  8     _ 

i=  -  -pr :  or,  -j  tan  —  1-j  +  C  =  -  Jet 


+  A*~~  2'  U1'  Avai±       A 
where  ,42  =  a-6.     For  the  integration,  see  (13),  page  193.    This  presupposes 


§  77.  THE  INTEGKAL  CALCULUS.  223 

that  a>6  ;  if  a<6  the  integration  is  Case  i.  of  page  213.  We  get  a  similar 
expression  for  the  rate  of  dissolution  of  a  solid  cylinder  of  metal  in  an  acid. 
To  evaluate  C,  note  that  x=0  when  £=0. 

(2)  L.  T.  Reicher  (Zeit.  phys.  Chem.,  16,  203,  1895)  in  studying  the  action 
of  bromine  on  fumaric  acid,  found  that  when  2=0,  his  solution  contained  8*8 
of  fumaric  acid,  and  when  2=95,  7*87  ;  the  concentration  of  the  acid  was  then 
altered  by  dilution  with  water,  it  was  then  found  that  when  t=0,  the  concen- 
tration was  3-88,  and  when  2= 132,  3*51.  Here  dCJdt=  (8-88  -  7'87)/95= 0-0106 ; 
dCJdt= 0-00227  ;  C1  =  (8-88 +  7'87)/2  =  8-375  ;  C2  =  3*7,  n  =  1-87  in  (12)  above 
The  reaction  is,  therefore,  of  the  second  order. 

(3)  In  the  absence  of  disturbing  side  reactions,  arrange  velocity  equations 
for  the  reaction  (A.  A.  Noyes  and  G.  J.  Cottle,  Zeit.  phys.  Chem.,  27,  578, 
1898) :— 2CH3 .  C02Ag  +  H .  C02Na  =  CH3 .  COOH  +  CH3 .  C02Na  +  C02+  2Ag. 
Assuming  that  the  silver,  sodium  and  hydrogen  salts  are  completely  dissociated 
in  solution,  the  reaction  is  essentially  between  the  ions : 

2Ag+  +  H.COO"  =  2Ag  +  C02  +  H  + 

therefore,  the  reaction  is  of  the  third  order.  Verify  this  from  the  following 
data.     When  a  (sodium  formate) =0-050,  b  (silver  acetate)  =0*100;  and  when 

t=     2,  4,  7,  11,  17,    ... 

(b  -  x)  x  103  =  81-03,        71-80,         63*95,  59-20,  56-25, . . . 

Show  that  if  the  reaction  be  of  the  second  order,  k  varies  from  1'88  to  2*67, 
while  if  the  reaction  be  of  the  third  order,  k  varies  between  31-2  and  28-0. 

(4)  For  the  conversion  of  acetochloranilide  into  £>-chloracetanilide,  J.  J. 
Blanksma  {Bee.  Trav.  Pays-Bos.,  21,  366,  1902  ;  22,  290,  1903)  has 

t=  0,  1,  2,  3,  4,  6,  8,...; 

a-x  =  49-3,      35-6,      25-75,     18-5,     13-8,      7'3,      4-8,... 
Show  that  the  reaction  is  of  the  first  order. 

(5)  An  homogeneous  spheroidal  solid  is  treated  with  a  solvent  which  dis- 
solves layer  after  layer  of  the  substance  of  the  sphere.  To  find  the  rate  of 
dissolution  of  the  solid.  Let  r0  denote  the  radius  of  the  sphere  at  the  be- 
ginning of  the  experiment,  when  t=0  ;  and  r  the  radius  of  the  time  t ;  let  <p 
denote  the  volume  of  one  gram  molecule  of  the  solid  ;  and  let  x  denote  the 
number  of  gram  molecules  of  the  sphere  whioh  have  been  dissolved  at  the 
time  t.  The  rate  of  dissolution  of  the  sphere  will  obviously  be  proportional 
to  the  surface  s,  and  to  the  amount  of  acid,  a  -  x,  present  in  the  solution  at 
the  time  t.  But,  remembering  that  the  volume  of  the  sphere  is  fa-r3,  the 
volume  of  the  x  gram  molecules  of  the  sphere  dissolved  at  the  time  t  will  be 

and  the  surface  s  of  the  sphere  at  the  time  t  will  be  litr*. 

an  expression  resembling  that  integrated  on  page  217,  Ex.  (4). 

(6)  L.  T.  Reicher  (Liebig's  Ann.,  228,  257,  1885 ;  232,  103,  1886)  measur- 
ing the  rate  of  hydrolysis  of  ethyl  acetate  by  sodium  hydroxide,  found  that 


when 

t=     0, 

393, 

669, 

1010 

a-  x  =  0-5638, 
b  -  x  =  03114, 

0-4866, 
0-2342, 

0-4467, 
01943, 

0-4113 
0-1589 

224  HIGHER  MATHEMATICS.  §  77. 


1265,  . . .  units ; 
0-3879,  . . . 
0-1354,  . . . 

Now  apply  these  results  to  equation  (2),  page  219,  and  show  that  k  is  ap- 
proximately constant  and  that,  in  consequence,  the  reaction  is  of  the  second 
order. 

(7)  Ethyl  acetate  is  hydrolized  in  the  presence  of  acidified  water  forming 
alcohol  and  acetic  acid.  Suppose  a  gram  molecules  of  acetic  acid  are  used  to 
acidify  the  water,  and  that  we  start  an  experiment  with  b  gram  molecules  of 
ethyl  acetate,  show  that  Wilhelmy's  law  leads  to 

dx        i  .  .  V         .  1    .       b(a  +  x) 

-^  m    \{a  +  x)  (b  -  x) ;    or,  j-logioa/&  _  A  =  constant ; 

with  the  additional  assumption  that  the  velocity  of  the  reaction  is  propor- 
tional to  the  amount  of  acetic  acid  present  in  the  system.  If  a  gram  molecules 
of  some  other  acid  are  used  as  "  catalytic  "  agent, 

dx  _     '  „         ,  1    ,       /     b       k<,a+  k,x  \ 

M  =  {k<fl  +  kxx)  (b  -  x) ;  or,  j.  log^g-^ j^—  )=  constant. 

See  W.  Ostwald,  Journ.  prakt.  Chem.,  [2],  28,  449,  1883,  for  experimental 
numbers.  Hint.  There  is  a  of  catalyzing  acid  present  and  the  velocity  of  the 
reaction  due  to  this  agent  will  be  k2a(b  -  x) ;  but  x  of  acetic  acid  has  also  been 
produced  ;  so  that  the  velocity  of  the  reaction  due  to  the  catalyzing  action  of 
acetic  acid  is  equal  to  kxx{b  -  x).  Now  apply  the  principle  of  the  mutual 
independence  of  different  reactions  of  page  70. 

(8)  It  was  once  thought  that  the  decomposition  of  phosphine  by  heat  was  in 
accordance  with  the  equation,  4PH3  =  P4  +  6H2 ;  now,  it  is  believed  that  the 
reaction  is  more  simple,  viz.,  PH3  =  P  +  3H,  and  that  the  subsequent  formation 
of  the  P4  and  H2  molecules  has  no  perceptible  influence  on  the  rate  of  the 
decomposition.  Show  that  these  suppositions  respectively  lead  to  the  follow- 
ing equations : 

^-^•••^Ho^-1}-    m  ***.  -•>:•••* 4 logrry 

In  other  words,  if  the  reaction  be  of  the  fourth  order,  k  will  be  constant,  and 
if  of  the  first  order,  k'  will  be  constant.  To  put  these  equations  into  a  form 
suitable  for  experimental  verification  let  a  gram  molecules  of  PH3  per  unit 
volume  be  taken.  Let  the  fraction  a;  of  a  be  decomposed  in  the  time  t. 
Hence,  (1  -  x)a  gram  molecules  of  phosphine  and  f  ax,  of  hydrogen  remain. 
Since  the  pressure  of  the  gas  is  proportional  to  its  density,  if  the  original 
pressure  of  PH3  be_p0  and  of  the  mixture  of  hydrogen  and  phosphine^,  then, 


_Pj     (l-x)a  +  %xa    ■      <  2px  1 


and 


HGAJ-^-N 


Po 


2ft1 

where  the  constants  are  not  necessarily  the  same  as  before.     D.  M.  Kooij 


§  78.  THE  INTEGRAL  CALCULUS.  225 

(Zeit.  phys.  Chem.,  12, 155,  1892)  has  published  the  following  data: — 
t=       0,  4,  14,  24,  46-3.    ... 

p^  758-01,  769-34         795-57,         819-16,         865-22, . . . 

Hence  show  that  ft7,  not  k  satisfies  the  required  condition.  The  decomposi- 
tion of  phosphine  is,  therefore,  said  to  be  a  reaction  of  the  first  order.  Of 
course  this  does  not  prove  that  a  reaction  is  really  unimolecular.  It  only 
proves  that  the  velocity  of  the  reaction  is  proportional  to  the  pressure  of  the 
gas — quite  another  matter.   See  J.  W.  Mellor's  Chemical  Statics  and  Dynamics. 

In  experimental  work  in  the  laboratory,  the  investigator  pro- 
ceeds by  the  method  of  trial  and  failure  in  the  hope  that  among 
many  wrong  guesses,  he  will  at  last  hit  upon  one  that  will  "  go  ". 
So  in  mathematical  work,  there  is  no  royal  road.  We  proceed  by 
instinct,  not  by  rule.  For  instance,  we  have  here  made  three  guesses. 
The  first  appeared  the  most  probable,  but  on  trial  proved  unmis- 
takably wrong.  The  second,  least  probable  guess,  proved  to  be 
the  one  we  were  searching  for.  In  his  celebrated  quest  for  the 
law  of  descent  of  freely  falling  bodies,  Galileo  first  tried  if  V,  the 
velocity  of  descent  was  a  function  of  s,  the  distance  traversed.  He 
found  the  assumption  was  not  in  agreement  with  facts.  He  then 
tried  if  V  was  a  function  of  t,  the  time  of  descent,  and  so  estab- 
lished the  familiar  law  V  =  gt.  So  Kepler  is  said  to  have  made 
nineteen  conjectures  respecting  the  form  of  the  planetary  orbits, 
and  to  have  given  them  up  one  by  one  until  he  arrived  at  the 
elliptical  orbit  which  satisfied  the  required  conditions. 

§  78.    Chemical  Equilibria — Incomplete  or  Reversible 
Reactions. 

Whether  equivalent  proportions  of  sodium  nitrate  and  potas- 
sium chloride,  or  of  sodium  chloride  and  potassium  nitrate,  are 
mixed  together  in  aqueous  solution  at  constant  temperature,  each 
solution  will,  after  the  elapse  of  a  certain  time,  contain  these  four 
salts  distributed  in  the  same  proportions.  Let  m  and  n  be  positive 
integers,  then 

(m  +  n)NaN03  +  (m  +  w)KCl  =  mNaCl  +  mKN03  +  wNaN03  +  nKCl ; 
(m  +  w)NaCl  +  (m  +  rc)KN03  -  wNaCl  +  raKN03  +  wNaN03  +  wKCl. 
This  is  more  concisely  written, 

NaCl  +  KN03^NaN03  +  KOI. 

The  phenomenon  is  explained  by  assuming  that  the  products  of 
the  reaction   interact  to  reform  the  original   components  simul- 

P 


226  HIGHER  MATHEMATICS.  §  78. 

taneously  with  the  direct  reaction.  That  is  to  say,  two  inde- 
pendent and  antagonistic  changes  take  place  simultaneously  in 
the  same  reacting  system.  When  the  speeds  of  the  two  opposing 
reactions  are  perfectly  balanced,  the  system  appears  to  be  in  a 
stationary  state  of  equilibrium.  This  is  another  illustration  of  the 
principle  of  the  coexistence  of  different  actions.  The  special  case 
of  the  law  of  mass  action  dealing  with  these  "incomplete"  or 
reversible  reactions  is  known  as  Guldberg  and  Waage's  law. 
Consider  a  system  containing  two  reacting  substances  A1  and  A2 
such  that 

A1  ^s  A2. 

Let  ax  and  a2  be  the  respective  concentrations  of  Al  and  A2.  Let 
x  of  A1  be  transformed  in  the  time  t,  then  by  the  law  of  mass  action, 

^  =  kY(aY  -  x). 

Further,  let  x'  of  A2  be  transformed  in  the  time  L  The  rate  of 
transformation  of  A2  to  Ax  is  then 

—  =  k2(a2-  x). 

But  for  the  mutual  transformation  of  x  of  Al  to  A2  and  x'  of  A2 
to  Alt  we  must  have,  for  equilibrium,  x  =  -  x' ;  and,  dx  =  -  dx' ; 

.-.  ^=  -  k2(a2  +  x). 

The  net,  or  total  velocity  of  the  reaction  is  obviously  the  algebraic 
sum  of  these  "  partial "  velocities,  or 

^  -=  KMi  -  x)  -  k2(a2  +  x).        .         .        (1) 

It  is  usual  to  write  K  =  kjk2.  When  the  system  has  attained  the 
stationary  state  dx/dt  -  0.  "Equilibrium,"  says  Ostwald,  "is  a 
state  which  is  not  dependent  upon  time."     Consequently 

Z=K±4     ....        (2) 

where  x  is  to  be  determined  by  chemical  analysis,  aY  is  the  amount 
of  substance  used  at  the  beginning  of  the  experiment,  a2  is  made 
zero  when  t  =  0.  This  determines  K.  Now  integrate  (1)  by 
the  method  of  partial  fractions  and  proceed  as  indicated  in  the 
subjoined  examples. 


§78.  THE  INTEGRAL  CALCULUS.  227 

Examples. — (1)  In  aqueous  solution  -y-oxybutyric  acid  is  converted  into 
y-butyrolactone,  and  y-butyrolactone  is  transformed  into  y-oxybutyric  acid 
according  to  the  equation, 

CRjOH  .  CHo .  0H2 .  COOH  =b  CH2  .  0H2 .  0H2 .  CO  +  H20. 

I 0. 1 

Use  the  preceding  notation  and  show  that  the  velocity  of  formation  of  the 
lactone  is,  dxfdt  =  hy(ax  -  x)  -  k2(a2  +  x),  and  K  m  kjk2  =  (a^  +  x)j{ax  -  x)» 
Now  integrate  the  first  equation  by  the  method  of  partial  fractions.  Evaluate 
the  integration  constant  for  x  =  0  when  t  =  0  and  show  that 

7  •  log  r~ r— -7T wr-  =  Constant.    ...         (3) 

t       8  (Zoj  -  a^  -  (1  +  K)x 

P.  Henry  (Zeit.  phys.  Ghem.,  10, 116,  1892)  worked  with  ax  =  18-23,  a2  =  0 ; 
analysis  showed  that  when  dx/dt  =  0,  x  =  13-28;  ai  -  x  =  4-95;  a.2  +  x  =  13-28 ; 
7l  =  2-68.  Substitute  these  values  in  (3);  reduce  the  equation  to  its  lowest 
terms  and  verify  the  constancy  of  the  resulting  expression  when  the  following 
pairs  of  experimental  values  are  substituted  for  x  and  t, 

t=    21,  50,  65,  80,  160   ...; 

x  =     2-39,        4-98,        6-07,        7-14,        10-28  . . . 
(2)  A  more  complicated  example  than  the  preceding  reaction  of  the  first 
order,  occurs  during  the  esterification  of  alcohol  by  acetic  acid : 

CH3 .  COOH  +  C2H5 .  OH  =h  CH3 .  COOC2H5  +  H  .  OH, 

a  reaction  of  the  second  order.  Let  Oj,  bx  denote  the  initial  concentrations  of 
the  acetic  acid  and  alcohol  respectively,  a^  b2  of  ethyl  acetate  and  water. 
Show  that,  dx/dt  =  ft^Oj  -  x)  (bx  -  x)  -  k^a^  +  a)  (b2  +  x).  Here,  as  else- 
where, the  calculation  is  greatly  simplified  by  taking  gram  molecules  such 
that  ax  =  1,  bx  =  1, 0-2  =  0,  b2  =  0.     The  preceding  equation  thus  reduces  to 

^  =  Ml  -  *)"  -  *¥* (4) 

For  the  sake  of  brevity,  write  kx/(k  -  &2)  =  m  and  let  o,  £  be  the  roots  of  the 
equation  <c2  -  2mx  +  m  =  0.     Show  that  (7)  may  be  written 

dx 

(E  -  .)  (S  -  fl)  =  ^  -  ^ 

Integrate  for  x  =  0  when  t  =  0,  in  the  usual  way.  Show  that  since 
a  =  m  +  \/m2  -  m  and  j8  =  m  -  s/m2  -  w,  page  353, 

1        (m  -  *Jm*  -  m)  (m  +  slm?  -  m  -  x)      rt/7        j       ,— 

7  lo8  ; /    ,  /    o         "  =  ^  -  fc2  Vm2  -  m.         5 

t         (m  +  v w2  -  m)  (m  -  Vro2  -  m  -  x) 

K  is  evaluated  as  before.     Since  m  =  ^{k^  -  k2);m  =  1/(1  -  kjkx).     M.  Ber- 

thelot  and  L.  Pean  St.  Gilles'  experiments  show  that  for  the  above  reaction, 

fc^  =  4 ;  m  =  * ;  slm1  -  m  =  $ ;  %(kx  -  k%)  =  0*00575 ;  or,  using  common 

logs,  %(kx  x  0*4343  -  k2)  =  0-0025.     The  corresponding  values  of  x  and  t  were 

t=      64,  103,  137,  167     ...; 

x  =  0-250,  0-345,  0-421,  0*474    ...; 

constant  =  0-0023,         0-0022,         0-0020,         0.0021  . . , 

V* 


h 


228  HIGHER  MATHEMATICS.  §  78. 

Verify  this  last  line  from  (5).  For  smaller  values  of  t,  side  reactions  are 
supposed  to  disturb  the  normal  reaction,  because  the  value  of  the  constant 
deviates  somewhat  from  the  regularity  just  found. 

(3)  Let  one  gram  molecule  of  hydrogen  iodide  in  a  v  litre  vessel  be  heated, 
decomposition  takes  place  according  to  the  equation :  2HI  =^t  Hj+Ig.  Hence 
show  that  for  equilibrium, 

dx      7    /l  -  2xV    (    x\* 

and  that  (1  -  2x)/v  is  the  concentration  of  the  undissociated  acid.  Put 
kjk2  =  K  and  verify  the  following  deductions, 

dx  _     1       ,     VE^l  -  2a?)  +  x  _  h%t 

K(l-2x)*-x*-2>jK'     *>JK(1-2x)-x~  v* 

Since,  when  t  =  0,  x  =  0,  G  =  0.    M.  Bodenstein  (Zeit.  phys.  Ghem.,  13,  56, 

1894 ;  22, 1,  1897)  found  E,  at  440°  =  0*02,  hence  »JK  =  0-141, 

1         1  +  5-Ijc 
•  '•  7  •  l08 1  _  Q.ifl.  =  constant, 

provided  the  volume  remains  constant.  The  corresponding  values  of  x  and  t  are 
to  be  found  by  experiment.  E.g.,  when  t  =  15,  x  =  00378,  constant  =  0-0171 ; 
and  when  t  =  60,  x  =  0-0950,  constant  =  0*0173,  etc. 

(4)  The  "  active  mass  "  of  a  solid  is  independent  of  its  quantity.  Hence, 
if  c  is  a  constant,  show  that  for  OaC03  -^  OaO  +  0O2,  Kc  =  p,  where p  denotes 
the  pressure  of  the  gas. 

(5)  Prove  that  the  velocity  equation  of  a  complete  reaction  of  the  first 
order,  A2  =  A2,  has  the  same  general  form  as  that  of  a  reversible  reaction, 
Aa  -^  A2,  of  the  same  order  when  the  concentration  of  the  substances  is  re- 
ferred to  the  point  of  equilibrium  instead  of  to  the  original  mass.  Let  | 
denote  the  value  of  x  at  the  point  of  equilibrium,  then,  dxldt=k1(a1  -x)-  k^c ; 
becomes  dx/dt  =  fc^  -  {)  -  &2£.  Substitute  for  fe2its  value  /^(Oj  -  {)/£,  when 
dx/dt  =  0, 

.  dx_k1al(t-x)m        dx 
..-gg.-i— ,  or,  -^  =  *(*-•).    .         .         .         (7) 

where  the  meanings  of  a,  k,  kx  will  be  obvious. 

(6)  Show  that  k  is  the  same  whether  the  experiment  is  made  with  the 
substance  Alf  or  Ag.  It  has  just  been  shown  that  starting  with  Alt  k  =  fe^/l ; 
starting  with  A2,  it  is  evident  that  there  is  %  -  £  of  A2  will  exist  at  the  point 
of  equilibrium.  Hence  show  dx/dt  =  k^i^fa  -  £)  -  x}/(a1  - 1) ;  k£=  k^  -  £), 
therefore,  as  before,  k^/fa  -  |)  =  k^aj^.  dx/dt  =  M^i^  -  £  -  «)/{.  Inte- 
grate between  the  limits  t  =  0  and  t  =  t,  x  =  x0  and  x  =  xx\  then  show,  from 
(7),  that 

<1osr    «logfli - 1 - *2~T~  =  kl  +  K  •     '     (8) 

C.  Tubandt  has  measured  the  rate  of  inversion  of  Z-menthone  into  d-men- 
thone,  and  vice  versd  (Dissertation,  Halle,  1904).  In  the  first  series  of  experi- 
ments x  denotes  the  amount  of  d-menthone  present  at  the  time  t ;  and  in  the 


§  79.  THE  INTEGRAL  CALCULUS.  229 

second  series,  the  amount  of  Z-menthone  present  at  the  time  t ;  |  is  the  value 
,  of  x  when  the  system  is  in  a  state  of  equilibrium,  that  is  when  t  is  infinite. 
First,  the  conversion  of  Z-menthone  into  d-menthone. 

t  =  0,       15,       30,       45,       60,       75,       90,       105,  oo ; 

x  =  0,    0-73,     1-31,     1-74,     2-06,     230,     2-48,      2-62,  3*09. 
Second,  the  conversion  of  d-menthone  into  Z-menthone. 

t  =  0,       15,       30,       45,       60,       75,       90,       105,  oo; 

x  =  0,    045,    0-76,     1-03,     1-22,     1-37,     1'47,     1'56,  1-84. 
Show  that  the  "  velocity  constant"  is  nearly  the  same  in  each  case,    k =0*008 
nearly. 

§  79.  Fractional  Precipitation. 

If  to  a  solution  of  a  mixture  of  two  salts,  A  and  B,  a  third 
substance  C,  is  added,  in  an  amount  insufficient  to  precipitate  all 
A  and  B  in  the  solution,  more  of  one  salt  will  be  precipitated,  as 
a  rule,  than  the '  other.  By  redissolving  the  mixed  precipitate  and 
again  partially  precipitating  the  salts,  we  can,  by  many  repetitions 
»f  the  process,  effect  fairly  good  separations  of  substances  otherwise 
intractable  to  any  known  process  of  separation. 

Since  Mosander  thus  fractioned  the  gadolinite  earths  in  1839, 
the  method  has  been  extensively  employed  by  W.  Crookes  (Chem. 
News,  54,  131,  155,  1886),  in  some  fine  work  on  the  yttria  and 
other  earths.  The  recent  separations  of  polonium,  radium  and 
other  curiosities  have  attracted  some  attention  to  the  process. 
The  "  mathematics  "  of  the  reactions  follows  directly  from  the  law 
of  mass  action.  Let  only  sufficient  C  be  added  to  partially  pre- 
cipitate A  and  B  and  let  the  solution  originally  contain  a  of  the 
salt  A,  b  of  the  salt  B.  Let  x  and  y  denote  the  amounts  of  A  and 
B  precipitated  at  the  end  of  a  certain  time  t,  then  a  -  x  and  b  -  x 
will  represent  the  amounts  of  A  and  B  respectively  remaining  in 
the  solution.     The  rates  of  precipitation  are,  therefore, 

-£  =  hx(a  -  x)  (g  -  z) ;  ^  -  k2(b  -  y)  (c  -  z), 

where  c  -  z  denotes  the  amount  of  C  remaining  in  the  solution  at 
the  end  of  a  certain  time  t. 

or,  multiplying  through  with  dt,  we  get 

k  J?-  =  ic  _^_ .  .  k  [d(a  -  x)  Z  k  [d(b  -  y) 

2a  -  x        1b  -  y'         2)   a  -  x  l)   b  -  y 


230 


HIGHER  MATHEMATICS. 


§80. 


On  integration,  k2\og(a  -  x)  =  k^.og{b  -  y)  +  log  0,  where  log  C 
is  the  integration  constant.  To  find  G  notice  that  when  x  =  0, 
y  =  0,  and  consequently  log  a**  =  log  Cbki ;  or,  0  =  a*2/£*i.- 


•*, 


(1) 


log 


b  —  y 

The  ratio  {a  -  x)/a  measures  the  amount  of  salt  remaining  in 
the  solution,  after  x  of  it  has  been  precipitated.  The  less  this  ratio, 
the  greater  the  amount  of  salt  A  in  the  precipitate.  The  same 
thing  may  be  said  of  the  ratio  (b  -  y)/b  in  connection  with  the 
salt  B.  The  more  k2  exceeds  klt  the  less  will  A  tend  to  accumulate 
in  the  precipitate  and,  the  more  kx  exceeds  k2,  the  more  will  A  tend 
to  accumulate  in  the  precipitate.  If  the  ratio  kjk2  is  nearly  unity, 
the  process  of  fractional  precipitation  will  be  a  very  long  one, 
because  the  ratio  of  the  quantities  of  A  and  B  in  the  precipitate 
will  be  nearly  the  same.  In  the  limiting  case,  when  k1  =  k2,  or 
kjk2  =  1,  the  ratio  of  A  to  B  in  the  mixed  precipitate"  will  be  the 
same  as  in  the  solution.  In  such  a  case,  the  complex  nature  of 
the  "earth"  could  never  be  detected  by  fractional  precipitation. 
The  application  to  gravimetric  analysis  has  not  yet  been  worked 
out. 

§  80.    Areas  enclosed  by  Curves.    To  Evaluate  Definite 

Integrals. 

Let  AB  (Fig.  100)  be  any  curve  whose  equation  is  known.     It 

is  required  to  find  the  area  of  the 
portion  bounded  by  the  curve ;  the 
two  coordinates  PM,  QN ;  and  that 
portion  of  the  #-axis,  MN,  included 
between  the  ordinates  at  the  ex- 
tremities of  that  portion  of  the  curve 
under  investigation.  The  area  can 
be  approximately  determined  by  sup- 
posing PQMN  to  be  cut  up  into  small 
strips — called  surface  elements — 
«"  perpendicular  to  the  #-axis  ;  finding 
the  area  of  each  separate  strip  on 
the  assumption  that  the  curve  bound- 
ing one  end  of  the  strip  is  a  straight 


§  80.  THE  INTEGRAL  CALCULUS.  231 

line ;  and  adding  the  areas  of  all  the  trapezoidal- shaped  strips 
together.  Let  the  surface  PrqQNM  be  cut  up  into  two  strips  by 
means  of  the  line  LB.     Join  PB,  BQ. 

Area  PQMN  =  Area  PBLM  +  Area  BQNL. 

But  the  area  so  calculated  is  greater  than  that  of  the  required 
figure.  The  shaded  portion  of  the  diagram  represents  the  magni- 
tude of  the  error.  It  is  obvious  that  the  narrower  each  strip  is 
made,  the  greater  will  be  the  number  of  trapeziums  to  be  included 
in  the  calculation  and  the  smaller  will 
be  the  error.  If  we  could  add  up  the 
areas  of  an  infinite  number  of  such  strips, 
the  actual  error  would  become  vanish- 
ingly  small.  Although  we  are  unable 
to  form  any  distinct  conception  of  this 
process,  we  feel  assured  that  such  an 
operation  would  give  a  result  absolutely 
correct.  But  enough  has  been  said  on 
this  matter  in  §  68.  We  want  to  know 
how  to  add  up  an  infinite  number  of  infinitely  small  strips. 

In  order  to  have  some  concrete  image  before  the  mind,  let 
us  find  the  area  of  PQNM  in  Fig.  101.  Take  any  small  strip 
PBSM ;  let  PM  =  y,  BS  =  y  +  8y ;  OM  =  x  ;  and  OS  =  x  4-  8x. 
Let  8A  represent  the  area  of  the  small  strip  under  consideration. 
If  the  short  distance,  PB,  were  straight  and  not  curved,  the  area, 
8A}  of  the  trapezium  PBSM  would  be,  (U),  page  604, 

8A  =  \8x(PM  +  BS)  =  8x(y  +  $y). 

By  making  8x  smaller  and  smaller,  the  ratio,  8A/8x  -  y  +  |&/, 
approaches,  and,  at  the  limit,  becomes  equal  to 

***-«.■"  «E^*S  •••<M-if.A».:      .      (i) 

This  formula  represents  the  area  of  an  infinitely  narrow  strip,  say, 
PM.  The  total  area,  A,  can  be  determined  by  adding  up  the. 
areas  of  the  infinite  number  of  infinitely  narrow  strips  ranged  side 
by  side  from  MP  to  NQ.  The  area  of  any  strip  is  obviously  length 
x  width,  or  y  x  dx.  Before  we  can  proceed  any  further,  we  must 
know  the  relation  between  the  length,  y,  of  any  strip  in  terms  of  its 
distance,  x,  from  the  point  0.  In  other  words,  we  must  have  the 
equation  of  the  curve  PQ.     For  instance,  the  area  of  any  indefinite 


232  HIGHER  MATHEMATICS.  §  80. 

portion  of  the  curve,  is 

A  =  \y  .  dx  +  0,  .  .  ,  .  (2) 
and  the  area  bounded  by  the  portion  situated  between  the  ordinates 
having  the  absciss©  a2  and  az  (Fig.  100)  is 

A=[a*y.dx+  C.  .         .         .         (3) 

Equation  (2)  is  an  indefinite  integral,  equation  (3)  a  definite 
integral.  The  value  of  the  definite  integral  is  determined  by  the 
magnitude  of  the  upper  and  lower  limits.  In  Fig.  100,  if  av  a2,  a3 
represent  the  magnitudes  of  three  absciss sb,  such  that  a2  lies 
between  a2  and  az, 

A  =  I   y .  dx  +  G  =  I   y .  dx  +     y .  dx  +  C. 

J «]  J  «1  J  <*2 

When  the  limits  are  known,  the  value  of  the  integral  is  found  by 
subtracting  the  expression  obtained  by  substituting  the  lower  limit 
in  place  of  x,  from  a  similar  expression  obtained  by  substituting 
the  upper  limit  for  x. 

In  order  to  fix  the  idea,  let  us  take  a  particular  case.  Suppose 
y  =  2x,  and  we  want  to  deal  with  that  portion  of  the  curve  between 
the  ordinates  a  and  b.     From  (3), 

r2a.da?=rz2  +  G;        .         .         ,         (4) 

The  vertical  line  in  the  preceding  equation,  (4),  resembles  Sarraus' 
symbol  of  substitution.  The  same  idea  is  sometimes  expressed 
by  square  brackets,  thus, 

J  2x.dx=  He2  +  cj  =  (62  +  G)  -  (a2  +  G)  =  (¥  -  a?). 

The  process  of  finding  the  area  of  any  surface  is  called,  in  the 
regular  text-books,  the  quadrature  of  surfaces,  from  the  fact  that 
the  area  is  measured  in  terms  of  a  square — sq.  cm.,  sq.  in.,  or 
whatever  unit  is  employed.  In  applying  these  principles  to 
specific  examples,  the  student  should  draw  his  own  diagrams. 
If  the  area  bounded  by  a  portion  of  an  ellipse  or  of  an  hyperbola 
is  to  be  determined,  first  sketch  the  curve,  and  carefully  note  the 
limits  of  the  integral. 

Examples. — (1)  To  find  the  area  bounded  by  an  ellipse,  origin  at  the 
centre.    Here 

x2      y2  b    i 

a?+  b*^1''  or,  y  =±~^a2  -  xr\ 


§  80.  THE  INTEGRAL  CALCULUS.  233 

Refer  to  Fig.  22,  page  100.     The  sum  of  all  the  elements  perpendicular  to  the 
x-axis,  from  OPx  to  0PV  is  given  by  the  equation 


/> 


dx, 


for,  when  the  curve  cuts  the  aj-axis,  x  =  a,  and  when  it  cuts  the  y-axis,  x  =  0. 
The  positive  sign  in  the  above  equation,  represents  ordinates  above  the  jc-axis. 
The  area  of  the  ellipse  is,  therefore, 

fa 

dx. 


=  *fav 

Jo 


Substitute  the  above  value  of  y  in  this  expression  and  we  get  for  the  sum  of 
this  infinite  number  of  strips, 

6f« 


A 

o 


aJ  o 
which  may  be  integrated  by  parts,  thus 

b  Vx  a?  x 

i  =  4-     -2^-^)  +  2sin 


a 

Jo 


The  term  within  the  braokets  is  yet  to  be  evaluated  between  the  limits  x  =  a 
and  x  =  0 

b  T  (a  a2  a         }      (0  a2  0         ^  1 

A  Ab  ^    ■  H 

•••^  =  4aX  2-sm~1; 

remembering  that  sin 90°  =  1,  sin"1!  =  90°  and  2 sin"1 1  =  180  =  ir.  The 
area  of  the  ellipse  is,  therefore,  vdb.  If  the  major  and  minor  axes  are  equal, 
a  =  b,  the  ellipse  becomes  a  circle  whose  area  is  7ra2.  It  will  be  found  that 
the  constant  always  disappears  in  this  way  when  evaluating  a  definite 
integral. 

(2)  Find  the  area  bounded  by  the  rectangular  hyperbola,  xy  =  a;   or, 
y  =  ajx,  between  the  limits  x  —  xx  and  x  *=  x2. 


f*2  t^a 

y.dx=       -dx\ 

Jx\  Jxi 


.'.  A  =a\    log  x  +  C  =  a{(log  x2  +  C)  -  ( log  ^  +  C)\  =  a  logl*. 

I*!  Xl 

If  xx  =  1,  and  sc2  =  x  ;  A  =  a  log^.     This  simple  relation  appears  to  be  the 
reason  natural  logarithms  are  sometimes  called  hyperbolic  logarithms. 

(3)  Find  the  area  bounded  by  the  curve  y  =  12(sc  -  l)/x,  when  the  limits 
are  12  cm.  and  3  cm.  Ansr.  91-36  sq.  cm.  The  integral  is  12J(a;  -l)x~1dx ; 
or  12[a5  -  log  a?]-1/  =  12(9  -  log  4),  etc.  Use  the  table  of  natural  logarithms, 
page  627. 

(4)  Show  that  the  area  bounded  by  the  logarithmic  curve,  x  =  logy,  is 
y  -  l.  Hint.  A  =  jdy  =  y  +  C.  Evaluate  G  by  noting  that  when  x  =  0, 
y  =  -  1,  A  =  0.     If  y  =  1,  A  =  0;  if  y  =  2,  A  =  1 ;  etc. 

If  polar  coordinates  are  employed,  the  differential  of  the  area 


234  HIGHER  MATHEMATICS.  §  81. 

assumes  the  form 

dA  m  \rU0 (5) 

Example. — Find  the  area  of  the  hyperbolic  spiral  between  0  and  +  r. 
See  Ex.  (2),  page  117.    r0  =  a;  de  =  -  a  .  drjr2;  consequently, 

.  f0    a     ,  1\°    ar     ar 

A'zj  *•*■■- 4  T=T 

After  this  the  integration  constant  is  not  to  be  used  at  any 
stage  of  the  process  of  integration  between  limits.  It  has  been 
retained  in  the  above  discussion  to  further  emphasize  the  rule : 
The  integration  constant  of  a  definite  integral  disappears  during 
the  process  of  integration.  The  absence  of  the  indefinite  integration 
constant  is  the  mark  of  a  definite  integral. 

§  81.    Mean  Values  of  Integrals. 

The  curve 

y  =  rsin#, 
represents  the  sinusoid  curve  for  the  electromotive  force,  y,  of  an 
alternating  current ;  r  denotes  the  maximum  current ;  x  denotes 
the  angular  displacement  made  in  the  time  t,  such  that  #  =  2tt£/T, 
where  T  denotes  the  time  of  a  complete  revolution  of  the  coil  in 
seconds.  The  value  of  y,  at  any  instant  of  time,  is  proportional  to 
the  corresponding  ordinate  of  the  curve.  When  x  =  90°,  t  =  \T,  the 
coil  has  made  a  quarter  revolution,  and  the  ordinate  is  a  maximum. 

When  x  =  270°,  or  when  t  =  f  T%  that 
is,  in  three-quarters  of  a  revolution, 
the  ordinate  is  a  minimum.  The 
curve  cuts  the  aj-axis  when  x  =  0°, 
180°,  and  360°,  that  is,  when  t  =  0, 
\T,  and  T.  For  half  revolution,  'the 
average  electromotive  force  beginning 
Fia  102#  when   x  =  0,   is   equal    to   the   area 

bounded  by  the  curve  OPC  (Fig.  102), 
and  the  #-axis,  cut  off  at  }£T,  that  is,  at  *-.  This  area,  Av  is 
evidently 

A1  =  I   r  sin  x .  dx  =  -    r  cos  x\    =  -  r  cos  ?r  +  r  cos  0  =  2r, 

because  costt  =  cos  180°  =  -  1,  and  cos  0°  =  1. 

The  area,  A2,  bounded  by  the  sine  curve  during  the  second 
half  revolution  of  the  coil  lies  below  the  rc-axis,  and  it  has  the  same 
numerical  value  as  in  the  first  half.     This  means  that  the  average 


§  81.  THE  INTEGRAL  CALCULUS.  235 

electromotive  force  during  the  second  half  revolu  fcion  is  numerically 
equal  to  that  of  the  first  half,  but  of  opposite  sign.  It  is  easy  to 
see  this. 


2r, 


A2  =  I    r&mx.dx=  -  \roosx\    =  -  tcosQtt  +  rcos-n-  = 

since  cos  360°  =  cos0°  =  1.  The  total  area,  A,  bounded  by  the 
sine  curve  and  the  #-axis  during  a  complete  revolution  of  the  coil 
is  zero,  since 

A  =  A1  +  A2  =  0. 

The  area  bounded  by  the  sine  curve  and  the  #-axis  for  a  whole 
period  2tt,  or  for  any  number  of  whole  periods,  is  zero. 

If  now  ylt  y2,  ys, . . .,  yn  be  the  values  of  f(x)  when  the  space 
from  a  to  b  is  divided  into  n  equal  parts  each  hx  wide,  b-  a  =  n$x, 
and  if 

y  =  /(«) ; 


Vi  -/(<*)  ;  y2  =/(«  +  Sx) ;  y3  =/(a  +  28b); . . . ;  yn  =f(b  -  n  -  Ux). 
The  arithmetical  mean  of  these  n  values  of  y  is,  by  definition, 
the  nth  part  of  their  sum.     Hence, 

Vi  +  V2  +  Vz  +  -  -  +  Vn  m  (Vi  +  V2  +  Vz  +  >  •  •  +  Vn)^ 
n  b  -  a  ' 

since  nSx  =  b  -  a.  If  x  now  assumes  every  possible  value  lying 
in  the  interval  between  b  and  a,  n  must  be  infinitely  great.  Hence 
the  sum  of  this  infinite  number  of  indefinitely  small  quantities  is 
expressed  by  the  symbol 

as  indicated  on  page  189.  The  arithmetical  mean  of  all  the  values 
which  f(x)  can  assume  in  the  interval  b  -  a  is,  therefore, 


f(x)dx  1      f, 

s^t ;  or,  -r f(x)dx. 

b  -  a  b  -  a)aJK  J 


This  is  called  the  mean  or  average  value  oif(x)  over  the  range 
b  -  a.  Geometrically,  the  mean  value  is  the  altitude  of  a  rectangle, 
on  the  base  b  -  a,  whose  area  is  equal  to  that  bounded  by  the  curve 
y  =  f(x),  the  two  ordinates  and  the  #-axis.  In  Fig.  102,  OA  is  the 
mean  value  of  y,  that  is,  of  r  sin  x}  for  all  values  of  x  which  may 
vary  continuously  from  0  to  w.     This  is  easy  to  see, 


236  HIGHEK  MATHEMATICS.  §  81. 

I   rainx  .dx  =  Area  OPC  =  Area  of  rectangle  OABG\ 
=  OC  x  OA  =  (b  -  a)  x  OA, 
where  a  denotes  the  abscissa  at  the  point  0,  and  b  the  abscissa  at 
C.     But  b  -  a  =  7T,  and  OA  =  ylt  consequently, 


-f" 


2r 
Mean  value  of  ordinate  =  —  |   raillX.dx=  —  =  0*6366r. 

7T 


Instruments  for  measuring  the  average  strength,  yv  of  an  alternat- 
ing current  during  half  a  complete  period,  that  is  to  say,  during  the 
time  the  current  flows  in  one  direction,  are  called  electrodynamo- 
meters.  The  electrodynamometer,  therefore,  measures  y1  =  OA 
(Fig.  102)  =  2r/7r  =  0"6366r.  But  MP  =  r  denotes  the  maximum 
current,  because  sin  a;  is  greatest  when  x  =  90°,  and  sin  90°  =  1. 
Hence,  y  =  r  sin  90°  =  r. 

Maximum  current  =  T  ',  Average  current  =  0#6366t*. 
There  is  another  variety  of  mean  of  no  little  importance  in  the 
treatment  of  alternating  currents,  namely,  the  square  root  of  the 
mean  of  the  squares  of  the  ordinates  for  the  range  from  x  =  0  to 
x  =  7t.  This  magnitude  is  called  the  mean  square  Yalue  off(x). 
With  the  preceding  function,  y  =  r  sin  x,  the 

r2 


--f 


Mean  value  of  V2  =  -  \   T  sin2aj .  dx  =  -g 


on  integration  by  parts  as  in  (12),  page  205.     Again 
Mean  square  value  of  2/2  =  J^T2  =  0*707lr. 
Examples. — (1)  In  calculations  involving    mean  values  care  must  be 
taken  not  to  take  the  wrong  independent  variable.     Find  the  mean  velocity  of 
a  particle  falling  from  rest  with  a  constant  acceleration,  the  velocities  being 
taken  at  equal  distances  of  time.     When  a  body  falls  from  rest,  V  =  gt, 


r/,«4Ji 


gt .  dU 


gt_V 


o  2  ~2' 

that  is  to  say,  the  mean  velocity,  Vt,  with  respect  to  equal  intervals  of  time 
is  one  half  the  final  velocity.  On  the  other  hand,  if  we  seek  the  mean 
velocity  which  the  body  had  after  describing  equal  intervals  of  space,  s, 
and  remembering  that  T2  =  2gs, 

that  is,  two-thirds  of  the  final  velocity. 

(2)  Show  that  if  a  particle  moves  with  a  constant  acceleration,  the  mean 
square  of  the  velocities  at  equal  infinitely  small  intervals  of  time,  is 
M^o2  +  "Po "Pi  +  ^i2)»  wnere  Vo  an<*  ^1  respectively  denote  the  initial  and  final 
velocities. 


§82. 


THE  INTEGRAL  CALCULUS. 


237 


(3)  The  relation  between  the  amount,  x,  of  a  substance  transformed  at 
the  time,  tt  in  a  unimolecular  chemical  reaction  may  be  written  x=a(l  -  e~ **) 
where  a  denotes  the  amount  of  substance  present  at  the  beginning  of  the 
reaction,  and  Ms  a  constant.  Show  that  V  =  ake  ~  ** ;  or,  V  =  k(a  -  x) 
according  as  we  refer  the  velocity  to  equal  intervals  of  time,  t ;  or  to  equal 
amounts  of  substance  transformed,  x.  Also  show  that  the  mean  velocity 
with  respect  to  equal  intervals  of  time  in  the  interval  tx  -  t0,  is 


t\  ~  ^o  J  to 


ake  ~  udt  — 


ale-**!  -  g~fao) 


"o  J  to  h~  <o  log  V  0  -  log  7i 

and  the  mean  velocity,  Vx,  with  respect  to  equal  amounts  of  substance  trans- 
formed, is 

V  1        [\,„     „w„  _  fe(ah  -  «b)  (2a  -  flfc  -  x0)  _  V0  +  V1 

v* = *r^o)XQk{a  - x)dx  ~ ^r^5 2— 

If  t0  =  0,  and  i,  is  infinite,  the  mean  velocity,  Vt,  converges  towards  zero. 
Several  interesting  relations  can  be  deduced  from  this  equation. 

Problems  connected  with  mean  densities,  centres  of  mass, 
moments  of  inertia,  mean  pressures,  and  centres  of  pressure  are 
treated  by  the  aid  of  the  above  principles. 

§  82.    Areas  Bounded  by  Curves.    Work  Diagrams. 

I. — The  area  enclosed  between  two  different  curves.  Let  PABQ 
and  PA'B'Q  (Fig.  103)  be  two  curves,  it  is  required  to  find  the 
area  PABQB'A'.  Let  yl  =  fx(x)  be  the  equation  of  one  curve, 
Vi  =  AM*  tne  equation  of  the  other.  First  find  the  abscissae  of 
the  points  of  intersection  of  the  two  curves.     Find  separately  the 

y 


103.  Fig.  104. 

areas  PABQMN  and  PA'B'QMN,  by  the  preceding  methods.  Let 
a  and  b  respectively  denote  the  abscissae  OM,  and  ON  (Fig.  103)f 
of  the  points  of  intersection,  P  and  Q,  of  the  two  curves, 
required  area,  A,  is,  therefore, 

Area  PABQB'A'  =  Area  PABQMN  -  Area  PA'B'QMN 


The 


4=1  yxdx  -  J  y2 .  dx. 


(1) 


238 


HIGHER  MATHEMATICS. 


§82. 


To  find  the  area  of  the  portion  ABB' A',  let  xx  be  the  abscissa 
of  AB  and  x0  the  abscissa  of  BS,  then, 


f*2  f  x2  f*2 

=       Vi-dx  -  \    y2.dx  =  \    (yx~  y2)dx. 
J*i  Jxi  jx\ 


(2) 


In  illustration  let  us  consider  the  area  included  between  the  two 
parabolas  whose  equations  are  ?/2  =  4a> ;  and  x2  =  4y.  The  curves 
obviously  meet  at  the  origin,  and  at  the  point  x  =  4,  cm.,  say, 
y  =  4  cm.  (16),  page  95.     Consequently, 


=  I  -j-dx  -  I  2  \/# .  dx  =  2 1  ( -o  -   \/#  Jda; 


2f  sq.  cm.      (3) 


Why  the  negative  sign  ?  On  plotting  it  will  be  seen  that  we  first 
integrated  along  the  line  OGB  (Fig.  104),  and  then  subtracted  from 
this  the  result  of  integrating  along  the  line  OAP.  We  ought  to 
have  gone  along  OAP  first.  It  is  therefore  necessary  to  pay  some 
attention  to  this  matter. 

Let  a  given  volume,  x,  of  a  gas  be  contained  in  a  cylindrical 
vessel  in  which  a  tightly  fitting  piston  can  be  made  to  slide  (Fig. 

105).  Let  the  sectional  area  of 
the  piston  be  unity.  Now  let 
the  volume  of  .gas  change  dx 
units  when  a  slight  pressure  X 
Fig.  105.  is  applied  to  the  free  end  of  the 

piston.      Then,  by  definition  of  work,  W, 

Work  =  Force  x  Displacement ;  or,  dW  =  X .dx. 
If  p  denotes  the  pressure  of  the  gas  and  v  the  volume,  we  have, 

dW  =  p  .dv. 
Now  let  the  gas  pass  from  one  condition  where  x  =  Xj  to  an- 
other state  where  x  =  xT     Let  the  corresponding  pressures  to 
which  the  gas  was  subjected  be  respectively  denoted  by  X1  and  X2. 

By  plotting  the  successive  values  of  X 
and  x,  as  x  passes  from  xx  and  x2,  we 
get  the  curve  ACB,  shown  in  Fig.  106. 
The  shaded  part  of  the  figure  represents 
the  total  work  done  on  the  system 
during  the  change. 

If  the  gas  returns  to  its  original 
state  through  another  series  of  succes- 
sive values  of  X  and  x  we  have  the 
curve  AVB    (Fig.    107).      The   total 


Fia.  106.— Work  Diagram. 


§82. 


THE  INTEGRAL  CALCULUS. 


239 


107.— Work  Diagram. 


work  done  by  the  system  will  then  be  represented  by  the  area 
ABDx2xv  If  we  agree  to  call  the  work  done  on  the  system 
positive ;  and  work  done  by  the  system  negative,  then  (Fig.  107), 

Wx  -  W2  =  Area  ACBx^Xx  -  Area  ADBx^X^  =  Area  ACBD. 
The  shaded  part  in  Fig.  107,  therefore,  represents  the  work  done 
on  the  system  during  the  above  cycle 
of  changes.  A  series  of  operations  by 
which  a  substance,  after  leaving  a  certain 
state,  finally  returns  to  its  original  con- 
dition, is  called  a  cycle,  or  a  cyclic 
process.  A  cyclic  process  is  represented 
graphically  by  a  closed  curve.  In  any 
cyclic  change,  the  work  done  on  the 
system  is  equal  to  the  "area  of  the 
cycle  ". 

Work  is  done  on  the  system  while  x  is  increasing  and  by  the 
system  when  x  is  decreasing.  Therefore,  if  the  curve  is  described 
by  a  point  moving  round  the  area  ACBD  in  the  direction  of  the 
hands  of  a  clock,  the  total  work  done 
on  the  system  is  positive ;  if  done  in 
the  opposite  direction,  negative.  We 
can  now  understand  the  negative 
sign  in  the  comparatively  simple  ex- 
ample, Fig.  104,  above.  We  should 
have  obtained  a  positive  value  if  we 
had  started  from  the  origin  and  taken 
the  curves  in  the  direction  of  the 
hands  of  a  clock. 

If  the  diagram  has  several  loops 
as  shown  in  Fig.  108,  the  total  work 
is  the  sum  of  the  areas  of  the  several 
loops  developed  by  the  point  travelling  in  the  same  direction  as  the 
hands  of  a  clock,  minus  the  sum  of  the  areas  developed  when  the 
point  travels  in  a  contrary  direction.  This  graphic  mode  of  re- 
presenting work  was  first  used  by  Clapeyron.  The  diagrams  are 
called  Clapeyron's  Work  Diagrams. 

In  Watt's  indioator  diagrams,  the  area  enclosed  by  the  curve  represents 
the  excess  of  the  work  done  by  the  steam  an  the  piston  during  a  forward 
stroke,  over  the  work  done  by  the  piston  when  ejecting  the  steam  in  the  re- 
turn stroke.    The  total  energy  communicated  to  the  piston  is  thus  represented 


Work  Diagram. 


240  HIGHER  MATHEMATICS.  §  83. 

by  the  area  enclosed  by  the  curve.     This  area  may  be  determined  by  one  of 
the  methods  described  in  the  next  chapter,  page  335,  §  110. 

II.  The  area  bounded  by  two  branches  of  the  same  curve  is  but 
a  simple  application  of  Equation  (1).  Thus  the  area,  A,  enclosed 
between  the  two  limbs  of  the  curve  y2  =  (x2  +  6)2  and  the  ordinates 
x  =  1,  x  —  2  is 

A  =±  \  (x2  +  6)dx  =  ±  8J  units, 

as  you  will  see  by  the  method  adopted  in  the  preceding  example. 

Example. — Show  that  the  area  between  the  parabola  y  =  x2  -  5x  +  6, 
the  ic-axis,  and  the  ordinates  x  =  1,  x  =  5,  is  5£  units.  Hint.  Plot  the  curve 
and  the  last  result  follows  from  the  diagram.  Of  course  you  can  get  the  same 
result  by  integrating  ydx  between  the  limits  x  =  5,  and  x  =  1. 

§  83.    Definite  Integrals  and  their  Properties. 

There  are  some  interesting  properties  of  definite  integrals  worth 
noting,  and  it  is  perhaps  necessary  to  further  amplify  the  remarks 
on  page  232. 

I.  A  definite  integral  is  a  function  of  its  limits.  If /'(a?)  denotes 
the  first  differential  coefficient  oif(x), 

J7(«) .  da>  =  [/(a)  ]\  „,  |V(*)  =  /(&)  -  f{a). 

This  means  that  a  definite  integral  is  a  function  of  its  limits,  not  of 
the  variable  of  integration,  or 

[f(x) .  dx  =  [f(y) .  dy  -  f/W  .&.-!;        .         (1) 

J"  Ja  Ja 

In  other  words,  functions   of   the   same  form,  when  integrated 
between  the  same  limits  have  the  same  value. 

Examples.— (1)  Show  /  e~xdx  =  /  e-*dz=e~a-e-*. 

(2)  Prove  j*_  x*.dx  =  i{(3)»  -  (  -  1)*}  =  9£. 
By  way  of  practice  verify  the  following  results : — 

(3)  J       sin  x  .  dx  =  -  I       cos  x  =  -  I  cos  ^  -  cos  0°  J  =  1. 

/£*•  ir       [%*  1/ir  \        />  t 

sin2* .  dx  =  | ;   /      sin2*  .  dx=^(  -^  -  1  ) ;   /     sin2*  .  dx  =  g. 


§83. 


THE  INTEGRAL  CALCULUS. 


241 


Hint  for  the  indefinite  integral.     Integrate  by  parts. 
dv  =  sin  x .  dx.     From  (1),  §  74, 


Put  u  =  sin  x, 


Jsin2^  .  dx  =  sin  x .  cos  x  +  Jcos2^  .  dx  =  sin  a? .  cos  x  +  j(l  -  sin2x)dx. 
Transpose  the  last  term  to  the  left-hand  side,  and  divide  by  2. 
,\  Jsin2^  .  dx  =  £(sin  x  .  cos  x  +  x)  +  C. 

II.  The  interchange  of  the  limits  of  a  definite  integral  causes 
the  integral  to  change  its  sign.     It  is  evident  that 


^"/(x)dx  -/(<*)  -  f(b)  =  -^f{x)dx, 


(2) 


or,  when  the  upper  and  lower  limits  of  an  integral  are  inter- 
changed, only  the  sign  of  the  definite  integral  changes.  This 
means  that  if  the  change  of  the  variable  from  b  to  a  is  reckoned 
positive,  the  change  from  a  to  b  is  negative.  That  is  to  say,  if 
motion  in  one  direction  is  reckoned  positive,  motion  in  the 
opposite  direction  is  to  be  reckoned  negative. 

III.  The  decomposition  of  the  integration  limits.  If  m  is  any 
interval  between  the  limits  a  and  b,  it  follows  directly  from  what 
has  been  said  upon  page  232,  that 

\Uf\x)dx  =  \a/{x)dx  +  fy'(x)dx  -/(a)  -f(m)  +f(m)  -f(b).    (3) 

Or  we  can  write 

\'f(x)dx  =  ^f(x)dx-  ^f'(x)dx=f(m).-f(b)  -f(m)+f(a).      (i) 

In  words,  a  definite  integral  extending  over  any  given  interval 
is  equal  to  the  sum  of  the  definite  integrals 
extending  over  the  partial  intervals.  Con- 
sequently, if  f\x)  is  a  finite  and  single- 
valued  function  between  x  =  a,  and  x  =  b, 
but  has  a  finite  discontinuity  at  some  point 
m  (Fig.  109),  we  can  evaluate  the  in- 
tegral by  taking  the  sum  of  the  partial 
integrals  extending  from  a  to  m,  and  from 
m  to  b. 


y 

R 

p 

0 

.  s 

X 

Fig.  109. 


When  any  function  has  two  or  more  values  for  any  assigned  real  or 
imaginary  value  of  the  independent  variable,  it  is  said  to  be  a  multi-valued 

Q 


242 


HIGHER  MATHEMATICS. 


§83. 


function.  Suoh  are  logarithmic,  irrational  algebraic,  and  inverse  trigonomet- 
rical functions.  For  example,  y  =  tan  ~  lx  is  a  multiple- 
valued  function,  because  the  ordinates  corresponding  to 
the  same  value  of  x  differ  by  multiples  of  -n.  Verify  this 
by  plotting.  Obviously,  if  x  =  a  and  x  =  b  are  the 
limits  of  integration  of  a  multiple-valued  function,  we 
must  make  sure  that  the  ordinates  x  =  a  and  x  =  6 
belong  to  the  same  branch  of  the  curve  y  =  f(x).  In 
Fig.  110,  if  x=OM,  y  is  multi-valued,  for y  may  be  MP, 
MQ,  or  MR.  The  imaginary  values  in  no  way  interfere 
with  the  ordinary  arithmetical  ones.  A  single-valued 
function  assumes  one  single  value  for  any  assigned  (real  or  imaginary)  value 
of  the  independent  variable.  For  example,  rational  algebraic,  exponential 
and  trigonometrical  functions  are  single-valued  functions. 

IV.  If  f(x)dx  be  one  function  of  y,  and  f'(a  -  x)dx  be  another 
function  of  y, 


f'(x)dx=\  f'(a  -  x)dx. 

Jo  Jo 


(5) 


For,  if  we  put  a  -  y  =  x ;  .*.  dx  =  -  dy,  and  substitute  x  =  a,  we 
see  at  once  that  y  =  0  ;  and  similarly,  if  x  =  0,  y  =  a. 

.-.  \af(x)dx=  -  [f(a  -  y)dy  =  [f'(a  -  x)dx, 

Jo  Jo  Jo 

from  (2)  and  (1)  abov<*  or  we  can  see  this  directly,  since 

I  f(x)dx  =  I  /'  {a  -  x)dx  -  -  [af\a  -  x)d(a  -  x)  =f(a)  -/(O). 

Jo  Jo  Jo 

This  result  simply  means  that  the  area  of  OPFO'  (Fig.  Ill) 
can  be  determined  either  by  taking  the  origin 
at  0  and  calling  00'  the  positive  direction  of 
the  #-axis  ;  or  by  transferring  the  origin  to 
the  point  0',  a  distance  a  from  the  old  origin 
0,  and  calling  O'O  the  positive  direction  of 
the  #-axis.  The  following  result  is  an  im- 
portant application  of  this, 


0f< " MO' 

Fig.  111. 


sinna; 
Jo 


dx 


x  )dx  =  1    cosw£C  .  dx. 


-J**8m-(|-*)fe-£"i 

Examples. — (1)  Verify  the  following  results : — 

cos  x  .  dx=  I     sin  x  .  dx—  1 ;  /     cos%c .  dx 
o  Jo  Jo 

(2)  Show  that    /"     f{x'i)dx=2iy(x2)dx. 


(6) 


J  0 


sin2£c .  da; =2' 


§  83.  THE  INTEGRAL  CALCULUS.  243 


(3)  Evaluate  /    sin  mx  .  sin  nxdx.     By  (28),  page  611, 

J  o 


2  sin  mx  .  sin  nx  =  cos(m  -  n)x  -  cos(m  +  ri)x. 
Jsin  mx  .  sin  nxdx  =  £Jcos(m  -  n)xdx  -  £jcos(m  +  n)xdx ; 

r  .  .  sin(m  -  n)x    sin(m  +  n)x 

sin  mx  .  sin  nxdx  —  -tt, f-  -  -777 ; — r- . 

2(m  -  n)  2(m  +  n) 

Therefore,  if  m  and  n  are  integral, 


/» 


/; 


sin  ma; .  cos  naafcc =0. 
1 


Remembering  that  sin  -k  =  sin  180°  =  0,  and  sin  0°  =  0,  if  m  =  n,  show  that 

Jl*ir^dxJ2f"(l-cos2nx)dx  =  [|-8i^Jo  -£. 

(4)  Show  that  the  integral  of  00s  mx  .  cos  no: .  dx,  between  the  limits  *■ 
and  0,  is  zero  when  m  and  n  are  whole  numbers  and  that  the  integral  is  £*•, 
when  m  =  n.     Hints.  From  (27),  page  612, 

2  00s  mx .  cos  nx=coB(m  -  n)x  +  cos(w  +  n)x. 


i 


/n  X  X 

a  sin  -x  .  cos  „  .  dx.     Ansr. 

2ajo  Mn5.^BinsJ--2-|o  sm^=fl. 

(6)  Show  that   /       cos  mx  .  cos  nx .  cte  =  0  ;  /        sin  ma; .  sin  nx  .dx=Q 


+  n 

cos  mx  .  sin  rac .  dx  =0.     Hint.  Use  the  results  of  Ex.  (3)  and  (4). 


Integrate/     -1=/     x~2da;=     --       ,  and  is  the  answer  -  2  ? 


V.  The  function  may  become  infinite  at  or  between  the  limits  of 
integration.  We  have  assumed  that  the  integrals  are  continuous 
between  the  limits  of  integration.  I  dare  say  that  the  beginner 
has  given  an  affirmative  answer  to  the  question  at  the  end  of  the 
last  example.  The  integral  jx  ~  2dx,  between  the 
limits  1  and  -  1  ought  to  be  given  by  the  area 
bounded  by  the  curve  y  =  x  - 2,  the  re-axis  and 
the  ordinates  corresponding  with  x  =  1,  and 
x  =  -  1.  Plot  the  curve  and  you  will  find  that 
this  result  is  erroneous.  The  curve  sweeps 
through  infinity,  whatever  that  may  mean,  as  x  FlGf  112^ 

passes  from  +  1  to  -  1  (Fig.  112).     The  method 
of  integration  is,  therefore,  unreliable  when  the  function  to  be 
integrated  becomes  infinite  or  otherwise  discontinuous  at  or  between 
the  limits  of  integration.     Consequently,  it  is  necessary  to  examine 

Q* 


244  HIGHER  MATHEMATICS.  §  83. 

certain  functions  in  order  to  make  sure  that  they  are  finite  and 
continuous  between  the  given  limits,  or  that  the  functions  either 
continually  increase  or  decrease,  or  alternately  increase  and  de- 
crease a  finite  number  of  times. 

This  subject  is  discussed  in  the  opening  chapters  of  B.  Eiemann 
and  H.  Weber's  Die  Partiellen  Differential-Gleichungen  der  mathe- 
matischen  Physik,  Braunschweig,  1900-1901,  to  which  the  student 
must  refer  if  he  intends  to  go  exhaustively  into  this  subject.  I  can, 
however,  give  a  few  hints  on  the  treatment  of  these  integrals.  It 
is  easy  to  see  that 


I  e~xdx  =  i  -  e1 


=  1  - 


and  if  n  is  made  infinitely  great,  the  integral  tends  towards  the 
limit  unity.     Hence  we  say  that 


I    e-*dx  =  1. 


If  the  function  is  continuous  for  all  values  of  x  between  a  and 
b,  except  when  x  =  b,  at  the  upper  limit,  it  is  obvious  that 

^J'{x)dx=Uh=^a   hf(x)dx  .        .         (7) 

if  h  is  diminished  indefinitely,  h,  of  course,  is  a  positive  number. 
And  in  a  similar  manner,  if  f(x)  is  continuous  for  all  values  of  x 
except  when  x  =  a,  at  the  lower  limit, 

[bf(x)dx  =  Lth,J      f'(x)dx.        .         .        (8) 

J  a  J  a+h' 

/1      dx                        fl~h      dx  [~  ~|1  -h 

,- =  Lt*=0  * =  Lt/*=o    _  2J1  -  x\ 

o  a/1  -  x  J  o     a/1  ~  x  L  -J° 

=  Lta  =  0{  -  2n/1  -  1  +  h  -  (  -  2)}  =  2  -  2slh. 
As  h  is  made  indefinitely  small,  the  integral  tends  towards  the  limit  2. 

f1     dx 
'  'J  a 


0  a/1    -   X 

fxdx  lldx  /l        \ 

(2)  Show  that   /    p  =  Lt*=0/    ^  =  Lt*=o(  ^  -11. 

As  h  is  made  very  small,  the  expression  on  the  right  becomes  infinite.     A 
definite  numerical  value  for  the  integral  does  not  exist. 

fa      dx  _         /"«-*     dx  -.■     '.    -if,       h\ 

(3)  Show  that  j-^  =  Lt1=,jo    -—  =  LWsrn      [l  -  a). 


§  84.  THE  INTEGRAL  CALCULUS.  245 

Since  when  h is  made  very  small  the  limit l  approaches  sin-  llt  or  £rr. 

fldx  ndx  1 

(4)   Show  that    /  —  =  Lt/i=0  /   —  =  log  1  -  log  h  =  log  ^  =  oo. 

When  the  function  f(x)  becomes  infinite  between  the  limits,  we 
write 


f\x)dx  =  LtA=0       f\x)dx  +  Ltw  =0         f'(x)dx. 

J  a  J  a  J  m-\-h> 


(9) 


if  f'(x)  only  becomes  infinite  at  the  one  point.     If  there  are  n  dis- 
continuities, we  must  obviously  take  the  sum  of  n  integrals. 

C1  dx  fldx  f~h'dx 

Examples.— (1)  J  _ -g  =  Lt»=o/Jfc^+  LtA/=0y  _jp-, 

=  LtA=0[-^+  Ltv=o[-j]_"  =  Lt»,0(l  -  £)+  LtA/=0(^  -  l), 

The  integral  thus  approaches  infinity  as  h  and  h'  are  made  very  small. 

/2    dx  f2        dx  r1-*    dx 

= u^4-^l^+Uh,4-^iV = Uh=li  -  x)+ "*-(»?  -  4 

as  7i  and  7*'  become  indefinitely  small,  the  limit  becomes  indefinitely  great, 
and  the  integral  is  indeterminate. 

f1     dx      6 
(3)  Show  that    J_i3^  =  2- 

It  would  now  do  the  beginner  good  to  revise  the  study  of  limits  by  the  aid 
of  say  J.  J.  Hardy's  pamphlet,  Infinitesimals  and  Limits,  Easton,  1900,  or  the 
discussions  in  the  regular  text-books. 

§  84.    To  find  the  Length  of  any  Curve. 

To  find  the  length,  I,  of  the  curve  AB  (Fig.  113)  when  the 
equation  of  the  curve  is  known.  This  is  equivalent  to  finding  the 
length  of  a  straight  line  of  the  same  length  as  the  curve  if  the  curve 
were  flattened  out  or  rectified,  hence  the  process  is  called  the  recti- 
fication of  curves.  Let  the  coordinates  of  A  be  (x0,  y0),  and  of 
B,  (xn)  yn).  Take  any  two  points,  P,  Q,  on  the  curve.  Make  the 
construction  shown  in  the  figure.  Then,  by  Euclid,  i.,  47,  if  P  and 
Q  are  sufficiently  close,  we  have,  very  nearly 

(PQf  =  (Zxf  +  (SyY  ;  or,  dl  =  J(dxf  +  {dyf 


1  Note  the  equivocal  use  of  the  word  limit.     There  is  a  difference  between  the 
limit "  of  the  differential  calculus  and  the  "  limit"  of  the  integral  calculus. 


246 


HIGHER  MATHEMATICS. 


§84. 


at  the  limit  when  the  length  of  the  chord  PQ  is  equal  to  the  length 
of  the  arc  PQ,  (1),  page  15.  Hence,  the  sum,  I,  of  all  the  small 
elements  dl  ranging  side  by  side  from  xY  to  x2  will  be 


-EV1 +©'■*• 


(i) 


In  order  to  apply  this  result  it  is  only  necessary  to  differentiate 
the  equation  of  the  curve  and  substitute 
the  values  of  dx  and  dy,  so  obtained,- in 
equation  (1).  By  integrating  this  equa- 
tion, we  obtain  a  general  expression  be- 
tween the  assigned  limits,  we  get  the 
length  of  the  given  portion  of  the  curve. 
If  the  equation  is  expressed  in  polar 
coordinates,  the  length  of  a  small  element, 
dl,  is  deduced  in  a  similar  manner.  Thus, 
dl  =  J  {drf  +  r\d$f.  ...  (2) 
The  mechanical  rectification  of  curves  in  practical  work  is  fre- 
quently done  by  running  a  wheel  along  the  curve  and  observing 
how  much  it  travels.  In  the  opisometer  this  is  done  by  starting 
the  wheel  from  a  stop,  running  it  along  the  path  to  be  measured  ; 
and  then  applying  it  to  the  scale  of  the  diagram,  running  it  back- 
wards until  the  stop  is  felt. 

Examples. — (1)  If  the  curve  is  a  common  parabola  y2=iax,  .'.  ydy=2adx; 
or,  {dx)2  =  2/2(%)2/4a2;  .-.  dl  =  J{y2  +  4a2)cfy/2a,  from  (1) ;  now  integrate,  as 
in  Ex.  (1),  page  203,  and  we  get  I  =  %y  Jy^+la2  +  2a2 log{(y  +  sly2  +  ±a2)j2a}  +  C. 
To  find  G,  put  y  =  0,  when  1  =  0;  .*.  C  =  -  2a2  log  2a. 


(2)  Show  that  the  perimeter  of  the  circle,  x2  +  y2 
the  length  of  the  arc  in  the  first  quadrant,  then  dyjdx 

.\l 


r2,  is  2ttt.     Let  I  be 

-xjy. 

.\  Whole  perimeter  =  4  x  %ttt  =■  2irr. 

(3)  Find  the  length  of  the  equiangular  spiral,  page  116,  whose  equation  is 
r  =  eO  ;  or,  0  =  log  r/log  e.  Ansr.  1=  si 2.  r.  Hint.  Differentiate ;  .-.  de=drjr, 
.-.  dl  =  sl2.  dr.     .\  I  =  \l2.r  +  G  ;  when  r  =  0,1  =  0,  G  =  0. 

(4)  The  length  of  the  first  whorl  of  Archimedes'  spiral  2ttt  =  a6  is  3*3885a. 
Verify  this.  Hint.  First  show  that  the  length  of  the  spiral  from  the  origin 
to  any  value  of  0  is  ^a/ir  x  {6  Jl  +  02  +  loge(0  +  sll  +  02)}.  For  the  first 
whorl,  0  =  2ir  =  6-2832  ;  sll  +  02  =  6 -363  ;  0  +  \/l  +  02  =  12-6462  ; 
loge(0  +  sll~+lP)  =  log,12-6462  =  2-5373.     Ansr.  =  a(3-1865  +  0*202). 


§  86.  THE  INTEGRAL  CALCULUS.  247 

(5)  Find  the  value  of  the  ratio 

_  I  _  Length  of  hyperbolic  arc  from  x  —  a  to  x  =  x 
~~  r  ~~  Distance  of  a  point  P(x,  y)  from  the  origin 

The  equation  of  the  rectangular  hyperbola  is  x1  -  y2  =  a2, .-.  y  =  sJx2  -  a2; 

.*.  dyjdx  =x/  s/x2  -  a2.     By  substitution  in  (1),  remembering  that  r=  six2  +  y2 ; 

y2  =  x2  -  a2 ;  .-.  r  =  *J<lx2  -  a2. 

L  rJ*Ek*.%  ,.i=  r1^==iog«_L^HZ. 

Ja\aj2-a2  r      J«  Va;2  -  a2  a 

We  shall  want  to  refer  back  to  this  result  when  we  discuss  hyperbolic  func- 
tions, and  also  to  show  that 

„u     x  +  s/x2  -  a2    /         a;\2      x2      ,    2xe»       _        „  a;      cM  +  *-«* 

The  reader  may  have  noticed  the  remarkable  analogy  between 
the  chemist's  "atom,"  the  physicist's  "particle,"  and  "molecule," 
and  the  mathematician's  "differential ".  When  the  chemist  wishes 
to  understand  the  various  transformations  of  matter,  he  resolves 
matter  into  minute  elements  which  he  calls  atoms  ;  so  here,  we 
have  sought  the  form  of  a  curve  by  resolving  it  into  small 
elements.  Both  processes  are  temporary  and  arbitrary  auxiliaries 
designed  to  help  the  mind  to  understand  in  parts  what  it  cannot 
comprehend  as  a  whole.  But  once  the  whole  concept  is  builded 
up,  the  scaffolding  may  be  rejected. 

§  85.  To  find  the  Area  of  a  Surface  of  Revolution. 

A  surface  of  revolution  is  a  surface  generated  by  the  rotation 
of  a  line  about  a  fixed  axis,  called  the  axis  of  revolution.  The 
quadrature  of  surfaces  of  revolution  is 
sometimes  styled  the  complanation  of 
surfaces.  Let  the  curve  APQ  (Fig.  114) 
generate  a  surface  of  revolution  as  it 
rotates  about  the  fixed  axis  Ox.  It  is 
required  to  find  the  area  of  this  surface. 

If  dx  and  dy  be  made  sufficiently  small,     q    L*      m m" 

we  may  assume  that  the  portion  (PQ)2,  or  F      114 

(diy  =  {dxf  +  (%)2,      .      (i) 

as  indicated  above.  The  student  is  supposed  to  know  that  the 
area  of  the  side  of  a  circular  cylinder  is  2-n-rh,  where  r  denotes  the 
radius  of  the  base  of  the  cylinder,  and  h  the  height  of  the  cylinder. 
The  surface,  ds,  of  the  cylinder  generated  by  the  revolution  of  the 


248 


HIGHER  MATHEMATICS. 


line  dl,  will  approach  the  limit 

ds  =  2irydl      ....         (2) 
as  the  length,  di,  at  P  is  made  infinitesimally  small.     Hence,  from 

(1),  and  (2),  

ds  =  2tt#  J(dxf  x  (dyf.  ...  (3) 
All  the  elements,  dl,  revolving  around  the  #-axis,  will  together  cut 
out  a  surface  having  an  area 


-M>V'+©v. 


W 


where  xY  and  x2  respectively  denote  the  abscissae  of  the  portion  of 
the  curve  under  investigation. 

Examples. — (1)  Find  the  surface  generated  by  the  revolution  of  the  slant 
side  of  a  triangle.  Hints.  Equation  of  the  line  OC  (Fig.  115)  is  y=mx\ 
.'.  dy  =  mdx,  ds  =  2vy  tjl+m?  .  dx,  s=j2irmN/l  +  m2  .  xdx  =  trmx2 „J1  +  m2  +  G. 
y  Reckon  the  area  from  the  apex,  where  x  =  0, 

therefore  C=0.  If  x  =  h= height  of  cone  = 
OB  and  the  radius  of  the  base  =  r  =  BC, 
then,  m  =  rfh  and 

=  ii7*  sJW  +  r2  =  2wr  x  £  slant  height. 
This  is  a  well-known  rule  in  mensuration. 

(2)  Show  that  the  surface  generated  by 
the  revolution  of  a   circle   is  47rr2.      Hint. 
Fig.  115.  x2  +  y2  =  r2.  dyjdx  =  _  ^ .  y  =  ^  _  ^  . 

.-.  2vjy  V(l  +  ^ly2)dx  becomes  2irr\dx  by  substituting  r2  =  x2  +  if.    The  limits 
of  the  integral  for  half  the  surface  are  x2  =  r,  and  x1  =  0. 

§  86.    To  find  the  Volume  of  a  Solid  of  Revolution. 

This  is  equivalent  to  finding  the  volume  of  a  cube  of  the  same 
capacity  as  the  given  solid.  Hence  the  process  is  named  the 
oubature  of  solids.  The  notion  of  differentials  will  allow  us  to 
deduce  a  method  for  finding  the  volume  of  the  solid  figure  swept 
out  by  a  curve  rotating  about  an  axis  of  revolution.     At  the  same 

time,  we  can  obtain  a  deeper  insight  into 
the  meaning  of  the  process  of  integra- 
tion. 

We  can,  in  imagination,  resolve  the 
solid  into  a  great  number  of  elementary 
parallel  planes,  so  that  each  plane  is  part 
of  a  small  cylinder.  Fig.  116  will,  per- 
haps, help  us  to  form  a  mental  picture  of 
the  process.     It  is  evident  that  the  total 


Fig.  116.— After  Cox. 


§  87.         THE  INTEGRAL  CALCULUS.  249 

volume  of  the  solid  is  the  sum  of  a  number  of  such  elementary 
cylinders  about  the  same  axis.  If  Sx  be  the  height  of  one  cylinder, 
y  the  radius  of  its  base,  the  area  of  the  base  is  iry2.  But  the  area 
of  the  base  multiplied  by  the  height  of  the  cylinder  is  the  volume 
of  each  elementary  cylinder,  that  is  to  say,  iry28x.  The  less  the 
height  of  each  cylinder,  the  more  nearly  will  a  succession  of  them 
form  a  figure  with  a  continuous  surface.  At  the  limit,  when 
Sx  =  0,  the  volume,  v,  of  the  solid  is 


rjV.da?,  (1) 


where  x  and  y  are  the  coordinates  of  the  generating  curve  ;  x1  and 
xn  the  abscissae  of  the  two  ends  of  the  revolving  curve  ;  and  the 
aj-axis  is  the  axis  of  revolution. 

The  methods  of  limits  can  be  used  in  place  of  the  method  of 
infinitesimals  to  deduce  this  expression,  as  well  as  (4)  of  the  pre- 
ceding section.  The  student  can,  if  he  wishes,  look  this  up  in 
some  other  text-book. 

Examples. — (1)  Find  the  volume  of  the  cone  generated  by  the  revolution 
of  the  slant  side  of  the  triangle  in  Ex.  1  of  the  preceding  section.  Here 
y  =  mx\  dv  =  7ry'2dx=irm2x'idx.  ,\  v  =  frjrm'2x*  +  C.  If  the  volume  be  reckoned 
from  the  apex  of  the  cone,  x  =  0,  and  therefore  C  =  0.  Let  x  =  h  and  ra=r/7fc, 
as  before,  and  the 

Volume  of  the  entire  cone  =  ^nr2h. 

(2)  Show  that  the  volume  generated  by  the  revolving  parabola,  t/3  =  4aa;, 
is  frry*xt  where  x  =  height  and  y= radius  of  the  base. 

(3)  Required  the  volume  of  the  sphere  generated  by  the  revolution  of  a 
circle,  with  the  equation  :  x2  +  y2  =  r2.  Volume  of  sphere  =  |^r3.  Hint. 
v=*TJ(r2  -  x"*)dx ;  use  limits  for  half  the  surface  x.>  =r,  xx=0. 

§  87.   Successive  Integration.    Multiple  Integrals. 

Just  as  it  is  sometimes  necessary,  or  convenient,  to  employ 
the  second,  third  or  the  higher  differential  coefficients  d2yldx2, 
d3y/dx* . . .  ,  so  it  is  often  just  as  necessary  to  apply  successive  in- 
tegration to  reverse  these  processes  of  differentiation.  Suppose 
that  it  is  required  to  reduce,  <Py/dx2  =  2,  to  its  original  primitive 
form.     We  can  write  for  the  first  integration 

,\  dy  =  (2x  +  CJdx ;  or,  y  =  |(2#  +  CJdx ;  .*.  y  =  x2  +  Cxx  +  C2. 


250 


HIGHER  MATHEMATICS. 


§87. 


In  order  to  show  that  d2yldx2  is  to  be  integrated  twice,  we  affix 
two  symbols  of  integration. 

y  =  IS2dx.dx,  .-.  y  =  x2  +  O^x  +  G2. 
Notice  that  there  are  as  many  integration  constants,   Cv  C2,  as 
there  are  symbols  of  integration. 

Examples. — (1)  Find  the  value  of  y  =  jjjsc3 . dx . dx.dx.    Ansr. 

(2)  Integrate  cPsfdfi  —  g,  where  g  is  a  constant  due  to  the  earth's  gravita- 
tion, t  the  time  and  s  the  space  traversed  by  a  falling  body. 


•••«-j]^+<*** 


To  evaluate  the  constants  Cx  and  C2,  when  the  body  starts  from  a  position 
of  rest,  s  =  0,  t  =  0,  Cj  =  0,  C2  =  0. 

In  finding  the  area  of  a  curve  y  =  f(x),  the  same  result  will  be 
obtained  whether  we  divide  the  area  Oab  into 
a  number  of  strips  parallel  to  the  y-axis,  as  in 
Fig.  117,  or  strips  parallel  to  the  #-axis,  Fig. 
118.  In  the  second  case,  the  reader  will  no 
doubt  be  able  to  satisfy  himself  that  the  area 


Fio.  117.— Surface 
Elements. 


=  I  a? .  dy  ; 


and,  in  the  second  case,  that 


Jo 


dx. 


(i) 


(2) 


There  is  another  way  of  looking  at  the  matter.  Suppose  the 
surface  is  divided  up  into  an  infinite  number  of  infinitely  small  rect- 
angles as  illustrated  in  Fig.  119.     The  area  of  each  rectangle  will 


Fig.  118.— Surface  Elements. 


0LTT 

Fig.  119.— Surface  Elements. 


be  dx.  dy.      The  area  of  the  narrow  strip  Obcd  is  the  sum  of  the 
areas  of  the  infinite  number  of  rectangles  ranged  side  by  side  along 


§  87.  THE  INTEGRAL  CALCULUS.  251 

this  strip  from  0  to  b.     The  length  of  this  strip,  y  =  b,  and  the 
width,  dx,  is  constant ;  consequently, 


Area  of  strip  Obcd  =  dx  I  dy. 

Jo 


The  total  area  of  the  surface  Obcd  is  obviously  the  sum  of  the 
areas  of  the  infinite  number  of  similar  strips  ranged  along  Oa. 
The  height  of  the  second  strip,  say,  dcef  obviously  depends  upon 
the  nature  of  the  curve  ba.  If  the  equation  of  this  line  be  repre- 
sented by  the  equation, 

H-j «) 

the  height  y  of  any  strip  at  a  distance  x  from  0  is  obtained  by 
solving  (3)  for  y.     Consequently, 

y  -  -(a  -  x) (4) 

The  area  of  any  strip  lying  between  0  and  a  is  therefore 

Area  of  any  strip  =  dx\  dy  =  dx\  aV    \  =  -(a~x)dx        (5) 

and  it  follows  naturally  that  the  area  of  all  the  strips,  when  each 
strip  has  an  area  b(a  -  x)dx/a,  will  be 

(ab(a~xh       Yabx-\bx2~\a    a?b     \tfb     ab 

Area  of  all  the  strips  =  1  -* J-dx  =     1 =  —  -  I =         ,q\ 

Jo     a  L       a       Jq      a       a       2   v> 


Combining  (5)  and  (6),  into  one  expression,  we  get 

Ma^x)  rn    ^(a  -  x) 

4  =   I  dx  I  (7) 


=   I  dx  \  dy  =   I    I  dx.dy, 

J  o      J  o  J  qJ  o 


which  is  called  a  double  integral.  This  integral  means  that  if  we 
divide  the  surface  into  an  infinite  number  of  small  rectangles — 
surface  elements — and  take  their  sum,  we  shall  obtain  the  re- 
quired area  of  the  surface. 

To  evaluate  the  double  integral,  first  integrate  with  respect  to 
one  variable,  no  matter  which,  and  afterwards  integrate  with 
respect  to  the  other.  If  we  begin  by  keeping  x  constant  and 
integrating  with  respect  to  y,  as  y  passes  from  0  to  b,  we  get  the 
area  of  the  vertical  strip  Obcd  (Fig.  119) ;  we  then  take  the  sum  of 
the  rectangles  in  each  vertical  strip  as  x  passes  from  0  to  a  in 


252  HIGHER  MATHEMATICS.  §  87. 

such  a  way  as  to  include  the  whole  surface  ObaO.  When  there 
can  be  any  doubt  as  to  which  differential  the  limits  belong,  the 
integration  is  performed  in  the  following  order :  the  right-hand 
element  is  taken  with  the  first  integration  sign  on  the  right,  and 
so  on  with  the  next  element.  It  just  happens  that  there  is  no 
special  advantage  in  resorting  to  double  integration  in  the  above 
example  because  the  single  integration  involved  in  (1)  or  (2)  would 
have  been  sufficient.  In  some  cases  double  integration  is  alone 
practicable.  The  application  of  the  integral  calculus  to  this  simple 
problem  in  mensuration  may  seem  as  incongruous  as  the  employ- 
ment of  a  hundred-ton  steam  hammer  to  crack  nuts.  But  I  have 
done  this  in  order  that  the  attention  might  be  alone  fixed  upon 
the  mechanism  of  the  hammer. 

Examples. — (1)  Show  that  if  the  curve  ab  (Fig.  119)  be  represented  by 
equation  (3),  then  the  area  of  the  surface  bounded  by  ab,  and  the  two  co- 
ordinate axes,  may  be  variously  represented  by  the  integrals 

hJQ(b  -  y)dy;  -]  Ja  -  x)dx;  JJ^  dy.dx;  J  J  dx.dy. 

(2)  Show  /    \  x. dx.dy =     x.dx  [VT=  3  /  x  • dx  =  3  \\  =  7$- 

fa    fb  a263 

(3)  Show  /     /  xy*.dx.dy=—. 

(4)  Show  that  the  area  bounded  by  the  two  parabolas  3y2  =  25#  *,  and 
5x2  =  9y  is  5  units. 

The  areas  of  curves  in  polar  coordinates  may  be  obtained  in  a 
similar  manner.  Divide  the  given  surfaces  up  into  slices  by  drawing 
radii  vectores  at  an  angle  dd  apart,  and  subdivide  these  slices  by 
drawing  arcs  of  circles  with  origin  as  centre.     Consider  any  little 

surface  element,  say,  PQBS  (Fig. 
120).  OPQ  may  be  regarded  as  a 
triangle  in  which  PQ  =  OQ  sin  (dO). 
But  the  limiting  value  of  the  sine 
of  a  very  small  angle  is  the  angl© 
itself,  and  since  OQ  =  r,  we  have 
FlG' 120'  QP  =  rdO.     Now  PS  is,  by  con- 

struction, equal  to  dr.  The  area  of  each  little  segment  is,  at  the 
limit,  equal  to  PQ  x  PS,  or 

dA=r.dr.dO.  ...  (8) 

The  total  area  will  be  found  by  first  adding  up  all  the  surface 


§  87.         THE  INTEGRAL  CALCULUS.  253 

elements  in  the  sector  OBC,  and  then  adding  up  all  the  sectors  like 
COB  which  it  contains,  or, 

A  =[*[%.  dr.  dO.       . 


}r)e1 


Example. — Find  the  area  of  the  circle  whose  equation  is  r  =  2a  cos  0, 
where  r  denotes  the  radius  of  the  circle.     Ansr. 

f2a  cos  0 


[•la cose  ri.tr  ,       _  „ 

=  /  I*  nr  .dr  .dd  =  ira\ 


We  can  also  imagine  a  solid  to  be  split  up  into  an  infinite 
number  of  little  parallelopipeds  along  the  three  dimensions  x,  y,  z. 
These  infinitesimal  figures  may  be  called  volume  elements.  The 
capacity  of  each  little  element  dx  x  dy  x  dz.  The  total  volume, 
v,  of  the  solid  is  represented  by  the  triple  integral 


\\\dx.dy 


(10) 


The  first  integration  along  the  #-axis  gives  the  length  of  an 
infinitely  narrow  strip ;  the  integration  along  the  y-axis  gives  the 
area  of  the  surface  of  an  infinitely  thin  slice,  and  a  third  integra- 
tion along  the  2-axis  gives  the  total  volume  of  all  these  little  slices, 
in  other  words,  the  volume  of  the  body. 

In  the  same  way,  quadruple  and  higher  integrals  may  occur. 
These,  however,  are  not  very  common.  Multiple  integration  rarely 
extends  beyond  triple  integrals. 

Examples. — (1)  Evaluate  the  following  triple  integrals  : — 

I     /     \  yz2 .  dx  .  dy  .  dz  ;    j    /    j  yz2  .dy  ,dz  .dx;    J     /    /  yz2 .  dz  .  dx .  dy. 

Ansrs.  2580,  1550,  1470  respectively. 
(2)  Show 


/•«  f/x2       y2\J      J        b    [»(x2     &2\  ah  (a2      62\ 

fr    f   V(r2-*2)    r    ^(r2~x2-v2)  A 

(3)  Evaluate  8  dx.dy.dz.    Ansr.  - 

J o  J o  Jo 


Note  sin  %w  =  1.  Show  that  this  integral  represents  the  volume  of  a  sphere 
whose  equation  is  x2  +  y2  +  z2  =  r2.  Hint.  The  "  dy  "  integration  is  the 
most  troublesome.  For  it,  put  r2  -  x2  =  c,  say,  and  use  Ex.  (1),  p.  203.  As  a 
result,  %y  sir2  -  x2  -  y2  +  £(r2  -  x2)  sin  -^{y/sJr2  -  x2}  has  to  be  evaluated 
between  the  limits  y  =  sj(r2  -  x2)  and  y  =  0.  The  result  is  l(r2-x2)ir.  The 
rest  is  simple  enough. 


254  HIGHER  MATHEMATICS.  §.88 

§  88.  The  Isothermal  Expansion  of  Gases. 

To  find  the  work  done  during  the  isothermal  expansion  of  a  gas, 
that  is,  the  work  done  when  the  gas  changes  its  volume,  by  ex- 
pansion or  compression,  at  a  constant  temperature.  A  contraction 
may  be  regarded  as  a  negative  expansion.  There  are  three  in- 
teresting applications. 

I. — The  gas  obeys  Boyle's  law,  pv  =  constant,  say,  o.  We  have 
seen  that  the  work  done  when  a  gas  expands  against  any  external 
pressure  is  represented  by  the  product  of  the  pressure  into  the 
change  of  volume.     The  work  performed  during  any  small  change 

of  volume,  is 

dW  =  p.dv.      .         .         .         .         (1) 
But  by  Boyle's  law, 

P>/(»)-|  •     '   •        •        •        (2) 

Substitute  this  value  of  p  in  (1);  and  we  get  dW  =  g  .  dv/v.     If  the 
gas  expands  from  a  volume  vx  to  a  new  volume  v2,  it  follows 

—  =  c      \ogv  +  G.     .:  W  =  p^log  -*   .         (3) 

From  (2),  vx  =  c/pv  and  also  v2  =  c/p2,  consequently 

W^p^log^.     .         .         .        .        (4) 

Equations  (3)  and  (4)  play  a  most  important  part  in  the  theory 
of  gases,  in  thermodynamics  and  in  the  theory  of  solutions.  The 
value  of  6  is  equal  to  the  product  of  the  initial  volume,  vv  and 
pressure,  plf  of  the  gas.     Hence  we  may  also  put 

W  =  2-3026^1log1(£  =  2-3026ftVog1(,;p 

V2  Pi 

for  the  work  done  in  compressing  the  gas. 

Example. — In  an  air  compressor  the  air  is  drawn  in  at  a  pressure  of 
14*7  lb.  per  square  inch,  and  compressed  to  77  lb.  per  square  inch.  The 
volume  drawn  in  per  stroke  is  1*52  cubic  feet,  and  133  strokes  are  made  per 
minute.  What  is  the  work  of  isothermal  compression  ?  Hint.  The  work 
done  is  the  compression  of  1*52  cubic  feet  x  133  =  202*16  cubic  feet  of  air  at 
14-7  lb.  to  77  lb.  per  square  inch,  or  14*7  x  144  =  2116-8  lb.  to  77  x  144  = 
110881b.  per  square  foot.  From  Boyle's  law,  p^  =  p2v2;  ,-.v2  x  77  = 
14*7  x  202-16 ;  or,  v2  =  38-598.  From  the  above  equation,  therefore,  the 
work  =  2-3026  x  2116-8  x  202-16  (log  202-16  -  log  38-598)  =  708757*28  foot 
pounds  per  minute  ;  or,  since  a  "  horse  power  "  can  work  33,000  foot  pounds 
per  minute,  the  work  of  isothermal  compression  is  21*48  H.P. 


§  88.  THE  INTEGRAL  CALCULUS.  255 

II. — The  gas  obeys  van  der  Waals'  law,  that  is  to  say, 

\P  +  ~2)(v  ~  ^)  ~  constant>  say,  C. 

As  an  exercise  on  what  precedes,  prove  that 

W=clogV-^4-a(±-±);       .         .  (S) 

6  vY  -  b         \Vj       v2J '  v  ' 

This  equation  has  occupied  a  prominent  place  in  the  development 

of  van  der  Waals'  theories  of  the  constitution  of  gases  and  liquids. 

Example. — Find  the  work  done  when  two  litres  of  carbon  dioxide  are 
compressed  isothermally  to  one  litre  ;  given  van  der  Waals'  a  =  0"00874: ; 
b  =  0-0023 ;  c  =  0*00369.  Substitute  in  (5),  using  a  negative  sign  for  con- 
traction. 

III. — The  gas  dissociates  during  expansion.  By  Guldberg  and 
Waage's  law,  in  the  reaction : 

N204-2N02, 
for  equilibrium,  if  x  denotes  that  fraction  of-  unit  mass  of  N204 
which  exists  as  N02,  we  must  have 

^1  -  x      xx 

V  V     V 

where  (1  -  x)/v  represents  the  concentration  of  the  undissociated 
nitrogen  peroxide.  The  relation  between  the  volume  and  degree 
of  dissociation  is,  therefore, 

*-r?s (6) 

If  n  represents  the  original  number  of  molecules  ;  (1  -  x)n  will 
represent  the  number  of  undissociated  molecules;  and  2xn  the 
number  of  dissociated  molecules.  If  the  relation  pv  =  g  does  not 
vary  during  the  expansion,  the  pressure  will  be  proportional  to  the 
number  of  molecules  actually  present,  that  is  to  say,  if  p  denotes 
the  pressure  when  there  was  no  dissociation,  and  p'  the  actual 
pressure  of  the  gas, 

p_ n 1 

p'  ~  (1  -  x)n  +  2xn  ~"  1  +  x 
The  actual  pressure  of  the  gas  is,  therefore,  p  =  (1  +  x)p ;  and 
the  work  done  is, 

dW  =  p' . dv  =  (1  +  x)p .dv  =  p.dv  +  xp.dv,    .        (7) 
From  Boyle's  law,  and  (6),  we  see  that 

c      cK(l  -  x) 

F       V  X* 


256  HIGHER  MATHEMATICS.  §  88. 

Substitute  this  value  of  p  in  (7).     Differentiate  (6)  and  we  obtain 

dv  _  2(1  -  x)x  +  x2  _   x(2  -  x)  , 

dx  "      K(l  -  xf     ;  ;"'  av  ~  JT(1  -  xTx; 

Now  substitute  this  value  of  dv  in  (7) ;  simplify,  and  we  get 

where  x1  and  x2  denote  the  values  of  x  corresponding  with  i^  and 
v2.     On  integration,  therefore, 

W -  c(log^  +  ^  -  «,  -  log^J)  .        .        (8) 
It  follows  directly  from  (6),  that 


vi  ~  T7-/1    \  »  and>  V2  = 


Substitute  these  values  of  v  in  (8),  and  the  work  of  expansion 


(9) 


Examples. — (1)  Find  the  work  done  during  the  isothermal  expansion  of 
dissociating  ammonium  carbamate  (gas) :  NH2COONH4  ^  2NH3  +  C02. 

(2)  In  calculating  the  work  done  during  the  isothermal  expansion  of 
dissociating  hydrogen  iodide,  2HI  ^  H2  +  1^,  does  it  make  any  difference 
whether  the  hydrogen  iodide  dissociates  or  not  ? 

(3)  A  particle  of  mass  m  moves  towards  a  centre  of  force  F  which  varies 
inversely  as  the  square  of  the  distance.  Determine  the  work  done  by  the 
force  as  it  moves  from  one  place  r2  to  another  place  rv  Work  =  force  x 
displacement 

•••-=/::-*=/;?*hh> 

If  r  is  infinite,  W  =  ra/r.  If  the  body  moves  towards  the  centre  of  attraction 
work  is  done  by  the  force ;  if  away  from  the  centre  of  attraction,  work  is 
done  against  the  central  force. 

(4)  If  the  force  of  attraction,  F,  between  two  molecules  of  a  gas,  varies 
inversely  as  the  fourth  power  of  the  distance,  r,  between  them,  show  that  the 
work,  W,  done  against  molecular  attractive  forces  when  a  gas  expands  into 
a  vacuum,  is  proportional  to  the  difference  between  the  initial  and  final 
pressures  of  the  gas.  That  is,  W  =  A(px  -  _p2),  where  A  is  the  variation  con- 
stant. By  hypothesis,  F=ajri ;  and  dW=F.  dr,  where  a  is  another  variation 
constant.     Hence, 


rrJV*-.£r 


§  89.  THE  INTEGRAL  CALCULUS.  257 

But  r  is  linear,  therefore,  the  volume  of  the  gas  will  vary  as  r3.     Hence, 
v  =  br5,  where  b  is  again  constant. 

•  *  W  ~  3\r\      f\)  '  8  U      «J* 

But  by  Boyle's  law,  pv  =  constant,  say,  c.     Hence  if  A  =  <2&/3c  =  constant, 

(5)  J/'  the  work  done  against  molecular  attractive  forces  when  a  gas 
expands  into  a  vacuum,  is 


where  a  is  constant ;  vv  v2,  refer  to  the  initial  and  final  volumes  of  the  gas, 
show  that  "  any  two  molecules  of  a  gas  will  attract  one  another  with  a  force 
inversely  proportional  to  the  fourth  power  of  the  distance  between  them  ". 
For  the  meaning  of  a/v2,  see  van  der  Waals'  equation. 

§  89.    The  Adiabatic  Expansion  of  Gases. 

When  the  gas  is  in  such  a  condition  that  no  heat  can  enter  or 
leave  the  system  during  the  change  of  volume — expansion  or  con- 
traction— the  temperature  will  generally  change  during  the  operation. 
This  alters  the  magnitude  of  the  work  of  expansion.  Let  us  first 
find  the  relation  between  p  and  v  when  no  heat  enters  or  leaves 
the  gas  while  the  gas  changes  its  volume.  Boyle's  relation  is 
obscured  if  the  gas  be  not  kept  at  a  constant  temperature. 

J. — The  relation  between  the  pressure  and  the  volume  of  a  gas 
when  the  volume  of  the  gas  changes  adiabatically.  In  example 
(5)  appended  to  §  27,  we  obtained  the  expression, 

«-($*♦($*•  •  •  » 

As  pointed  out  on  page  44,  we  may,  without  altering  the  value  of 
the  expression,  multiply  and  divide  each  term  within  the  brackets 
bydT.     Thus, 

But  (bQfiT)p  is  the  amount  of  heat  added  to  the  substance  at  a 
constant  pressure  for  a  small  change  of  temperature  ;  this  is  none 
other  than  the  specific  heat  at  constant  pressure,  usually  written 
Cp.  Similarly  (dQfdT),  is  the  specific  heat  at  constant  volume, 
written  0„.     Consequently, 


dQ 


-^dv  +  GQdp.       .        .        (3) 


258  HIGHER  MATHEMATICS.  §  89. 

This  equation  tells  that  when  a  certain  quantity  of  heat  is  added 
to  a  substance,  one  part  is  spent  in  raising  the  temperature  while 
the  volume  changes  under  constant  pressure,  and  the  other  part  is 
spent  in  raising  the  temperature  while  the  pressure  changes  under 
constant  volume.     Eor  an  ideal  gas  obeying  Boyle's  law, 

"-» ■■■■J -CD.'*-®; 

Substitute  these  values  in  (3),  and  we  get 

after  dividing  through  with  6  =  pv/B.  By  definition,  an  adiabatic 
change  takes  place  when  the  system  neither  gains  nor  loses  heat. 
Under  these  conditions,  dQ  =  0 ;  and  remembering  that  the  ratio 
of  the  two  specific  heats  Cp/Gv  is  a  constant,  usually  written  y ; 

Cv   dv      dp      „  Cdv       [dp 

.       .'.  7^ +  -£■  =  0  ;   or,  y   —  +      —  =  Constant. 

G,     V  p  '      '  'J  V         )p 

or,  y  log  V  +  logp  =  const.  ;  or,  log  vy  +  log p  =  const.  ;   . '.  log  (pvy)  =  const. 

•\  pvy  =  c (5) 

A  most  important  relation  sometimes  called  Poisson's  equation. 

By  integrating  between  the  limits  pv  p2  ;  and  vv  v2  in  the  above 
equation,  we  could  have  eliminated  the  constant  and  obtained  (5) 
in  another  form,  namely, 


Pg  =  (h\y 
Pi       W  " 


(6) 


The  last  two  equations  tell  us  that  the  adiabatic  pressure  of  a  gas 
varies  inversely  as  the  yth  power  of  the  volume.  Now  substitute 
vl  =  TlBlpl ;  and  v2  =  T2B/p2,  in  (6),  the  result  is  that 

fc^-fr- ©"-(S)(?J "-®'-®~  "> 

and  the  relation  between  the  volume  and  temperature  of  a  gas  under 
adiabatic  conditions  assumes  the  form, 

r.       A>,\r-» 


Z7,      \vt 


Y 


i      •  ...        (8) 

V 

This  equation  affirms  that  for  adiabatic  changes,  the  absolute  tem- 
perature of  a  gas  varies  inversely  as  the  (y  -  l)th  power  of  the 
volume.     A  well-known  thermodynamic  law. 

Again,   since  "weight  varies  directly  as  the  volume,"   if  wl 


§  89.  THE  INTEGRAL  CALCULUS.  259 

denotes  the  weight  of  v  volumes  of  the  gas  at  a  pressure  p  l  and  w2 
the  weight  of  the  same  volume,  at  a  pressure  p2,  we  see  at  once, 
from  (6),  that 

II. — The  work  performed  when  a  gas  is  compressed  under  adia- 
batic  conditions.  From  (5),  p  =  c/vy  ;  and  we  know  that  the  work 
done  when  v  volumes  of  a  gas  are  compressed  from  vl  to  v2,  is 

f'2  p2  dv  r  v-<y-i>"|"2 

•••^'F-ife^-^)    •   •  (10) 

From  (5),  c  =  p^  =  p2v2y.  We  may,  therefore,  represent  this 
relation  in  another  form,  viz.  : 

0    (    1  1     \        1    /ffiV     ffiV\        1    (V<Pl      ffiV\ 

If  a  gas  expands  adiabatically  from  a  pressure  px  to  a  pressure 
p2,  we  get,  from  (5)  and  (11), 

w  =  —rjiPi1'*  -  ft1**)**  — r^^Tj-ft^iw1"  (12) 

provided  we  work  with  unit  volume,  v1  =  1,  of  gas,  so  that  _p:  =  c. 
If  _p1v1  =  BTX ;  and_p2-y2  =  i2T2,  are  the  isothermal  equations 
for  Tj0  and  T2°,  we  may  write, 

^--ItW-ZY),       •        •        •      (13) 

which  states  in  words,  that  the  work  required  to  compress  a  mass 
of  gas  adiabatically  while  the  temperature  changes  from  T-f  to  T2°, 
will  be  independent  of  the  initial  pressure  and  volume  of  the  gas. 
In  other  words,  the  work  done  by  a  perfect  gas  in  passing  along 
an  adiabatic  curve,  from  one  isothermal  to  another,  page  111,  is 
constant  and  independent  of  the  path. 

Examples. — (1)  Two  litres  of  a  gas  are  compressed  adiabatically  to  one 
litre.  What  is  the  work  done?  Given  y  =  1-4  ;  atmospheric  pressure^?  = 
1-03  kilograms  per  sq.  cm.  Ansr.  16-48  kilogram  metres.  Hint.  v9  =  %ox ; 
from  (6),  p2  =  £>i2Y ;  vx  =  2000  c.c.  From  a  table  of  common  logs, 
log  2?  =  log  21'4  =  0-30103  x  1-4  =  0-4214  j  or  21"4  =  2-64.     From  (11), 


260 


HIGHER  MATHEMATICS. 


§ 


W: 


-(1-32  -  1),  etc. 


V2P1  •  21'4  ~  *>iPi  =  PjPift  x  21'4  -  1)  _  1-03  x  2000, 
7-1  1-4-1  ~         0-4 

(2)  To  continue  illustration  3,  §  20,  page  62.  We  have  assumed  Boyle's 
law  p2p1  =  p^p2.  This  is  only  true  under  isothermal  conditions.  For  a  more 
correct  result,  use  (5)  above.  For  a  constant  mass,  m,  of  gas,  m  =  pv,  hence 
show  that  for  adiabatic  conditions, 


L 

y  e 


y 


f-&**l''.-*:7?    y       ■    ■    (U) 

is  the  more  correct  form  of  Halley's  law  for  the  pressure,  p2,  of  the  atmos- 
phere at  a  height  h  above  sea-level.     Atmospheric  pressure  at  sea-level  =  pv 

(3)  From  the  preceding  example  proceed  to  show  that  the  rate  of  diminu- 
tion of  temperature,  T,  is  constant  per  unit  distance,  h,  ascent.  In  other 
words,  prove  and  interpret 

T'-T=W1~h (!5> 

(4)  A  litre  of  gas  at  0°  0.  is  allowed  to  expand  adiabatically  to  two  litres. 
Find  the  fall  of  temperature  given  y  =  1-4.     Ansr.  66°  C.  (nearly).     Hints. 


vl  =  2v2;  from  (8),  T2  x  20'4  =  273  ;  20-1  =  1-32, 


207°  ;  there  is  there- 


fore a  fall  of  273  -  207°  absolute,  =  66°  C. 

(5)  To  continue  the  discussion  §§  15  and  64,  suppose  the  gas  obeys  van  der 
Waais'  law : 


(*+£)(V 


b)=RT. 


(16) 


where  R,  a,  6,  are  known  constants.  The  first  law  of  thermodynamics  may 
be  written 

dQ=  C,.dT+(p  +  a/v2)dv,  .        .        .        (17) 

where  the  specific  heat  at  constant  volume  has  been  assumed  constant.  To 
find  a  value  for  Cp,  the  specific  heat  at  constant  pressure.  Expand  (16). 
Differentiate  the  result.  Cancel  the  term  2ab  .  dv/v3  as  a  very  small  order  of 
magnitude  (§  4).  Solve  the  result  for  dv.  Multiply  through  with  p  +  a/v2. 
Since  ajv2  is  very  small,  show  that  the  fraction  (p  +  a/v2)/(p  -  afv2)  is  very 
nearly  1  +  2a/pv2.    Substitute  the  last  result  in  (17),  and 

dQ  =  {c„  +  B(l  +  ^)}dT  -  (l  +  j£)  (V  -  b)dp. 

By  hypothesis  Cv  is  constant, 

Cp  Rf        2a  \ 

•••cf.  =  1  +  c.(1+^) <18> 

For  ideal  gases  a  =  0,  and  we  get  Mayer's  equation,  §  27.  From  Boynton, 
p.  114; 


For. 

Air. 

Hydrogen. 

Carbon 
Dioxide. 

a 

B/O. 

y  Calculated  (18) 

y  Observed           .... 

0-002812 
0-4 

1-40225 
1-403 

0-0000895 
0-4 

1-40007 
1-4017 

0-00874 
0-2857 
1-2907 
1-2911 

§89. 


THE  INTEGRAL  CALCULUS. 


261 


(6)  Show  van  der  Waals'  equation  for  adiabatic  conditions  is 

(p  +  £)(v  -  b)y  =  BTt        ....      (19) 
and  the  work  of  adiabatic  expansion  is 

^=B(T1-r2){^1-a(i-i)}..        .        .      (20) 

(7)  Calculate  the  work  done  by  a  gas  which  is  compressed  adiabatically 
from  a  state  represented  by  the  point  A  (Fig.  121)  along  the  path  AB  until  a 
state  B  is  reached.  It  is  then  allowed  to  expand 
isothermally  along  the  path  BG  until  a  state  C  is 
reached.  This  is  followed  by  an  adiabatic  expan- 
sion along  CD  ;  and  by  an  isothermal  contraction 
along  DA  until  the  original  state  A  is  reached. 
The  total  work  done  is  obviously  represented  by 
the  sum  of 

-  AabB  +  BCcb  +  CDdc  -  DdaA. 
By  evaluating  the  work  in  each  operation  as  indi- 
cated in  thelasttwosections,on  the  assumption  that 
the  equation  of  AB  is  pvy=cx ;  of  BC>  pv=c^\  CD 


P»T  =  63 

gas,  is 


DA,pv  =  c4 


:Cj ;  oi  &u,pv  =  c2; 
Hence  show  that  the  external  work,  Wt  done  by  the 


W 


flog 


(8)  Compare  the  work  of  isothermal  and  adiabatic  compression  in  the 
example  on  page  254.  Take  y  for  air  =  1-408.  Hint.  From  (6),  14-7  x 
206-161'408  =  77  x  vj-m;  .-.  v2  =  62-36  cubic  feet;  and  from  (10),  (62-36  x 
11,088  -  202-16  x  2116-8)/0-408  =  645-871  foot  lbs.  per  minute  =  19-75  H.P. 
The  required  ratio  is  therefore  as  1  :  0-91. 

(9)  If  a  gas  flows  adiabatically  from  one  place  where  the  pressure  is  px  to 
another  place  where  the  pressure  is  p2,  the  work  of  expansion  is  spent  in 
communicating  kinetic  energy  to  the  gas.  Let  V  be  the  velocity  of  flow. 
The  kinetic  energy  gained  by  the  gas  is  equal  to  the  work  done.  But  kinetic 
energy  is,  by  definition,  \mV\  where  m  is  the  mass  of  the  substance  set  in 
motion;  but  we  know  that  mass  =  weight  -f  g,  hence,  if  wx  denotes  the 
weight  of  gas  flowing  per  second  from  a  pressure  px  to  a  pressure  p2t 

mV*  _wxV2 
*'•     2     ~    2g 

If  a  denotes  the  cross  sectional  area  of  the  flowing  gas,  obviously,  w1  =  aVw2, 
where  w2  denotes  the  weight  of  unit  volume  of  the  gas  at  a  pressure  p2.  Let 
vJPi  =  2>    From  (9),  w2  =  w^n.    Hence  the  weight  of  gas 


=  W; 


Vw*, 


w,  =  a  Vw* 


-W^P^W- 


Now  multiply  through  with  plW  ;  then  with  the  denominator  of  pxjpx ',  then 
with  w2/wx,  or,  what  is  the  same  thing,  with  q}W  ;  substitute  w2  =  w1g1/y,  and 
multiply  through  with  the  last  result.     The  weight  of  gas  which  passes  per 


262  HIGHER  MATHEMATICS.  §  90. 

second  from  a  pressure  px  to  a  pressure  p2  is  then 


w 
Wj  will  be  a  maximum  when 


i  =  aVW^.;+;). 


V 

2  =  i(y  +  i)1 "  y. 


For  dry  steam,  y  =  1*18,  and  hence, 

log<#  =  -  8-7  x  logel-065  =  1-762  ;  .-.  q  =  0-58 ;  or,  p2  =  0-58^ ; 
or  there  will  be  a  maximum  flow  when  the  external  pressure  is  a  little  less 
than  half  the  supply  pressure.     This  conclusion  was  verified  by  the  experi- 
ments of  Navier. 


§  90.    The  Influence  of  Temperature  on  Chemical  and 
Physical  Changes. 

On  page  82,  (18),  we  deduced  the  formula, 

m,-m  ■  ■  •  » 

by  a  simple  process  of  mathematical  reasoning.  The  physical 
signification  of  this  formula  is  that  the  change  in  the  quantity 
of  heat  communicated  to  any  substance  per  unit  change  of  volume 
at  constant  temperature,  is  equal  to  the  product  of  the  absolute 
temperature  into  the  change  of  pressure  per  unit  change  of  temper- 
ature at  constant  volume. 

Suppose  that  1  -  x  grams  of  one  system  A  is  in  equilibrium 
with  x  grams  of  another  system  B.  Let  v  denote  the  total  volume, 
and  T  the  temperature  of  the  two  systems.  Equation  (1)  shows 
that  (dQ/~dv)T  is,  the  heat  absorbed  when  the  very  large  volume  of 
system  A  is  increased  by  unity  at  constant  temperature  T,  less  the 
work  done  during  expansion.  Suppose  that  during  this  change  of 
volume,  a  certain  quantity  (bxfiv)T  of  system  B  is  formed,  then,  if 
q  be  the  amount  of  heat  absorbed  when  unit  quantity  of  the  first 
system  is  converted  into  the  second,  the  quantity  of  heat  absorbed 
during  this  transformation  is  q(dx/'dv)r.  q  is  really  the  molecular 
heat  of  the  reaction. 

The  work  done  during  this  change  of  volume  is  p  .  dv  ;  but  dv 
is  unity,  hence  the  external  work  of  expansion  is  p.  Under  these 
circumstances, 


§  90.  THE  INTEGRAL  CALCULUS.  263 

from  (1).     Now  multiply  and  divide  the  numerator  by  the  inte- 
grating factor,  T2 ; 

-®)M$);       ■     ■     * 

If,  now,  wx  molecules  of  the  system  A  ;  and  n2  molecules  of  the 
system  B,  take  part  in  the  reaction,  we  must  write,  instead  of 
pv  =  BT, 

pv  =  BT{nx(l  -x)  +  rye]  J  or,  f  =  — Li ^ — • 

The  reason  for  this  is  well  worth  puzzling  out.     Differentiate  with 
respect  to  (pjT)  and  x ;  divide  by  ~b  T ;  and 


Substitute  this  result  in  equation  (3),  and  we  obtain 

By  Guldberg  and  Waage's  statement  of  the  mass  law,  when  nx 
molecules  of  the  one  system  react  with  n2  molecules  of  the  other, 


©M1^)' 


Hence,  taking  logarithms, 

log  K  +  (n2  -  n^)  log  v  =  n2  log  x  -  nx  log  (1  -  x). 
Differentiate  this  last  expression  with  respect  to  T,  at  constant 
volume ;  and  with  respect  to  v,  at  constant  temperature, 

aiogZ 


Gx\  n2  -  nx  f^x\ 

v\x  +  1  -x) 


58  + 


Introduce  these  values  in  (4)  and  reduce  the  result  to  its  simplest 
terms,  thus, 

DT      "  BT*  •         •         V  .     W 

This  fundamental  relation  expresses  the  change  of  the  equilibrium 
constant  K  with  temperature  at  constant  volume  in  terms  of  the 
molecular  heat  of  the  reaction. 


264 


HIGHER  MATHEMATICS. 


§90. 


Equation  (5),  first  deduced  by  van't  Hoff,  has  led  to  some  of 
the  most  important  results  of  physical  chemistry.  Since  B  and 
T  are  positive,  K  and  q  must  always  have  the  same  sign.  Hence 
van't  Hoffs  principle  of  mobile  equilibrium  follows  directly,  viz., 
If  the  reaction  absorbs  heat,  it  advances  with  rise  of  temperature  ; 
if  the  reaction  evolves  heat  it  retrogrades  with  rise  of  temperature  ; 
and  if  the  reaction  neither  absorbs  nor  evolves  heat,  the  state  of 
equilibrium  is  stationary  with  rise  of  temperature. 

According  to  the  particular  nature  of  the  systems  considered  q 
may  represent  the  so-called  heat  of  sublimation,  heat  of  vaporiza- 
tion, heat  of  solution,  heat  of  dissociation,  or  the  thermal  value  of 
strictly  chemical  reactions  when  certain  simple  modifications  are 
made  in  the  interpretation  of  the  "  concentration  "  K.  If,  at  tem- 
perature Tj  and  T2)  K  becomes  Kx  and  E2,  we  get,  by  the  integration 
of  (5), 


log 


*i 


(1  _  L\ 


(6) 


The  thermal  values  of  the  different  molecular  changes,  calculated 
by  means  of  this  equation,  are  in  close  agreement  with  experiment. 
For  instance : 


Heat  of 

2  in  calories. 

Calculated. 

Observed. 

Vaporization  of  water 
Solution  of  benzoic  acid  in  water 
Sublimation  of  NH4SH      . 
Combination  of  BaGIa  -f  2H20  . 
Dissociation  of  N204   .... 
Precipitation  of  AgGl 

10100 
6700 

21550 
3815 

12900 

15992 

10296 
6500 

21640 
3830 

12500 

15850 

A  sufficiently  varied  assortment  to  show  the  profound  nature  of 
the  relation  symbolized  by  equations  (5)  and  (6). 

Numerical  Example. — Calculate  the  heat  of  solution  of  mercuric  chloride 
from  the  change  of  solubility  with  change  of  temperature.  If  Cj,  c2  denote  the 
solubilities  corresponding  to  the  respective  absolute  temperatures  Tx  and  T2, 

%  =  6-57  when  Tx  =  273°  +  10°  ;  ^  =  11*84  when  T2  =  273°  +  50°. 

Since  the  solubility  of  a  salt  in  a  given  solvent  is  constant  at  any  fixed  tem- 
perature, we  may  write  c  in  place  of  the  equilibrium  constant  K.  From  (6), 
therefore, 


§  90.  THE  INTEGRAL  CALCULUS.  265 

8C!      2V27!       TJ'  '        8  6-57        2\283      323/ 
.*.  q  =  log  1-8  x  45,704*5  =  2,700  (nearly) ;  q  (observed)  =  3,000  (nearly). 
Use  the  Table  of  Natural  Logarithms,  Appendix  II.,  for  the  calculation. 

Le  Chatelier  has  extended  van't  Hoff's  law  and  enunciated  the 
important  generalization :  "  any  change  in  the  factors  of  equilibrium 
from  outside,  is  followed  by  a  reversed  change  within  the  system  ". 
This  rule,  known  as  "  Le  Chatelier 's  theorem,"  enables  the  chemist 
to  foresee  the  influence  of  pressure  and  other  agents  on  physical 
and  chemical  equilibria. 


CHAPTER  V. 

INFINITE  SERIES  AND  THEIR  USES. 

"In  abstract  mathematical  theorems,  the  approximation  to  truth  is 
perfect. ...  In  physical  science,  on  the  contrary,  we  treat  of 
the  least  quantities  which  are  perceptible." — W.  Stanley  Jevons. 

§  91.    What  is  an  Infinite  Series  ? 

Mark  off  a  distance  AB  (Fig.  122)  of  unit  length.  Bisect  AB  at 
Oj ;  bisect  0YB  at  02 ;  0%B  at  03  ;  etc. 

L_ ! I  ill 

A  0,  02       03  04  B 

Fig.  122. 

By  continuing  this  operation,  we  can  approach  as  near  to  B  as  we 
please.  In  other  words,  if  we  take  a  sufficient  number  of  terms 
of  the  series, 

A01  +  0X02  +  020z  + . . ., 
we  shall  obtain  a  result  differing  from  AB  by  as  small  a  quantity 
as  ever  we  please.     This  is  the  geometrical  meaning  of  the  infinite 
series  of  terms, 

1-1+  (i)2  +  (if  +  (i)4  + ..-  to  infinity.  .  (1) 

Such  an  expression,  in  which  the  successive  terms  are  related 
according  to  a  known  law,  is  called  a  series. 

Example. — I  may  now  be  pardoned  if  I  recite  the  old  fable  of  Achilles 
and  the  tortoise.  Achilles  goes  ten  times  as  fast  as  the  tortoise  and  the  latter 
has  ten  feet  start.  When  Achilles  has  gone  ten  feet  the  tortoise  is  one  foot 
in  front  of  him ;  when  Achilles  has  gone  one  foot  farther  the  tortoise  is  ^  ft. 
in  front ;  when  Achilles  has  gone  j-^  ft.  farther  the  tortoise  is  ^  ft.  in  front ; 
and  so  on  without  end ;  therefore  Achilles  will  never  catch  the  tortoise.  There 
is  a  fallacy  somewhere  of  course,  but  where  ? 

When  the  sum  of  an  infinite  series  approaches  closer  and  closer 
to  some  definite  finite  value,  as  the  number  of  terms  is  increased 


§  91.  INFINITE  SERIES  AND  THEIR  USES.  267 

without  limit,  the  series  is  said  to  be  a  convergent  series.  The 
sum  of  a  convergent  series  is  the  "  limiting  value  "  of  §  6.  On 
the  contrary,  if  the  sum  of  an  infinite  series  obtained  by  taking  a 
sufficient  number  of  terms  can  be  made  greater  than  any  finite 
quantity,  however  large,  the  series  is  said  to  be  a  divergent  series. 
For  example, 

1  +  2  +  3+4+.. .to  infinity.  .  .  (2) 

Divergent  series  are  not  much  used  in  physical  work,  while  con- 
verging series  are  very  frequently  employed.1 

The  student  should  be  able  to  discriminate  between  convergent 
and  divergent  series.  I  shall  give  tests  very  shortly.  To  simplify 
matters,  it  may  be  assumed  that  the  series  discussed  in  this  work 
satisfy  the  tests  of  convergency.  It  is  necessary  to  bear  this  in 
mind,  otherwise  we  may  be  led  to  absurd  conclusions.  E.  W.  Hob- 
son's  On  the  Infinite  and  Infinitesimal  in  Mathematical  Analysis, 
London,  1902,  is  an  interesting  pamphlet  to  read  at  this  stage  of 
our  work. 

Let  S  denote  the  limiting  value  or  sum  of  the  converging 
series, 

S  =  a  +  ar  +  ar2  +  . .  .  +  arn  +  arn+1  +  ...  ad  inf.  (3) 

When  r  is  less  than  unity,  cut  off  the  series  at  some  assigned  term, 
say  the  nth,  i.e.,  all  terms  after  arn~l  are  suppressed.  Let  sn  de- 
note the  sum  of  the  n  terms  retained,  o-n  the  sum  of  the  suppressed 
terms.     Then, 

sn  =  a  +  ar  +  ar2  +  . . .  +  arn_1.       .         .         (4) 
Multiply  through  by  r, 

rsn  =  ar  +  ar2  +  ar3  +  . . .  +  ar11. 
Subtract  the  last  expression  from  (4), 

$J1  -  r)  -  a(l  -  t~)  ;  or,  sn  =  aj^-  .         (5) 

Obviously  we  can  write  series  (3),  in  the  form, 

S  =  s„  +  o-„ (6) 

The  error  which  results  when  the  first  n  terms  are  taken  to  repre- 
sent the  series,  is  given  by  the  expression 

o-n  =  S  -  sn. 
This  error  can  be  made  to  vanish  by  taking  an  infinitely  great 

1  A  prize  was  offered  in  France  some  time  back  for  the  best  essay  on  the  use  of 
diverging  series  in  physical  mathematics. 


268  HIGHER  MATHEMATICS.  §  9L 

numbei  of  terms,  or,  LtM=00orn  =  0.     But, 

1  -  rn  a  arn 


?w  =  a. 


1-r      1-r      1-r 
When  n  is  made  infinitely  great,  the  last  term  vanishes, 

T,         arn 
.-.  LtM=oc5— —  =  0. 

The  sum  of  the  infinite  series  of  terms  (3),  is,  therefore,  given  by 
the  expression 

s  =  ^b (7) 

Series  (3)  is  generally  called  a  geometrical  series.    If  r  is 

either  equal  to  or  greater  than  unity,  S  is  infinitely  great  when 

n  =  od,  the  series  is  then  divergent. 

To  determine  the  magnitude  of  the  error  introduced  when  only 

a  finite  number  of  terms  of  an  infinite  series  is  taken.     Take  the 

infinite  number  of  terms, 

S  =  YZTi  =  1  +  r  +  r2  +  . . .  +  r""1  +  j~y      •         (8) 

The  error  introduced  into  the  sum  S,  by  the  omission  of  all  terms 
after  the  wth,  is,  therefore, 

*-& « 

When  r  is  positive,  <rn  is  positive,  and  the  result  is  a  little  too 
small ;  but  if  r  is  negative 

<v-±r^7      '•     •     •     (10) 

which  means  that  if  all  terms  after  the  wth  are  omitted,  the  sum 
obtained  will  be  too  great  or  too  small,  according  as  n  is  odd  or 
even. 

Examples. — (1)  Suppose  that  the  electrical  conductivity  of  an  organic 
acid  at  different  concentrations  has  to  be  measured  and  that  the  first 
measurement  is  made  on  50  c.c.  of  solution  of  concentration  c.  25  c.c.  of 
this  solution  are  then  removed  and  25  c.c.  of  distilled  water  added  instead. 
This  is  repeated  five  more  times.  What  is  the  then  concentration  of  the  acid 
in  the  electrolytic  cell  ?  Obviously  we  are  required  to  find  the  7th  term  in  the 
series  c{l  +  £  +  (J)2  +  (J)3  +...},  where  the  nth  term  is  c^)**-1.     Ansr.  (£)6c. 

(2)  If  the  receiver  of  an  air  pump  and  connections  have  a  volume  a,  and 
the  cylinder  with  the  piston  at  the  top  has  a  volume  b,  the  first  stroke  of 
the  pump  will  remove  b  of  the  air.  Hence  show  that  the  density  of  the  air 
in  the  receiver  after  the  third  stroke  will  be  0*75  of  its  original  density  if 
a  =  1000  c.c,  and  6  =  100  c.c.     Hint.  After  the  first  stroke  the  density  of  the 


§  92.  INFINITE  SERIES  AND  THEIR  USES.  269 

air,  pv  will  be  px(a  +  b)  =  pQa,  when  p0  denotes  the  original  density  of  the  air; 
after  the  second  stroke  the  density  will  be  p.2,  and  p.2(a+b)=pla;  .*.  p2(a+b)2=p0a.'~ 
After  the  nth  stroke,  pn(a+  b)n=p0an ;  or  pn(1000  +  100)3=10003 ;  .-.  />„=0-73. 

§  92.   Washing  Precipitates. 

Applications  of  the  series  to  the  washing  of  organic  substances 
with  ether  ;  to  the  washing  of  precipitates  ;  to  Mallet's  process  for 
separating  oxygen  from  air  by  shaking  air  with  water,  etc.,  are 
obvious.  We  can  imagine  a  precipitate  placed  upon  a  filter  paper, 
and  suppose  that  C0  represents  the  concentration  of  the  mother 
liquid  which  is  to  be  washed  from  the  precipitate ;  let  v  denote  the 
volume  of  the  liquid  which  remains  behind  after  the  precipitate  has 
drained ;  vx  the  volume  of  liquid  poured  on  to  the  precipitate  in  the 
filter  paper. 

Examples. — (1)  A  precipitate  at  the  bottom  of  a  beaker  which  holds  v  c.c. 
of  mother  liquid  is  to  be  washed  by  decantation,  i.e.,  by  repeatedly  filling  the 
beaker  up  to  say  the  vx  c.c.  mark  with  distilled  water  and  emptying.  Sup- 
pose that  the  precipitate  and  vessel  retain  v  c.c.  of  the  liquid  in  the  beaker  at 
each  decantation,  what  will  be  the  percentage  volume  of  mother  liquor  about 
the  precipitate  after  the  nth  emptying,  assuming  that  the  volume  of  the 
precipitate  is  negligibly  small  ?  Ansr.  lOOivjv^11-1.  Hint.  The  solution  in 
the  beaker,  after  the  first  filling,  has  vjv1  c.c.  of  mother  liquid.  On  emptying, 
v  of  this  vjvx  c.c.  is  retained  by  the  precipitate.  On  refilling,  the  solution  in 
the  beaker  has  (v2/vi)/'yi  °f  motaer  liquor,  and  so  we  build  up  the  series, 

(2)  Show  that  the  residual  liquid  which  remains  with  the  precipitate 
after  the  first,  second  and  nth  washings  is  respectively 

vC,  =  —?—  vC0;  vG2  =  f-JL-VvOo ;  vCn  =  (-^—\vCQ. 

It  is  thus  easy  to  see  that  the  residue  of  mother  liquid  vGn  which 
contaminates  the  precipitate,  as  impurity,  is  smaller  the  less  the 
value  of  v/(v  +  vx) ;  this  fraction,  in  turn,  is  smaller  the  less  the 
value  of  v,  and  the  greater  the  value  of  vv  Hence  it  is  inferred 
that  (i)  the  more  perfectly  the  precipitate  is  allowed  to  drain — 
lessening  v;  and  (ii)  the  greater  the  volume  of  washing  liquid 
employed — increasing  vx — the  more  perfectly  effective  will  be  the 
washing  of  the  precipitate. 

Example. — Show  that  if  the  amount  of  liquid  poured  on  to  the  precipi- 
tate at  each  washing  is  nine  times  the  amount  of  residual  liquid  retained  by 
the  precipitate  on  the  filter  paper,  then,  if  the  amount  of  impurity  con- 
taminating the  original  precipitate  be  one  gram,  show  that  0*0001  gram  of 
impurity  will  remain  after  the  fourth  washing. 


270  HIGHER  MATHEMATICS.  §  92. 

What  simplifying  assumptions  have  been  made  in  this  discussion? 
We  have  assumed  that  the  impurity  on  the  filter  paper  is  reduced 
a  i^th  part  when  vx  volumes  of  the  washing  liquid  is  poured  on  to 
the  precipitate,  and  the  latter  is  allowed  to  drain.  We  have  ne- 
glected the  amount  of  impurity  which  adheres  very  tenaciously,  by 
surface  condensation  or  absorption.  The  washing  is,  in  conse- 
quence, less  thorough  than  the  simplified  theory  would  lead  us  to 
suppose.  Here  is  a  field  for  investigation.  Can  we  make  the 
plausible  assumption  that  the  amount  of  impurity  absorbed  is  pro- 
portional to  the  concentration  of  the  solution?  Let  us  find  how 
this  would  affect  the  amount  of  impurity  contaminating  the  pre- 
cipitate after  the  nth  washing. 

Let  a  denote  the  amount  of  solution  retained  as  impurity  by 
surface  condensation,  let  b  denote  the  concentration  of  the  solution. 
If  we  make  the  above-mentioned  assumption,  then 

b  =  ka, 

where  k  is  the  constant  of  proportion.  Let  v  c.c.  of  washing  liquid 
be  added  to  the  precipitate  which  has  absorbed  a  c.c.  of  mother 
liquid.  Then  a0  -  ax  c.c.  of  impurity  passes  into  solution,  and  with 
the  v  c.c.  of  solvent  gives  a  solution  of  concentration  (a0  -  a-^/v ; 
the  amount  of  impurity  remaining  with  the  precipitate  will  be 

*=p-*i      .     .     .     (id 

When  this  solution  has  drained  off,  and  v  more  c.c.  of  washing 
liquid  is  added,  the  amount  of  impurity  remaining  with  the  pre- 
cipitate will  be 

^-^2  =  K-  •         •         •         (12) 

Eliminate  aY  from  this  by  the  aid  of  (11),  and  we  get 


2      kv  +  1 
for  the  second  washing ;  and  for  the  nth.  washing, 

1 

an  =    7 Ta0* 

kv  +  1  ° 

But  all  this  is  based  upon  the  unverified  assumption  as  to  the  con- 
stancy of  k,  a  question  which  can  only  be  decided  by  an  appeal  to 
experiment.     See  E.  Bunsen,  Liebig's  Ann.,  138,  269,  1868. 


§  93.  INFINITE  SERIES  AND  THEIR  USES.  271 

§  93.  Tests  for  Convergent  Series. 

Mathematicians  have  discovered  some  very  interesting  facts  in 
their  investigations  upon  the  properties  of  infinite  series.  Many  of 
these  results  can  be  employed  as  tests  for  the  convergency  of  any 
given  series.  I  shall  not  give  more  than  three  tests  to  be  used  in 
this  connection. 

I.  If  the  series  of  terms  are  alternately  positive  and  negative, 
and  the  numerical  value  of  the  successive  terms  decreases,  the  series 
is  convergent.     For  example,  the  series 

1      1*1      t  +  t      ,M 
may  be  expressed  in  either  of  the  following  forms : — 

a-i)+(W)+(W)+..-;  i-(W)-(w)-(£-f)--- 

Every  quantity  within  the  brackets  is  of  necessity  positive.  The 
sum  of  the  former  series  is  greater  than  1  -  J,  and  the  sum  of  the 
latter  is  less  than  1 ;  consequently,  the  sum  of  the  series  must  have 
some  value  between  1  and  J.  In  other  words,  the  series  is  con- 
vergent. If  a  series  in  which  all  the  terms  are  positive  is  conver- 
gent, the  series  will  also  be  convergent  when  some  or  all  of  the 
terms  have  a  negative  value.  Otherwise  expressed,  a  series  with 
varying  signs  is  convergent  if  the  series  derived  from  it  by  making 
all  the  signs  positive  is  convergent. 

II.  If  there  be  two  infinite  series. 

u0  +  ui  +  U2  +  . .  .  Un  +  .  .  .  ;   and  V0  +  V±  +  V2  +  .  .  .  Vn  +  . . . 

the  first  of  which  is  known  to  be  convergent,  and  if  each  term  of 
the  other  series  is  not  greater  than  the  corresponding  term  of  the 
first  series,  the  second  series  is  also  convergent.  If  the  first  series  is 
divergent,  and  each  term  of  the  second  series  is  greater  than  the 
corresponding  term  of  the  first  series,  the  second  series  is  divergent. 
This  is  called  the  comparison  test.  The  series  most  used  for 
reference  are  the  geometrical  series 

a  +  ar  +  ar2  +  . . .  +  arn  +  . . . 
which  is  known  to  be  convergent  when  r  is  less  than  unity,  and 
divergent  when  r  is  greater  than  or  equal  to  unity ;  and 

.        I      11 

2m  +  Sm  +  4™  +  * " 
which  is  known  to  be  convergent  when  m  is  greater  than  unity ; 
and  divergent,  if  m  is  equal  to  or  greater  than  unity. 


272  HIGHER  MATHEMATICS.  §  93. 

Example. — Show  that  the  series  l  +  ^-  +  ^  +  ^+...is  convergent  by 
comparison  with  the  geometrical  series  1  +  ^  +  tV  +  ^  +  . . . 

III.  An  infinite  series  is  convergent  if  from  and  after  some 
fixed  term  the  ratio  of  each  term  to  the  preceding  term  is  numerically 
less  than  some  quantity  which  is  itself  less  than  unity.  For  in- 
stance, let  the  series,  beginning  from  the  fixed  term,  be 

a1  +  a2  +  az  +  ... 
Let  sn  denote  the  sum  of  the  first  n  terms.     We  can  therefore 
write 

sn  =  a1  +  a2  +  a3  +  a4  +  . . .     • 

By  rearranging  the  terms  of  the  series,  we  get 

(-       a9      a,  a0      a,  a~  \ 

a1      a2  al      a3  a1  J 

The  fraction  -~*  is  called  the  ratio  test.     Suppose  the  ratio  test 

a2  a3  a. 

—  be  less  than  r  ',    —  be  less  than  r  \    —  be  less  than    r  \    ... 

aY  a2  a3 

that  is,  from  (3)  and  (5),  page  267, 

1  -  rn 

Sn  be  less  than  a,-, • 

1 1  —  r 
Hence,  from  (7),  if  r  is  less  than  unity, 

ai 

Sn  be  less  than  ^ -• 

»  1  -  r 

Thus  the  sum  of  as  many  terms  as  we  please,  beginning  with  a, 
is  less  than  a  certain  finite  quantity  r,  and  therefore  the  series 
beginning  with  ax  is  convergent. 

Examples. — (1)  The  series  1  +  \x  +  -^x2  +  f,sc3  +  . . .,  is  convergent  be- 
cause the  test-ratio  =  x/n  becomes  zero  when  n  =  ao. 

(2)  The  series  1  +  $x  +  \  -fa2  +  \  -\  -fa3  +  . . .  is  convergent  when  x  is 
less  than  unity. 

It  is  possible  to  have -a  series  in  which  the  terms  increase  up 
to  a  certain  point,  and  then  begin  to  decrease.     In  the  series 

1  +  2x  +  3x2  +  4ic3  +  . . .  +  nxn+1  +  . . ., 
for  example,  we  have 


an  nx        /.,  1    \ 

— —  = T  =   1  + Ax. 

an_1      n  -  1      V         n-lj 


If  n  be  large  enough,  the  series  can  be  made  as  nearly  equal  to  x 
as  we  please.  Hence,  if  re  is  less  than  unity,  the  series  is  con- 
vergent.    The  ratio  will  not  be  less  than  unity  until 


§  94.  INFINITE  SEKIES  AND  THEIR  USES.  273 

»-  1  ,  1 

be  less  than   1  '.    i.e.,    until  71  >  3 • 

nx  '       '  1  -  x 

9  1 

If  x  =  jtx,  for  example,  j-t —  =  10,  and  the  terms  only  begin  to 

decrease  after  the  10th  term. 

These  tests  will  probably  be  found  sufficient  for  all  the  series 
the  student  is  likely  to  meet  in  the  ordinary  course  of  things.  If 
the  test-ratio  is  greater  than  unity,  the  series  is  divergent ;  and  if 
this  ratio  is  equal  to  unity,  the  test  fails. 

§  9$.    Approximate  Calculations  in  Scientific  Work. 

A  good  deal  of  the  tedious  labour  involved  in  the  reduction  of 
experimental  results  to  their  final  form,  may  be  avoided  by  at- 
tention to  the  degree  of  accuracy  of  the  measurements  under 
consideration.  It  is  one  of  the  commonest  of  mistakes  to  extend 
the  arithmetical  work  beyond  the  degree  of  precision  attained  in 
the  practical  work.  Thus,  Dulong  calculated  his  indices  of  refrac- 
tion to  eight  digits  when  they  agreed  only  to  three.  When  asked 
"  Why?"  Dulong  returned  the  ironical  answer:  "  I  see  no  reason 
for  suppressing  the  last  decimals,  for,  if  the  first  are  wrong,  the  last 
may  be  all  right"  1 

In  a  memoir  "  On  the  Atomic  Weight  of  Aluminium,"  at 
present   before  me,  I  read,   "  0*646  grm.   of  aluminium  chloride 

gave  2-0549731  grms.  of  silver  chloride "    It  is  not  clear  how  the 

author  obtained  his  seven  decimals  seeing  that,  in  an  earlier  part 
of  the  paper,  he  expressly  states  that  bis  balance  was  not  sensitive 
to  more  than  0*0001  grm.  A  popular  book  on  '-The  Analysis  of 
Gases,"  tells  us  that  1  c.c.  of  carbon  dioxide  weighs  0*00196633 
grm.  The  number  is  calculated  upon  the  assumption  that  carbon 
dioxide  is  an  ideal  gas,  whereas  this  gas  is  a  notorious  exception. 
Latitude  also  might  cause  variations  over  a  range  of  +  0*000003 
grm.  The  last  three  figures  of  the  given  constant  are  useless. 
"  Superfluitas,"  said  R.  Bacon,  "  impedit  multum . . .  reddit  opus 
abominabile." 

Although  the  measurements  of  a  Stas,  or  of  a  Whitworth,  may 
require  six  or  eight  decimal  figures,  few  observations  are  correct 
to  more  than  four  or  five.  But  even  this  degree  of  accuracy  is 
only  obtained  by  picked  men  working  under  special  conditions. 
Observations  which  agree  to  the  second  or  third  decimal  place  are 
comparatively  rare  in  chemistry. 

S 


274  HIGHER  MATHEMATICS.  §  94. 

Again,  the  best  of  calculations  is  a  more  or  less  crude  approxi- 
mation on  account  of  the  "simplifying  assumptions"  introduced 
when  deducing  the  formula  to  which  the  experimental  results  are 
referred.  It  is,  therefore,  no  good  extending  the  "calculated 
results  "  beyond  the  reach  of  experimental  verification.  "It  is 
unprofitable  to  demand  a  greater  degree  of  precision  from  the 
calculated  than  from  the  observed  results — but  one  ought  not  to 
demand  a  less  "  (H.  Poincare's  Mdcanique  C&leste,  Paris,  1892). 

The  general  rule  in  scientific  calculations  is  to  use  one  more 
decimal  figure  than  the  degree  of  accuracy  of  the  data.  In  other 
words,  reject  as  superfluous  all  decimal  figures  beyond  the  first 
doubtful  digit.  The  remaining  digits  are  said  to  be  significant 
figures. 

Examples. — In  1/540,  there  are  four  significant  figures,  the  cypher  indi- 
cates that  the  magnitude  has  been  measured  to  the  thousandth  part ;  in 
0-00154,  there  are  three  significant  figures,  the  cyphers  are  added  to  fix  the 
decimal  point ;  in  15,400,  there  is  nothing  to  show  whether  the  last  two 
cyphers  are  significant  or  not,  there  may  be  three,  four,  or  five  significant 
figures. 

In  "casting  off"  useless  decimal  figures,  the  last  digit  retained 
must  be  increased  by  unity  when  the  following  digit  is  greater 
than  four.  We  must,  therefore,  distinguish  between  9*2  when  it 
means  exactly  9*2,  and  when  it  means  anything  between  9  14  and 
9"25.  In  the  so-called  "exact  sciences,"  the  latter  is  the  usual 
interpretation.  Quantities  are  assumed  to  be  equal  when  the 
differences  fall  within  the  limits  of  experimental  error. 

Logarithms. — There  are  very  few  calculations  in  practical 
work  outside  the  range  of  four  or  five  figure  logarithms.  The 
use  of  more  elaborate  tables  may,  therefore,  be  dispensed  with. 
There  are  so  very  many  booklets  and  cards  containing,  "  Tables  of 
Logarithms "  upon  the  market  that  one  cannot  be  recommended 
in  preference  to  another. 

Addition  and  Subtraction. — In  adding  such  numbers  as  9-2 
and  0*4913,  cast  off  the  3  and  the  1,  then  write  the  answer,  9*69, 
not  9-6913.  Show  that  5*60  +  20*7  +  103-193  =  129*5,  with  an 
error  of  about  0*01,  that  is  about  0*08  per  cent. 

Multiplication  and  Division. — The  product  2*25tt  represents 
the  length  of  the  perimeter  of  a  circle  whose  diameter  is  2*25 
units ;  7r  is  a  numerical  coefficient  whose  value  has  been  calculated 
by  Shanks  {Proc.  Roy.  Soc,  22,  45,  1873),  to  over  seven  hundred 


§  94.  INFINITE  SERIES  AND  THEIE  USES.  275 

decimal  places,  so  that  tt  -  3-141592,653589,793. ...  Of  these  two 
numbers,  therefore,  2*25  is  the  less  reliable.  Instead  of  the 
ludicrous  7-0685808625 . . .,  we  simply  write  the  answer,  7*07. 
Again,  although  W.  K.  Oolvill  has  run  out  J2  to  110  decimal 
places  we  are  not  likely  to  want  more  than  half  a  dozen  significant 
figures. 

It  is  no  doubt  unnecessary  to  remind  the  reader  that  in  scientific 
computations  the  standard  arithmetical  methods  of  multiplication 
and  division  are  abbreviated  so  as  to  avoid  writing  down  a  greater 
number  of  digits  than  is  necessary  to  obtain  the  desired  degree  of 
accuracy.  The  following  scheme  for  "  shortened  multiplication 
and  division,"  requires  little  or  no  explanation  : — 

Shortened  Multiplication.  Shortened  Division. 

9-774  365-4)3571-3(9-774 

3288-6 


365-4 

2932-2 

586-4 
48-9 

3-9 

282-7 
255-8 

26-9 
25-5 


3571-4 


1-4 

The  digits  of  the  multiplier  are  taken  from  left  to  right,  not 
right  to  left.  One  figure  less  of  the  divisor  is  used  at  each  step  of 
the  division.  The  last  figure  of  the  quotient  is  obtained  mentally. 
A  "bar"  is  usually  placed  over  strengthened  figures  so  as  to  allow 
for  an  excess  or  defect  of  them  in  the  result. 

W.  Ostwald,  in  his  Hand-  und  Hilfsbuch  zur  Ausfiihrung 
physikochemiker  Messungen,  Leipzig,  1893,  has  said  that  "the 
use  of  these  methods  cannot  be  too  strongly  emphasized.  The 
ordinary  methods  of  multiplication  and  division  must  be  termed 
unscientific."  Full  details  are  given  in  E.  M.  Langley's  booklet, 
A  Treatise  on  Computation,  London,  1895. 

The  error  introduced  in  approximate  calculations  by  the  "  casting 
off"  of  decimal  figures. 
Some  care  is  required  in  rounding  off  decimals  to  avoid  an 
excess  or  defect  of  strengthened  figures  by  making  the  positive 
and  negative  errors  neutralize  each  other  in  the  final  result.  It  is 
sometimes  advisable,  in  dealing  with  the  5  in  a  M  train"  of  arith- 
metical operations,  to  leave  the  last  figure  an  even  number.  E.g., 
3*75  would  become  3'8,  while  3-85  would  be  written  3-8. 

S* 


276  HIGHER  MATHEMATICS.  §  95. 

The  percentage  error  of  the  product  of  two  approximate  numbers 
is  very  nearly  the  algebraic  sum  of  the  percentage  error  of  each. 
If  the  positive  error  in  the  one  be  numerically  equal  to  the  negative 
error  in  the  other,  the  product  will  be  nearly  correct,  the  errors 
neutralize  each  other. 

Example. — 19*8  x  3-18.  The  first  factor  may  be  written  20  with  a  + 
error  of  1  %,  and,  therefore,  20  x  3-18  =  63-6,  with  a  +  error  of  1  %.  This 
excess  must  be  deducted  from  63*6.  We  thus  obtain  62*95.  The  true  result 
is  62-964. 

The  percentage  error  of  the  quotient  of  two  approximate  numbers 
is  obtained  by  subtracting  the  percentage  error  of  the  numerator 
from  that  of  the  denominator.  If  the  positive  error  of  the  numer- 
ator is  numerically  equal  to  the  positive  error  of  the  denominator, 
the  error  in  the  quotient  is  practically  neutralized. 

There  is  a  well-defined  distinction  between  the  approximate 
values  of  a  physical  constant,  which  are  seldom  known  to  more 
than  three  or  four  significant  figures,  an^the  approximate  value  of 
the  incommensurables  7r,  e,  J2, . . .  which  can  be  calculated  to  any 
desired  degree  of  accuracy.  If  we  use  -^  in  place  of  3  -1416  for  ir, 
the  absolute  error  is  greater  than  or  equal  to  3-14-26  -  3*1416,  and 
equal  to  or  less  than  3*1428  -  3*1416  ;  that  is,  between  -0012  and 
•0014.  In  scientific  work  we  are  rarely  concerned  with  absolute 
errors. 

§  95.  Approximate  Calculations  by  Means  of  Infinite  Series. 

The  reader  will,  perhaps,  have  been  impressed  with  the  fre- 
quency with  which  experimental  results  are  referred  to  a  series 
formula  of  the  type : 

y  =  A  +  Bx  +  Cx2  +  Dx*  + . . .,  .  .  (1) 
in  physical  or  chemical  text-books.  For  instance,  I  have  counted 
over  thirty  examples  in  the  first  volume  of  Mendeleeff's  The 
Principles  of  Chemistry,  and  in  J.  W.  Mellor's  Chemical  Statics 
and  Dynamics  it  is  shown  that  all  the  formulae  which  have  been 
proposed  to  represent  the  relation  between  the  temperature  and 
the  velocity  of  chemical  reactions  have  been  derived  from  a  similar 
formula  by  the  suppression  of  certain  terms.  The  formula  has  no 
theoretical  significance  whatever.  It  does  not  pretend  to  accurately 
represent  the  whole  course  of  any  natural  phenomena.  All  it 
postulates  is  that  the  phenomena  in  question  proceed  continuously. 
In  the  absence  of  any  knowledge  as  to  the  proper  setting  of  the 


§95. 


INFINITE  SEKIES  AND  THEIR  USES. 


277 


"law"  connecting  two  variables,  this  formula  may  be  used  to 
express  the  relation  between  the  two  phenomena  to  any  required 
degree  of  approximation.  It  is  only  to  be  looked  upon  as  an 
arbitrary  device  which  is  used  for  calculating  corresponding  values 
of  the  two  variables  where  direct  measurements  have  not  been 
obtained.  A,  B,  C, . . .  are  constants  to  be  determined  from  the  ex- 
perimental data  by  methods  to  be  described  later  on.  There  are 
several  interesting  features  about  this  expression.- 

I.  When  the  progress  of  any  physical  change  is  represented  by 
the  above  formula,  the  approximation  is  closer  to  reality  the  greater 
the  number  of  terms  included  in  the  calculation.  This  is  best  shown 
by  an  example.  The  specific  gravity  s  of  an  aqueous  solution  of 
hydrogen  chloride  is  an  unknown  function  of  the  amount  of  gas  p 
per  cent,  dissolved  in  the  water.     (Unit :  water  at  4°  =  10,000.) 

The  first  two  columns  of  the  following  table  represent  cor- 
responding values  of  p  and  s,  determined  by  Mendeleeff.  It  is 
desired  to  find  a  mathematical  formula  to  represent  these  results 
with  a  fair  degree  of  approximation,  in  order  that  we  may  be  able 
to  calculate  p  if  we  know  s,  or,  to  determine  s  if  we  know  p.  Let 
us  suppress  all  but  the  first  two  terms  of  the  above  series, 

s  =  A  +  Bp, 
where  A  and  B  are  constants,  found,  by  methods  to  be  described 
later,  to  be  A  =  9991-6,   B  =  50*5.      Now  calculate  s  from  the 
given  values  of  p  by  means  of  the  formula, 

s  =  9991-6  +  50-5^,    ....        (2) 

and  compare  the  results  with  those  determined  by  experiment. 
See  the  second  and  third  columns  of  the  following  table : — 


Percentage 
Composition 

Specific  Gravity  s. 

Found. 

Calculated. 

1st  Approx. 

2nd  Approx. 

5 
10 
15 
20 
25 

10242 
10490 
10744 
11001 
11266 

10244 

10497 
10749 
13002 
11254 

10240 
10492 
10746 
11003 
11263 

Formula  (2),  therefore,  might  serve  all  that  is  required  in,  say, 


278  HIGHER  MATHEMATICS.  §  95. 

a  manufacturing  establishment,  but,  in  order  to  represent  the  con- 
nection between  specific  gravity  and  percentage  composition  with 
a  greater  degree  of  accuracy,  another  term  must  be  included  in 
the  calculation,  thus  we  write 

s  =  A+  Bp  +  Cp2, 
where  B  is  found  to  be  equivalent  to  49-43,  and  G  to  0-0571.  The 
agreement  between  the  results  calculated  according  to  the  formula  : 
s  -  9991-6  +  49-43^  +  0-0571^,  .  .  (3) 
and  those  actually  found  by  experiment  is  now  very  close.  This 
will  be  evident  on  comparing  the  second  with  the  fourth  columns 
of  the  above  table.  The  term  0*0571p2  is  to  be  looked  upon  as  a 
correction  term.  It  is  very  small  in  comparison  with  the  preced- 
ing terms. 

If  a  still  greater  precision  is  required,  another  correction  term 
must  be  included  in  the  calculation,  we  thus  obtain 

y  =  A  +  Bx  +  Gx2  +  Dx*. 
Such  an  expression  was  employed  by  T.  E.  Thorpe  and  A.  W. 
Eiicker  (Phil.  Trans.,  166,  ii.,  1,  1877)  for  the  relation  between  the 
volume  and  temperature  of  sea-water ;  by  T.  E.  Thorpe  and  A.  E. 
Tutton  (Journ.  Chem.  Soc,  57,  545,  1890)  for  the  relation  between 
the  temperature  and  volume  of  phosphorous  oxide  ;  and  by  Eapp 
for  the  specific  heat  of  water,  <r,  between  0°  and  100°.  Thus  Eapp 
gives 

<r  -  1-039935  -  0-0070680  +  O-OOO2125502  -  0-00000154^, 
and  Hirn  (Ann.  Ghim.  Phys.  [4],  10,  32,  1867)  used  yet  a  fourth 
term,  namely,  f 

v  =  A  +  BB  +  G62  +  D0Z  +  EOS 
in  his  formula  for  the  volume  of  water,  between  100°  and  200°. 

The  logical  consequence  of  this  reasoning,  is  that  by  including 
every  possible  term  in  the  approximation  formula,  we  should  get 
absolutely  correct  results  by  means  of  the  infinite  converging 
series : 

y  =  A  +  Bx  +  Cx2  +  Dxz  +  Ex*  +  Fx*  +  . . .  +  ad  infin.  (4) 
It  is  the  purpose  of  Maclaurin's  theorem  to  determine  values  of 
A,  B,  G, .  .  .  which  will  make  this  series  true. 

II.  The  rapidity  of  the  convergence  of  any  series  determines 
how  many  terms  are  to  be  included  in  the  calculation  in  order  to 
obtain  any  desired  degree  of  approximation.     It  is  obvious  that 


§95.  INFINITE  SERIES  AND  THEIR  USES.  279 

the  smaller  the  numerical  value  of  the  "correction  terms"  in  the 
preceding  series,  the  less  their  influence  on  the  calculated  result. 
If  each  correction  term  is  very  small  in  comparison  with  the 
preceding  one,  very  good  approximations  can  be  obtained  by  the 
use  of  comparatively  simple  formula  involving  two,  or,  at  most, 
three  terms,  e.g.,  p.  87.  On  the  other  hand,  if  the  number  of 
correction  terms  is  very  great,  the  series  becomes  so  unmanageable 
as  to  be  p  ;actically  useless. 

Equation  (1)  may  be  written  in  the  form, 
y  =  A(l  +  bx  +  ex2  +  . . .), 
where  A,  b,  c, . . .  are  constants  ;  A  is  the  value  of  y  when  x  =  0. 

As  a  general  rule,  when  a  substance  is  heated,  it  increases  in 
volume,  v  ;  its  mass,  m,  remains  constant,  the  density,  p,  therefore, 
must  necessarily  decrease.     But, 

Mass  =  Density  X  Volume ;  or,  m  =  pV. 
The  volume  of  a  substance  at  0°  is  given  by  the  expression 

v  -  v0{l  +  a$), 
where  vQ  represents  the  volume  of  the  substance  when  0  is  0°  0. 
a  is  the  coefficient  of  cubical  expansion:    Evidently, 

p0         V         VQ(1  +  ad)       ■'*.«,     .  .  Po 

p        V0  V0  y        r        1  +  aO 

True  for  solids,  liquids,  and  gases.  For  simplicity,  put  p0  m  1.  By 
division,  we  obtain 

p  =  1  -  a$  +  (aO)*  -  (a0)3  +  .  .  . 

For  solids  and  some  liquids  a  is  very  small  in  comparison*  with 
unity.  For  example,  with  mercury  a  =  0*00018.  Let  6  be  small 
enough 

p  =  1  -  0-00018(9  +  (O-OOO180)2  -  . . . 
.-.  p  =  1  -0-000180  +  0000000,032402 

If  the  result  is  to  be  accurate  to  the  second  decimal  place  (1  per  100), 
terms  smaller  than  0-01  should  be  neglected  ;  if  to  the  third  decimal 
place  (1  per  1000 j,  omit  all  terms  smaller  than  0  001,  and  so  on. 
It  is,  of  course,  necessary  to  extend  the  calculation  a  few  decimal 
places  beyond  the  required  degree  of  approximation.  How  many, 
naturally  depends  on  the  rapidity  of  convergence  of  the  series. 
If,  therefore,  we  require  the  density  of  mercury  correct  to  the 
sixth  decimal  place,  the  omission  of  the  third  term  can  make  no 
perceptible  difference  to  the  result. 


280  HIGHER  MATHEMATICS.  §  96. 

Examples. — (1)  If  h0  denotes  the  height  of  the  barometer  at  0°  0.  and 
h  its  height  at  6°,  what  terms  must  be  included  in  the  approximation 
formula,  h  =  h0(l  +  0*000160),  in  order  to  reduce  a  reading  at  20°  to  the 
standard  temperature,  correct  to  1  in  100,000  ? 

(2)  In  accurate  weighings  a  correction  must  be  made  for  the  buoyancy  of 
the  air  by  reducing  the  "  observed  weight  in  air  "  to  "  weight  in  vacuo  n.*  Let 
W  denote  the  true  weight  of  the  body  (in  vacuo),  w  the  observed  weight  in 
air,  p  the  density  of  the  body,  px  the  density  of  the  weights,  p2  the  density  of 
the  air  at  the  time  of  weighing.     Hence  show  that  if 

V      PJ       V      ft/  1_£2        V      ft     pJ  \P    ft/ 

p 

which  is  tbe  standard  formula  for  reducing  weighings  in  air  to  weighings  in 
vacuo.  The  numerical  factor  represents  the  density  of  moderately  moist  air 
at  the  temperature  of  a  room  under  normal  conditions. 

(3)  If  a  denotes  the  coefficient  of  cubical  expansion  of  a  solid,  the  volume 
of  a  solid  at  any  temperature  6  is,  v  =  v0(l  +  ad),  where  v0  represents  the 
volume  of  the  substance  at  0°.  Hence  show  that  the  relation  between  the 
volumes,  vx  and  v2,  of  the  solid  at  the  respective  temperatures  of  Q±  and  B2  is 
i?!  =  v2(l  +  adx  -  ad2).     Why  does  this  formula  fail  for  gases  ? 

ia\  ci-  I  1       a       a2 

(4)  Since  =_ +       +       +. 

w  x-axx2x3  ' 

the  reciprocals  of  many  numbers  can  be  very  easily  obtained  correct  to  many 

decimal  places.     Thus 

Jt  =  1^  =  I^  +  Io|oO  +  1,000,000+  •••  =0-01 +  0-0003 +  0-000009+  ... 

(5)  We  require  an  accuracy  of  1  per  1,000.  What  is  the  greatest  value  of 
x  which  will  permit  the  use  of  the  approximation  formula  (1  +  x)z  =  1  +  Sx  ? 
There  is  a  collection  of  approximation  formulae  on  page  601. 

(6)  From  the  formula,  (1  +  x)n  =  1  ±  nx,  where  n  may  be  positive  or 
negative,  integral  or  fractional,  calculate  the  approximate  values  of  \/999, 
1/  v/l^,  (1-001)3,  dvoS,  mentally.  Hints.  In  the  first  case  n  =  J ;  in  the 
second,  n  =  -  J ;  in  the  third,  n  =  3.  In  the  first,  x [.«■  -  1 J  in  the  second, 
x  =  002 ;  in  the  third,  x  =  0-001,  etc. 

§  96.    Maclaurin's  Theorem. 

Maclaurirts  theorem  determines  the  law  for  the  expansion  of  a 
function  of  a  single  variable  in  a  series  of  ascending  powers  of 
that  variable.     Let  the  variable  be  denoted  by  x,  then, 

u  =  f{x). 

1  A  difference  of  45  mm.  in  the  height  of  a  barometer  during  an  organic  combus- 
tion analysis,  may  cause  an  error  of  0*6  °/0  in  the  determination  of  the  C02,  and  an 
error  of  0*4  % in  tne  determination  of  the  H20.  See  W.  Crookes,  "The  Determination 
of  the  Atomic  Weight  of  Thallium,"  Phil.  Trans.,  163,  277,  1874. 


§  96.  INFINITE  SERIES  AND  THEIR  USES.  281 

Assume  that  f(x)  can  be  developed  in  ascending  powers  of  x,  like 
the  series  used  in  the  preceding  section,  namely, 

u  =  f(x)  =  A  +  Bx  +  Cx2  +  Dx3  +  . . .,  .         (1) 

where  A,  B,  G,  D  . . . ,  are  constants  independent  of  x}  but  de- 
pendent on  the  constants  contained  in  the  original  function.  It  is 
required  to  determine  the  value  of  these  constants,  in  order  that 
the  above  assumption  may  be  true  for  all  values  of  x. 

There  are  several  methods  for  the  development  of  functions  in 
series,  depending  on  algebraic,  trigonometrical,  or  other  processes. 
The  one  of  greatest  utility  is  known  as  Taylor's  theorem.  Mac- 
laurin's l  theorem  is  but  a  special  case  of  Taylor's.  We  shall  work 
from  the  special  to  the  general. 

By  successive  differentiation  of  (1), 

du      df(x)       ■       _         " 

<Pu      df'(x) 

3F  =  T=20  +  2-3Da;  +  -";     •       •       (3) 

&u      df"(x)      „  i-  • 

^=^i  =  2-3-I)  + W 

By  hypothesis,  (1)  is  true  whatever  be  the  value  of  x,  and, 
therefore,  the  constants  A,  B,  C,  D, . . .  are  the  same  whatever 
value  be  assigned  to  x.  Now  substitute  x  =  0  in  equations  (2), 
(3),  (4).  Let  v  denote  the  value  assumed  by  u  when  x  =  0. 
Hence,  from  (1), 

i-/(0)-4 

from(2)'    £  =/'(0)  =  i.s, 

,„.     &v  lft     I  .        (5) 

from  (3),    ^=/"(0)  =  1.2C, 

Substitute  the  above  values  of  A,  B,  C, . . .,  in  (1)  and  we  get 
dv  x      d2v  x2       d3v  x3 

1  The  name  is  here  a  historical  misnomer.  Taylor  published  his  series  in  1715. 
In  1717,  Stirling  showed  that  the  series  under  consideration  was  a  special  case  of 
Taylor's.  Twenty-five  years  after  this  Maclaurin  independently  published  Stirling's 
series.  But  then  "both  Maclaurin  and  Stirling,"  adds  De  Morgan,  "would  have 
been  astonished  to  know  that  a  particular  case  of  Taylor's  theorem  would  be  called 
by  either  of  their  names  ". 


A 

=  v; 

B 

dv 
*dx; 

1  d2v  . 

C 

~  2 1  dx2 ' 

1  d3v 

D 

■  3 !  dx3' 

282  HIGHER  MATHEMATICS.  §  97. 

The  series  on  the  right-hand  side  is  known  as  Maclaurin's  Series. 
The  first  term  is  what  the  series  becomes  when  x  =  0 ;  the  second 
term  is  what  the  first  derivative  of  the  function  becomes  when 
x  =  0,  multiplied  by  x ;  the  third  term  is  the  product  of  the  second 
derivative  of  the  function  when  x  =  0,  into  x1  divided  by  factorial 
2... 

"/M(0)  "  means  that  f(x)  is  to  be  differentiated  n  times,  and  x 
equated  to  zero  in  the  resulting  expression.  Using  this  notation 
the  series  assumes  the  form 

u = /(o)  +  /'(0)| + /*(0)j^ + tmf^s +  •  •  •    (7) 

§  97.    Useful  Deductions  from  Maclaurin's  Theorem. 

The  following  may  be  considered  as  a  series  of  examples  of  the 
use  of  the  formula  obtained  in  the  preceding  section.  Many  of  the 
results  now  to  be  established  will  be  employed  in  our  subsequent 
work. 

I.  Binomial  Series. 

In  order  to  expand  any  function  by  Maclaurin's  theorem,  the 
successive  differential  coefficients  of  u  are  to  be  computed  and  x 
then  equated  to  zero.  This  fixes  the  values  of  the  different  con- 
stants. 

Let  u  =  {a  +  x)n, 

du/dx   =n(a  +  x)n~1j  .'.f(0)  =  nan-1; 

d2u/dx2  =  n(n  -  1)  (a  +  x)n  ~  2,  .-.  /"(0)  =  n{n  -  l)an  -  2 ; 

d*u/dx*  =  n(n  - 1)  (n  -  2)  (a  +  x)n  - 3,     .-.  /"'(0)  =  n(n  -  1)(»  -  2)an~ 3, 

and  so  on.     Now  substitute  these  values  in  Maclaurin's  series  (6), 

n       ,         n(n  -  1)        „  „ 
(a  +  x)n=  an  +  -^an-lx  +      i   2    an~  x  +•••>•         C1) 

a  result  known  as  the  binomial  series,  true  for  positive,  negative, 
or  fractional  values  of  n. 

Examples. — (1)  Prove  that 

(a  -  x)n  =  an  -  jan~lx  +      1  \    V~2a2  - (2) 

When  n  is  a  positive  integer,  and  n  =  rr„  the  infinite  series  is  cut  off  at  a 
point  where  n  -  m  =  0.     A  finite  number  of  terms  remains. 

(2)  Establish  (1  +  a;-)1/2  =  1  +  <c2/2  -  a?4/8  +  £C6/16 

(3)  Show  (1  -  x2)  -i/2  =  l  +  jc2/2  +  3x^8  +  5a;6/16  +  . . . 

(4)  Show  (1  +  x1)  ~  2  =  1  -  x2  +  x4  -  . . .  Verify  this  result  by  actual 
division. 


§  97.  INFINITE  SERIES  AND  THEIR  USES.  283 

II.  Trigonometrical  Series. 

Suppose  u  =f(x)  =  sin  x.     Note  that  fatjdx  =  d(sm  x)/dx  =  cos  x ; 

d2u/dx2  =  d2(sinx)/dx2  =  d(GO$x)/dx  =  -  sin  a;,  etc.;  and  that  sin  0  =  0, 

-  sin  0  =  0,  cos  0  =  1,  -  cos  0  =  -  1. 

Hence,  we  get  the  sine  series, 

x      xz       x5       x7 
Binaj-j-  3T+5T-7T+ (3) 

In  the  same  manner  we  find  the  cosine  series 

o>*2  /j»4  />»6 

oosa;==1-2T  +  IT~6T+ <*> 

These  series  are  employed  for  calculating  the  numerical  values 
of  angles  between  0  and  \tt.  All  the  other  angles  found  in  ' '  trigo- 
nometrical tables  of  sines  and  cosines,"  can  be  then  determined  by 
means  of  the  formulae,  page  611, 

sin(j7r  -  x)  =  cos  x ;  cos( ^-n-  -  x)  —  sin  x. 
Now  let 

u  =  f(x)  =  tan  re.     .-.  u  cos  x  =  sin  a;. 

From  page   67,  by  successive   differentiation  of    this   expression, 

remembering  that  ux  =  dujdx,  u2  =  d2u/dx2,  . . .,  as  in  §  8, 

.*.  i^cosx  -  ^sin#  =  cos  a;  ; 

.'.  u2cosx  -  2^8^^  -  ugosx  =  -  sin  x ; 

.*.  W3COS  x  -  3%2sin  x  -  3^003  x  +  u  sin  x  =  -  cos  x. 

By  analogy  with  the  coefficients  of  the  binominal  development  (1), 

or  Leibnitz'  theorem,  §  21, 

n  .  n(n  -  1) 

UnGOSX  -  ■^Un_1SlD.X *3 — o — ^n-2C0S  X   +    '  •  •   =   n^   deriv*   Sln  X' 

Now  find  the  values  of  u,  uv  u2,  u3  ...  by  equating  x  =  0  in 
the  above  equations,  thus, 

/(0)  =  /"(0)  -  ....  0  j  /'(0)  =  1,  /"'(0)  =  2,  .; . 
Substitute  these  values  in  Maclaurin's  series  (7),  preceding  section. 
The  result  is,  the  tangent  series : 

x      2a?3      16a;5  x      a;3      2a;5 

tan  x  =  j  +  q-j-  +  -g-r-  +  . . . ;  or,  tan  x  =  T  +  IT  +  15+-"  (5) 

IZT.  Inverse  Trigonometrical  Series. 

Let  0  =  tan-Ja\     By  (3),  §  17  and  Ex.  (4)  above, 

.-.  dO/dx  =  (1  +  x2)  -1  =  1  -  x2  +  x*  -  xQ  +  . . . 
By  successive  differentiation  and  substitution  in  the  usual  way,  we 
find  that 


284  HIGHER  MATHEMATICS.  §  97. 

tan-1^  =  x  -  j  +  j  -  ...,         .         .         (6) 

or,  from  the  original  equation, 

0  =  tan 6  -  |tan3<9  +  itan50  -...,.  .  (7) 
which  is  known  as  Gregory's  series.  This  series  is  known  to 
be  converging  when  0  lies  between  -  \tt  and  \w  ;  and  it  has 
been  employed  for  calculating  the  numerical  value  of  ir.  Let 
0  =  45°  =  Jw,  .'.  x  - 1.     Substitute  in  (6), 

7T_1        XI        1        _1_         1 
4"  3  +  5       7  +  9      11  +  13       •" 

The  so-called  Leibnitz  series.     We  can  obtain  the  inverse  sine 

series 

,  lx3      Sx5       5  x7 

sm     x  =  x+2J  +  85+\ET+ ^ 

in  a  similar  manner.  Now  write  x  =^  J,  sin_1aj  =  }tt.  Substitute 
these  values  in  (8).  The  resulting  series  was  used  by  Newton  for 
the  computation  of  *. 

IV.  The  Niimerical  value  of  ir. 

This  is  a  convenient  opportunity  to  emphasize  the  remarks  on 
the  unpracticable  nature  of  a  slowly  converging  series.  It  would 
be  an  extremely  laborious  operation  to  calculate  -n-  accurately  by 
means  of  this  series.  A  little  artifice  will  simplify  the  method, 
thus, 


V      3y  +  V5     7;  +  V9     liy  +  '-,;4~1.3+5.7'f9.11 


1  +£+  ' 


8  1.3*5.7  9 .  U  *  . 
which  does  not  involve  quite  so  much  labour.  It  will  be  observed 
that  the  angle  x  is  not  to  be  referred  to  the  degree-minute-second 
system  of  units,  but  to  the  unit  of  the  circular  system  (page  606), 
namely,  the  radian.  Suppose  x  =  J^,  then  tan  ~lx  =  30°  =  g-7r. 
Substitute  this  value  of  x  in  (6),  collect  the  positive  and  negative 
terms  in  separate  brackets,  thus 

To  further  illustrate,  we  shall  compute  the  numerical  value  of 
ir  to  five  correct  decimal  places.     At  the  outset,  it  will  be  obvious 


§  97.  INFINITE  SERIES  AND  THEIR  USES.  285 

that  (1)  we  must  include  two  or  three  more  decimals  in  each  term 
than  is  required  in  the  final  result,  and  (2)  we  must  evaluate  term 
after  term  until  the  subsequent  terms  can  no  longer  influence  the 
numerical  value  of  the  desired  result.     Hence  : 

Terms  enclosed  in  the  first  brackets.  Terms  enclosed  in  the  second  brackets. 

0-57735  03  0-06415  01 

0-01283  00  0-00305  48 

0-00079  20  0-00021  60 

0-00006  09  0-00001  76 

0-00000  52  0-00000  15 

0-00000  05  0-00000  02 


0-59103  89  -  0-06744  02 

.-.  7T  -  6(0-59103  89  -  0-06744  02)  =  3-14159  22. 

The  number  of  unreliable  figures  at  the  end  obviously  depends 
on  the  rapidity  of  the  convergence  of  the  series.  Here  the  last  two 
figures  are  untrustworthy.  But  notice  how  the  positive  errors  are, 
in  part,  balanced  by  the  negative  errors.  The  correct  value  of  tt  to 
seven  decimal  places  is  3*1415926.  There  are  several  shorter  ways 
of  evaluating  tt.     See  Encyc.  Brit.,  Art.  "  Squaring  the  Circle  ". 

V.  Exponential  Series. 
Show  that 

x      x2      x3  11 

^  =  1  +  l  +  ^!+r!  +  ""e  =  1  +  1  +  2l  +  3!+---     (9) 

by  Maclaurin's  series.  An  exponential  series  expresses  the  de- 
velopment of  ex,  ax,  or  some  other  exponential  function  in  a  series 
of  ascending  powers  of  x  and  coefficients  independent  of  x. 

Examples.— (1)  Show  that  if  k  =  log  a, 

k°-x2      k3x* 
a*  =  l  +  kx  +  -2Y  +  TT  + <10) 

(2)  Represent  Dalton's  and  Gay  Lussac's  laws,  from  the  footnote,  page 
91,  in  symbols.  Show  by  mathematical  reasoning  that  if  second  and  higher 
powers  of  afl  are  outside  the  range  of  measurement,  as  they  are  supposed  to  be 
in  ordinary  gas  calculations,  Dalton's  law,  v  =  v0e*o,  is  equivalent  to  Gay  Lus- 
sac's, v  =  vQ(l  +  a6). 

/*•  g*2  /*»5  /wi 

(3)  Show  e-x=1  __.  +  ___  +  __ (11) 

VI.  Euler's  Sine  and  Cosine  Series. 

If  we  substitute  J  -  1.x,  or,  what  is  the  same  thing,  ix  in 
place  of  x}  we  obtain, 


286  HIGHER  MATHEMATICS.  §  98. 

'         .,  iX         X2  lX3         #4         IX5 

eLX  =  1  a — 1 —  a 

T  1       2!       3!       4!  +  5!       "" 

/^         X2        X*  \         (X        x3        xb  \        /10X 

••^=(1-2!+i'!----)  +  {l-3!  +  5T---->     (12) 

By  reference  to  page  283,  we  shall  find  that  the  first  expression  in 

brackets,  is  the  cosine  series,  the  second  the  sine  series.     Hence, 

eiX  =  cos  a;  +  isin#.      ".         .         .         (13) 

In  the  same  way,  it  can  be  shown  that 

-        IX       x2       ix3       x*       lX5 

e~ lX  =  1 1 1 

1       2!  +  3!       4!       5!  ' 

.   —  t»       /i        x2       x^  \        (x      x3       x5  \       ,„., 

Or, 

e  ~ IX  =  cos  x  -  i  sin  x.      .         .         .         (15) 

Combining  equations  (13)  and  (15),  we  get 

J(g«  _  q-ix)  =  i  sin  x ;  \{&-x  +  e  ~  IX)  =  cos  x.  .  (16) 
The  development  by  Maclaurin's  series  cannot  be  used  if  the 
function  or  any  of  its  derivatives  becomes  infinite  or  discontinuous 
when  x  is  equated  to  zero.  For  example,  the  first  differential 
coefficient  oif{x)  =  jJx,  is  \x ""  *,  which  is  infinite  for  x  =  0,  in  other 
words,  the  series  is  no  longer  convergent.  The  same  thing  will  be 
found  with  the  functions  log  x,  cot  x,  1/x,  a1  ,x  and  sec  "  1x.  Some 
of  these  functions  may,  however,  be  developed  as  a  fractional  or 
some  other  simple  function  of  x,  or  we  may  use  Taylor's  theorem. 

§  98.    Taylor's  Theorem. 

Taylor  s  theorem  determines  the  law  for  the  expansion  of  a 
function  of  the  sum,  or  difference  of  two  variables  into  a  series 
of  ascending  powers  of  one  of  the  variables.     Now  let 

Assume  that 

uY  =  f{x  +  y)  =  A  +  By  +  Cy2  +  Dy*  + (1) 

where  A,  B,  C,  D,  . . .  are  constants,  independent  of  y,  but  de- 
pendent upon  x  and  also  upon  the  constants  entering  into  the 
original  equation.  It  is  required  to  find  values  for  A,  B}  C,  . . . 
which  will  make  the  series  true.  Since  the  proposed  development 
is  true  for  all  values  of  x  and  y,  it  will  also  be  true  for  any  given 
value  of  x,  say  a.    Now  let  A',  B',  O,  . . .  be  the  respective  values 


§  98.  INFINITE  SERIES  AND  THEIR  USES.  287 

of  A,  B,  0,  ...  in  (1)  when  x  =  a.     Hence,  we  start  with  the  as- 
sumption that 

v!  -/(a  +  y)  -  A'  +  B'y  +  G'y2  +  D'y*  +  ...     .         (2) 
Put  z  =  a  +  y,  hence,  y  =  z  -  a,  and  Maclaurin's  theorem  gives 

us 

u'  =f(z)  =  A'  +  B\z  -  a)  +  G\z  -  af  +  D\z  -  af  +  . . . 

Now  write  down  the  successive  derivatives  with  respect  to  z. 
?g.  =  f'{z)  =  B'  +  2C(s  -  a)  +  3D'(*  -  af  +  . . . 

2gi  =  /"(*)  =  20'  +  2 .  3Z>'(*  -  a)  +  3  .  4S(*--  a-)2  +  . . . 

^  ./"'(*)  =  2. 3D'  +  S* 8.4*0!  -  «)  +  ••• 

While  Maclaurin's  theorem  evaluates  the  series  upon  the  assump- 
tion that  the  variable  becomes  zero,  Taylor's  theorem  deduces  a 
value  for  the  series  when  x  =  a.     Let  z  =  a,  then  y  =  0,  and  we 
get 
f(a)  =  A' ;  f(a)  -  B' ;  /"(a)  =  20' ;  .-.  0'  =  if  (a) ;  /'"(a)  =  2 . 3D'; 

Substitute  these  values  of  A',  B't  G\  . . .  in  equation  (2),  and  we 
get 

u'  =  f(a  +  ./)  =  /(a)  +  /'(a)f  +  /»g  +  /'"(a)f3,  +  . . .      (3) 

for  the  proposed  development  when  x  assumes  a  given  particular 
value.     But  a  is  any  value  of  x ;  hence,  if 

»-/(?) (4) 

Substitute  these  values  of  A,  B,  G,  D  in  the  original  equation 
and  we  obtain 

du  y      <Pu    y2        dfiu      y3 

•»  -/(*+*>-»  +  2&\  +  mm  *  &  rkra +  ••■  <s> 

The  series  on  the  right-hand  side  is  known  as  Taylor's  series. 
The  first  term  is  what  the  given  function  becomes  when  y  =  0 ; 
the  second  term  is  the  product  of  the  first  derivative  of  the  function 
when  y  =  0,  into  y ;  the  third  term  is  the  product  of  the  second 
derivative  of  the  function  when  y  =  0,  into  y2  divided  by  factorial 
2  . . .  In  (5),  u  =  f(x)  is  obtained  by  putting  y  =  0.  Thus,  in  the 
development  of  (x  +  yf  by  Taylor's  theorem, 
u  =  f(x)  =  xb ;  du/dx  =  f'(x)  =  5x* ;  d2u/dx2  =  /"(a?)  =  4 .  5x* ;  ... 
/.  (x  +  y)b  =  x5  +  5x*y  +  10xsy2  +  10x2y*  +  5xy*  +  y5. 


288  HIGHER  MATHEMATICS.  §  98. 

Instead  of  (5),  we  may  write  Taylor's  series  in  the  form, 

«fi  -a* + y) =/(*)  +/'o4 + /"(^o + r&uks + ■  ■  ■ (6) 

Or,  interchanging  the  variables, 

«i  -  A*  +  y)=  Ay)  +  f(y)i  +  W)£%  +  /"Wfno  +  •  •  •  (7> 

I  leave  the  reader  to  prove  that 

f(x  -  y)  -  /(*)  -  /' (*)f  +  /"(as)  jg  -  /'"(^  +  . . .         (8) 

Maclaurin's  and  Taylor's  series  are  slightly  different  expressions 
for  the  same  thing.  The  one  form  can  be  converted  into  the  other 
by  substituting  f(x  +  y)  for  f{x)  in  Maclaurin's  theorem,  or  by 
putting  y  =  0  in  Taylor's. 

Examples. — (1)  Expand  v^  =  (x  +  y)n  by  Taylor's  theorem. ■   Put  y  =  0 

and  u  =  X",  as  indicated  above, 

du  .     d2u 

.-.  j-x  =  nx*-1 ;  ^  =*(tt  -  l)an-2,  etc. 

Substitute  the  values  of  these  derivatives  in  (7). 

.-.  Mj  =  (jc  +  y)n  =  sb*1  +  nx"  ~  ly  +  %n(n  -  l)xn  -  2y2  +  . . . 

(2)  If  k  =  log  a ;  ^  =  a*  +  *  =  a*(l  +  ky  +  \k2y2  +  |fcy  +  ...). 

(3)  Show  (a  +  y  +  «)t  =  (a?  +  a)*  +  £?/(x  +  a)  -  i  -  . . .  If  x  =  -  a,  the 
development  fails. 

/        w2       v4  \  f        y3  \ 

(4)  Show  sin  (a;  +  y)  =  sin  zf  1  -  ^  +  ^y  -  . . .  )  +  cos  xi  y  -  g-j  +  ...)• 

(5)  The  numerical  tables  of  the  trigonometrical  functions  are  calculated  by 

means  of  Taylor's  or  by  Maclaurin's  theorems.     For  example,  by  Maclaurin's 

theorem, 

x3       xP  X2       x^ 

sina  =  a;-g-j  +  g-j--...;  cosicrrl-^-f  +  ^j-... 

But  35°  =  -610865  radians,  and  .-.  sin  35°  =  sin  -610865.     Consequently, 

sin  35°  =  -610865  -  £(-610865)3  +  ^(^lOSeS)5  -  . . .  =  -57357  . . . 

In  the  same  way,  show  that  cos  35°  =  '81915 . . .     Again  by  Taylor's  theorem, 

sin  36°  =  sin  (35°  +  1°) ; 

cos  35°  sin  35° 

.-.  sin  36°  =  sin  35°  +  -jy— (-017453)  -  — 2j-(-017453)2  -  . . .  =  -58778  .  . . 

(6)  Taylor's  theorem  is  used  in  tabulating  the  values  of  a  function  for  dif- 
ferent values  of  the  variable.  Suppose  we  want  the  value  of  y  =  a;(24  -  x2)  for 
values  of  x  ranging  from  2-7  to  3*3.  First  draw  up  a  set  of  values  of  the 
successive  differential  coefficients  of  y. 

f(x)  =  /(3)  =  <b(24  -  x2)  =  45  ;  f{x)  =  f(B)  =  24  -  3x2  =  -  3 ; 
f'{x)  =  /"(3)  =  -  6*  =  -  18 ;  f"(x)  =  /'"(3)  =  -  6. 
By  Taylor's  theorem, 

/(3  ±  h)  =/(3)  ±f(B)h  +  tf"(3)h2  ±  tf'"{3)h?  =  45  +  3h  -  9h\+  h\ 


§  98.  INFINITE  SEKIES  AND  THEIR  USES.  289 


2-7  =  3- 

-  0-3  ;  2*8  =  3  -  0-2  ; . 

..,3-3  =  3  +  0-3.     Hence, 

/(2-7)  =  45  +  0-9  - 

-  0-81  +  0-027  =  45-117. 

/(2-8)  =  45  +  0-6  - 

-  0-36  +  0-008  =  45-148. 

/(2*9)  =  45  +  0-3  - 

-  0-09  +  0-001  =  45-211. 

/(3*0)  =  45 

=  45-000. 

/(3-1)  =  45  -  0-3  - 

-  0-09  -  0-001  =  44-609. 

/(3-2)  =  45  -  0-6  - 

-  0-36  -  0-008  =  44-032. 

/(3*3)  =  45  -  0-9  ■ 

-  0-81  -  0-027  =  43-263. 

(7)  Show  log (x  +  y)  =  log*  +f --£r  +  |^r-  ... 

'*)  Expand  log  (n  +  h)  and  also  log  (n  +  1).  Observe  that  we  can  neglect 
terms  containing  second  powers  of  h,  if  h  is  less  than  unity,  and  n  is  large. 
Thus,  if  h  <  1,  and  n  is  10,000,  hfn  <  0-0001 ;  the  second  term  of  the  ex- 
pansion is  less  than  0*000000,005  ;  and  the  next  term  still  less  again.  By 
division  of  the  two  expansions,  we  get  the  important  result, 

log  (n  +  h)  -  log  n  _  h 

log  (n  +  1)  -  log  n       1 '  ' 

or, 

Incr.  when  log  ft  becomes  log(n  +  h)  :  Incr.  when  logn  becomes  log(n  +  1)  —  h  :  1, 
provided  the  differences  between  two  numbers  n  and  h  are  such  that  n  is  of 
the  order  of  10,000  when  a*  is  less  than  unity.  This  formula,  known  as  the 
rule  of  proportional  parts,  is  used  for  finding  the  exact  logarithm  of  a  number 
containing  more  digits  than  the  table  of  logarithms  allows  for,  or  for  finding 
the  number  corresponding  to  a  logarithm  not  exactly  coinciding  with  those  in 
the  tables.     The  following  examples  wil  make  this  clear : — 

(9)  Find  the  logarithm  of  46502-32,  having  given 

log  46501  =  4-6674623  ;  log  46502  =  4-6674716  ;  difference  =  0-0000093. 
Let  h  denote  the  quantity  to  be  added  to  the  smaller  of  the  given  logs.     The 
problem  may  be  stated  thus, 

log  n  =  log  46501  =  4-6674623  ; 

log  (n  +  1)  =  log  (46501  +  1)      =  4-6674623  +  0-0000093  ; 

log  (n  +  h)  =  log  (46501  +  0-32)  =  4-6674623  +  x. 

By  (9),  that  is,  by  simple  rule  of  three  :  if  a  difference  of  1  unit  in  a  number 
corresponds  with  a  difference  of  0*0000093  in  the  logarithm,  what  difference 
in  the  logarithm  will  arise  when  the  number  is  augmented  by  0*32  ? 

.-.  1  :  0-32  =  0-0000093  :  tr,  .-.  x  =  0-00000298 . . . 
The  required  logarithm  is,  therefore,  4*6674653. 

Again,  find  the  number  whose  logarithm  is  4*6816223,  having  given 

log  48042  =  4-6816211 ;  log  48043  =  4-6816301. 

Since  a  difference  of  unity  in  the  number  causes  a  difference  of  0*0000090 
in  the  logarithm,  what  will  be  the  difference  in  the  number  when  the  logarithms 
differ  by  0*0000012  ? 

.-.  1  :  h  =  00000090  :  00000012  ;  .-.  h  =  0*13.     The  number  is  48042-13, 

(10)  Show  log  (1  +  y)  =  y  -  $y*  +  #■  -  |fl*  +  .,,. 

T 


290  HIGHER  MATHEMATICS.  §  98. 

This  series  may  be  employed  for  evaluating  log  2,  but  as  the  series  happens 
to  be  divergent  for  numbers  greater  than  2,  and  very  slowly  convergent  for 
numbers  less  than  2,  it  is  not  suited  for  general  computations. 

(11)  Show  log  (1  -  y)  =  -  (y  +  \y>  +  | y*  +  \y*  +  . . .). 

If  y  =  4,  the  development  gives  a  divergent  series  and  the  theorem  is  then 
said  to  fail.    The  last  four  examples  are  logarithmic  series. 

A  series  suitable  for  finding  the  numerical  values  of  logarithms  may  here 
be  indicated  as  a  subject  of  general  interest,  but  of  no  particular  utility  since 
we  oan  purchase  "ready-made  tables  from  a  penny  upwards".  But  the 
principle  involved  has  useful  applications. 

Subtract  the  series  in  Ex.  (11)  from  that  in  Ex.  (10)  and  we  get 

a  series  slowly  convergent  when  y  is  less  than  unity.  Let  n  have  a  value 
greater  than  unity.     Put 

n  +  1      1  +  y        ,  1 

-»-  =  FTP  sothaty  =  5— y. 

Henoe,  when  n  is  greater  than  unity,  y  is  less  than  unity.  By  substitution, 
therefore, 

log  (»  +  1)  .  log  n  +  afc^  +  3(2»  +  I)"  +  •  ■  •} 
This  series  is  rapidly  convergent.    It  enables  us  to  compute  the  numerical 
value  of  log  (n  +  1)  when  the  value  of  log  n  is  known.     Thus  starting  with 
n  =  1,  log  n  =  0,  the  series  then  gives  the  value  of  log  2,  hence,  we  get  the 
value  of  log  3,  then  of  log  4,  etc. 

(12)  Put  y  =  -  x  in  Taylor's  expansion,  and  show  that 

f(x)=f(0)+f(x).x-y"(x).x*  +  ..., 
known  as  Bernoulli's  series  (of  historical  interest,  published  1694). 

Mathematical  text-books,  at  this  stage,  proceed  to  discuss  the 
conditions  under  which  the  sum  of  the  individual  terms  of  Taylor's 
series  is  really  equal  to  f(x  +  y).  When  the  given  function  f(x  +  y) 
is  finite,  the  sum  of  the  corresponding  series  must  also  be  finite,  in 
other  words,  the  series  must  either  be  finite  or  convergent.  The 
development  is  said  to  fail  when  the  series  is  divergent. 

It  is  not  here  intended  to  show  how  mathematicians  have  suc- 
ceeded in  placing  Taylor's  series  on  a  satisfactory  basis.  That 
subject  belongs  to  the  realms  of  pure  mathematics.1  The  reader 
may  exercise  "belief  based  on  suitable  evidence  outside  personal 
experience,"  otherwise  known  as  faith.  This  will  require  no  great 
mental  effort  on  the  part  of  the  student  of  the  physical  sciences. 
He  has  to  apply  the  very  highest  orders  of  faith  to  the  fundamental 

1  If  the  student  is  at  all  curious,  Todhunter,  or  Williamson  on  "  Lagrange's 
Theorem  on  the  Limits  of  Taylor's  Series,"  is  always  available, 


§  99.  INFINITE  SERIES  AND  THEIR  USES.  291 

principles — the  inscrutables — of  these  sciences,  namely,  to  the 
theory  of  atoms,  stereochemistry,  affinity,  the  existence  and  pro- 
perties of  interstellar  ether,  the  origin  of  energy,  etc.,  etc.  What 
is  more,  "  reliance  on  the  dicta  and  data  of  investigators  whose 
very  names  may  be  unknown,  lies  at  the  very  foundation  of  physical 
science,  and  without  this  faith  in  authority  the  structure  would  fall 
to  the  ground  ;  not  the  blind  faith  in  authority  of  the  unreasoning 
kind  that  prevailed  in  the  Middle  Ages,  but  a  rational  belief  in  the 
concurrent  testimony  of  individuals  who  have  recorded  the  results 
of  their  experiments  and  observations,  and  whose  statements  can 
be  verified . .  .".1 

The  rest  of  this  chapter  will  be  mainly  concerned  with  direct 
or  indirect  applications  of  infinite  converging  series. 

§  99.    The  Contact  of  Curves. 

The  following  is  a  geometrical  illustration  of  one  meaning  of 
the  different  terms  in  Taylor's  development.  If  four  curves  Pa, 
Pb,  Pc,  Pd,  . . .  (Fig.  123)  have  a  common  p 

point  P,  any  curve,  say  Pc,  which  passes 


between  two  others,  Pb,  Pd,  is  said  to  have         fn  /i, 

a  closer  contact  with  Pb  than   Pd.     Now     fig.  123.— Oontaot  of 
let  two  curves  P0P  and  PQPY  (Fig.  124)  re-  Curves- 

ferred  to  the  same  rectangular  axes,  have  equations, 

y  =  f(x)  ;  and,  yx  =  f^xj.  .         .         (1) 

Let  the  abscissa  of  each  curve  at  any  given  point,  be  increased  by 
a  small  amount  h,  then,  by  Taylor's  theorem, 

f(x  +  h)  =  y  +  a£h  +  a42l  +  --'i 

A(%;  +  *)-*+^'+^  }?+:•»      •      (2) 

If  the  curves  have  a  common  point  P0,  x  =  xv  and  y  =  y1  at 
the  point  of  contact.  Since  the  first  differential  coefficient  repre- 
sents the  angle  made  by  a  tangent  with  the  a?-axis,  if,  at  the  point 

1  Excerpt  from  the  Presidential  Address  of  Dr.  Carrington  Bolton  to  the  Washing- 
ton Chemical  Society,  English  Mechanic,  5th  April,  1901. 

m  ♦ 


292 


HIGHER  MATHEMATICS. 


§100. 


the  curves  will  have  a  common  tangent  at  P0. 
contact  of  the  first  order.     If,  however, 


This  is  called  a 


dy       dy1 

*v  y  =  Vi ;  3j  =  ^ ; 


and 


<Py     dfy 


da;2     da^2 

the  curves  are  said  to  have  a  contact  of  the  second  order,  and 

so  on  for  the  higher  orders  of  contact. 

If  all  the  terms  in  the  two  equations  are  equal  the  two  curves 
will  be  identical ;  the  greater  the  number  of 
equal  terms  in  the  two  series,  the  closer  will 
be  the  order  of  contact  of  the  two  curves.  If 
the  order  of  contact  is  even,  the  curves  will 
intersect  at  their  common  point ;  if  the  order 
of  contact  is  odd,  the  curves  will  not  cross  each 
other  at  the  point  of  contact. 


<tf 


x 


M, 


0  I 

Fig.  124.— Contact  of 
Curves. 


Examples. — (1)  Show  that  the  curves  y=  -  x2,  and 
y  =  3x  -  x2  intersect  at  the  point  x  =  0,  y  =  0.  Hint. 
The  first  differential  coefficients  are  not  equal  to  one  another  when  we  put 
x  =  1.  Thus,  in  the  first  case,  dyjdx  =  -2a;  =  -2  =  0;  and  in  the  second, 
dyjdx  =  3  -  2x  =  1. 

(2)  Show  that  the  tangent  crosses  a  curve  at  a  point  of  inflexion.  Let 
the  equation  of  the  curve  be  y  =  f(x) ;  of  the  tangent,  Ax  +  By  +  G  =  0. 
The  necessary  condition  for  a  point  of  inflexion  in  the  ourve  y  =  f(x)  is  that 
d'Hjfdx2  =  0.  But  for  the  equation  of  the  tangent,  cPy/dx2  is  also  zero. 
Hence,  there  is  a  contact  of  the  second  order  at  the  point  of  inflexion,  and 
the  tangent  crosses  the  curve. 


§  100.    Extension  of  Taylor's  Theorem. 

Taylor's  theorem  may  be  extended  so  as  to  include  the  expan- 
sion of  functions  of  two  or  more  independent  variables.     Let 

a-/(^y)i  (1) 

where  x  and  y  are  independent  of  each  other.  Suppose  each 
variable  changes  independently  so  that  x  becomes  x  +  h,  and  y 
becomes  y  +  k.  First,  let  f(x,  y)  change  to  f(x  +  h,  y).  By 
Taylor's  theorem 

"du,       'dhi  h2 


/(^  +  ^1/)  =  ^  +  ^  +  ^29TT  + 


~bx22\ 


(2) 


If  y  now  becomes  y  +  k,  each  term  of  equation  (2)  will  change  so 

that 

7)11-       d2^  k* 
U  becomes  U  +  ^k  +  ^  eft  +  •  •  »5 


§  101.  INFINITE  SERIES  AND  THEIR  USES.  293 

t)w  'du       Wu  7>2u  <)%        <)% 

*x  becomes  Si  +  toty*  +  '";  ^becomes  W  +  W^k  +  •  •" 
by  Taylor's  theorem.     Now  substitute  these  values  in  (2)  and  we 
obtain,  if  u'  denotes  the  value  of  u  when  x  becomes  x  +  h,  and  y 
becomes  y  +  k, 

u'  =f(x  +  h,y  +  k)\ 

~dui      ~b2U   k  ~d2U  _  _  <)W7       c)%  h 

8u  =  u'-u=f(x  +  h,  y  +  k)-f{x,y); 

,       3«,     Du,     1/ft,,     .»,,     ft,,\  /ox 

The  final  result  is  exactly  the  same  whether  we  expand  first 
with  respect  to  y  or  in  the  reverse  order. 

By  equating  the  coefficients  of  hk  in  the  identical  results  ob- 
tained by  first  expanding  with  regard  to  h,  (2)  above,  and  by  first 
expanding  with  regard  to  k,  we  get 

TixDy      ~dy~dx' 
which  was  obtained  another  way  in  page  77.     The  investigation 
may  be  extended  to  functions  of  any  number  of  variables. 

§  101.  The  Determination  of  Maximum  and  Minimum  Values 
of  a  Function  by  means  of  Taylor's  Series. 

I.  Functions  of  one  variable. 

Taylor's  theorem  is  sometimes  useful  in  seeking  the  maximum 
and  the  minimum  values  of  a  function,  say, 

u  =  f(x). 
It  is  required  to  find  particular  values  of  x  in  order  that  y  may 
be  a  maximum  or  a  minimum.     If  x  changes  by  a  small  amount 
h,  Taylor's  theorem  tells  us  that 

dun       1  d2u_,  _      1  d%,  „ 

a*  ±  h)  -  m  -  ±  as*  + 1 »»  ±  g  Mh° +  ■  •  •  •    (^ 

according  as  h  is  added  to  or  subtracted  from  x. 

First,  it  must  be  proved  that  h  can  be  made  so  small  that  the 

dy 
term  -r-h  will  be  greater  than  the  sum  of  all  succeeding  terms  of 

either  series.  Assume  that  Taylor's  series  may  be  written, 

f(x  +  h)  =  u  +  Ah  +  Bh2  +  Ghs  +  . . . , 
where  A,  B,  C, .  . .  are  coefficients  independent  of  h  but  dependent 


294  HIGHER  MATHEMATICS.  §  101. 

upon  x,  then,  if  Bh  =  Bh  +  Gh2  +  . . .  -  (B  +  Gh  +  . .  .)h,  and 

fix  +  h)  =  u  +  h(A  +  Bh).        .         .        (2) 
Consequently,  for  sufficiently  small  values  of  h,  it  will  be  obvious 
that  Bh  must  be  less  than  A. 
Let  us  put 

8u=f(x±h)  -f{x). 

If  u  is  really  a  maximum,  ever  so  small  a  change — increase  or 
decrease — in  the  value  of  x  will  diminish  the  value  of  u  ;  and  fix) 
must  be  greater  than  fix  ±  h).     Hence,  for  a  maximum, 

Su  =  fix  ±h)  —  fix)  must  be  negative. 
Again,  if  u  is  really  a  minimum,  then  u  will  be  augmented  when  x 
is  increased  or  diminished  by  h.     In  other  words,  if  u  is  a  minimum, 

Su  =  fix  ±  h)  -  fix)  must  be  positive. 

Illustration.— The  function  u  =  4<c3  -  3sca  -  18sc  will  be  a  maximum 
when  x  =  -  1.  In  that  case,  f(x)  =  11 ;  if  we  put  some  small  quantity, 
say  £,  in  place  of  h,  then  f{x  +  h)  =  +  ^,  and  f(x  -  h)  =  +  ^-.  Hence, 
f(x±  h)  -  f{x)  will  be  either  -  ^&  or  -  ^-.  You  can  also  show  in  the  same 
manner  that  u  will  be  a  minimum  when  x  =  $ . 

Now  if  h  is  made  small  enough,  we  have  just  proved  that  the 
higher  derivatives  in  equations  (1)  will  become  vanishingly  small ; 
and  so  long  as  the  first  derivative,  du/dx,  remains  finite,  the  al- 
gebraic sign  of  Su  will  be  the  same  as 

*        du7 
ax 
At  a  turning  point — maximum  or  minimum — we  must  have,  as 
explained  in  an  earlier  chapter, 

%-* 

dx 
Substituting  this  in  the  above  series, 

1    d%79     1    d%7„ 

remains.  Now  h  may  be  taken  so  small  that  the  derivatives 
higher  than  the  second  become  vanishingly  small,  and  so  long  as 
dhijdx2  remains  finite,  the  sign  8u  will  be  the  same  as  that  of 

*        d2u    h2 

But  h2,  being  the  square  of  a  number,  must  be  positive.  The  sign 
of  the  second  differential  coefficient  will,  in  consequence,  be  the 
same  as  that  of  8u.  But  u  =  fix)  is  a  maximum  or  a  minimum 
according  as  Su  is  negative  or  positive.     This  means  that  y  will  be 


§101. 


INFINITE  SEftlES  AND  THEIR  USES. 


295 


a  maximum  when  dy/dx  =  0  and  d2y/dx2  is  negative,  and  a  mini- 
mum, if  d2y/dx2  is  positive. 

If,  however,  the  second  differential  coefficient  vanishes,  the 
reasoning  used  in  connection  with  the  first  differential  must  be 
applied  to  the  third  differential  coefficient.  If  the  tjiird  derivative 
vanishes,  a  similar  relation  holds  between  the  second  and  fourth 
differential  coefficients.    See  Table  I . ,  page  168.    Hence  the  rules : — 

1.  y  is  either  a  maximum  or  a  minimum  for  a  given  value  of  x 
only  when  the  first  non-vanishing  derivative,  for  this  value  of  x,  is 
even. 

2.  y  is  a  maximum  or  a  minimum  according  as  the  sign  of  the 
non-vanishing  derivative  of  an  even  order,  is  negative  or  'positive. 

In  practice,  if  the  first  derivative  vanishes,  it  is  often  con- 
venient to  test  by  substitution  whether  y  changes  from  a  positive 
to  a  negative  value.     If  there  is  no  change  of  sign,  there  is  neither 

1  maximum  nor  a  minimum.     For  example,  in 

y  =  Xs  -  Sx2  +  3x  +  7 ;  .'.  ^|  =  3a;2  -  6x  +  3. 

For  a  maximum  or  a  minimum,  we  must  have 
x2  -  2x  +  1  =  0 ;  .-.  x  =  1. 
If  x  =  0,  y  =  7  ;   if  x  =  1,  y  =  8  ;   if  x  =  2,  y  =  9.      There  is  no 
change  of  sign  and  x  =  1  will  not  make  the  function  a  maximum 
or  a  minimum. 

Examples. — (1)  Test  y  =  x3  -  12a;2  -  60a;  for  maximum  or  minimum 
values.  dy/dx  m  8a;2  -  24a;  -  60  ;  .%  x2  -  8x  -  20  =  0, 
or  x  =  -  2,  or  +  10.  dPy/dx2  =  6a;  -  24  ;  or,  x  =  +  4. 
Since  d^/dx2  is  positive  when  x  =  10  is  substituted, 
x  =  10  will  make  y  a  minimum.  When  -  2  is  substi- 
tuted, 6?y\dx2  becomes  negative,  hence  x  =  -  2  will  make 
y  a  maximum.  This  can  easily  be  verified  by  plotting 
(Fig.  125),  for,  if 

x  =  -  3,         -  2,  -  1,      . . .  +  9,  +10, 

y  =  +  45,       +64  (max.),      +48, 

(2)  What  value  of  x  will  make  y  t 
expression,  y  =  Xs  -  6x2  +  11a;  +6? 
dy/dx  =  3x2  -  12a?  +  11  =  0;  .\ 
x  =  2  +  n/J  ;  dttyjdx2  =  6a;  -  12.  If 
x  =  2  +  n/J;  dh/jdx*  m  6\/J  =  +  2\/3; 
and  if  x  =  2  -  \/£,  dhf/dx2  =-  2\/3. 
Hence  2  +  \/^  makes  y  a  minimum,  and 

2  -  s/%  makes  y  a  maximum  (see  Pig.  126). 

(8)  Show  that  a;3  -  9a;2  +  15a;  -  3  is  a 


+y 

-X 

s> 

o 

+x 

/ 

\ 

/ 

-y 

v 

\j 

Fig.  125. 

+  11,... 
783,       -  800  (min.),     -  781, . . . 
maximum  or  a  minimum  in  the 


f 

0LJ — 

— r-|        +y- 

L-  x         -y 

Fig.  126. 


Fig.  127. 


296  HIGHER  MATHEMATICS.  §  101. 

maximum  when  x  =  1,  and  a  minimum  when  x  =  5.  The  graph  is  shown  in 
Fig.  127. 

II.  Functions  of  tivo  variables. 

To  find  particular  values  of  x  and  y  which  will  make  the 
function, 

«  -/(?>  y)» 

a  maximum  or  a  minimum.  As  before,  when  x  changes  by  a 
small  amount  h,  and  y  by  a  small  amount  k,  if  f(x,  y)  is  greater 
than  f(x  ±  h,  y  ±  k),  for  all  values  of  h  or  k,  then  f(x,  y)  is  a 
maximum.     Hence,  if 

Su  =  f(x  ±  h,  y  ±  k)  -  f(x,  y)  is  negative, 
u  will   be   a   maximum ;    whereas   when  fix,    y)   is   less '  than 
f{x  ±h,y±  k), 

Su  =  f(x  ±  h,  y  ±  k)  -  f(x,  y)  is  positive, 

and  u  will  be  a  minimum. 

Illustration. — The  function  u  =  xhj  +  xy2  -  Sxy  will  be  a  minimum 
when  x  =  1  and  y  =  1.  In  that  case  f(x,  y)  =  -  1 ;  and  if  we  put  h  =  %,  and 
k  =  J — any  other  small  quantity  will  do  just  as  well — then  f(x  +  ht  y  +  k)=0; 
and  f{x  -  h,y  -  k)  =  -\.    Hence,  fix  ±h,y  ±  k)  -  fix,  y)  =  +  1,  or  +  £. 

Also,  let 

Su  =  f(x  +h,y  +  k)  -  f(x,  y). 

Let  us  now  expand  this  function  as  indicated  in  the  preceding 
section,  and  we  get 

hu  =  lih  +  Tyk  +  2^2  +  2*xtyhk  +  Jf?)  +  '  •  ■  (3) 
By  making  the  values  of  h  and  k  small  enough,  the  higher  orders 
of  differentials  become  vanishingly  small.  But  as  long  as  T^uftx 
and  "bufiy  remain  finite,  the  algebraic  sign  of  ~du  will  be  that  of 

©.»  *  ©.*• 

At  a  turning  point — maximum  or  a  minimum — we  must  have 
t)w        7)u 

&  +  p*° w 

and,  since  h  and  k  are  independent  of  each  other,  and  the  sign  of 
Su,  in  (4),  depends  on  the  signs  of  h  and  k,  u  can  have  a  maximum 
or  a  minimum  value  only  when 

©f-0;"d^).-a    •        •        •        (5) 


§101. 


INFINITE  SERIES  AND  THEIE  USES. 


297 


We  can  perhaps  get  a  clearer  mental  picture  of  what  we  are 
talking  about  if  we  imagine  an  undulating  surface  lying  above  the 
icy-plane.  At  the  top  of  an  isolated  hill,  P  (Fig.  128),  u  will  be  a 
maximum  ;  at  the  bottom  of  a  valley  or  lake,  Q,  u  will  be  a  mini- 
mum. The  surface  can  only  be 
horizontal  at  the  point  where 
"bufix  and  ~dufdy  are  both  zero. 
At  this  point,  u  will  be  either  a 
maximum  or  a  minimum.  It 
is  easy  to  see  that  if  APBG 
is  a  surface  represented  by 
u  =  f(x,  y),  ~du/~dx  is  the  slope 
of  the  surface  along  AP,  and 
'du/'dy,  the  slope  along  BP. 
The  line  Pb  represents  the  slope  Tiufbx  at  P,  and  Pa  the  slope 
'bu/'by  at  P. 

If  u  is  really  a  maximum,  it  follows  from  our  previous  work, 
page  159,  that  ib2u/'dx2  and  Wufiy2  must  be  negative,  just  as  surely 
as  if  P  is  really  the  top  of  a  hill,  movement  in  the  directions  Pb, 
or  Pa  must  be  down  hill.  And  similarly,  if  we  are  really  at  the 
bottom  of  a  valley,  ^2u/~dx2  and  Wufoy2  must  be  both  positive. 

Let  us  now  examine  the  sign  of  8u  in  (3)  when  'du/'dx  and  ^ufoy 
are  made  zero ;  h  and  k  can  be  made  so  small  that 


Fig.  128. 


Su 


=  J°Ah2  +  2 


^hk  +  —k2 
~dxty  ty2     , 


(6) 


2\dx2' 

remains.     For  the  sake  of  brevity,  write  the  homogeneous  quad- 
ratic (6)  in  the  form 

ah2  +  2bhk  +  ck2.  ...         (7) 

Add  and  subtract  b2k2/a ;  rearrange  terms,  and  we  get  the  equiva- 
lent form 


H{ah  +  bk)2  +  (ac  -  b2)k2\, 


(8) 


which  enables  us  to  see  at  a  glance  that  for  small  values  of  h  and 
k  the  sign  of  (7),  or  (6),  is  independent  of  h  and  k  only  when 
ac  -  b2  is  positive  or  zero,  for  if  ao  -  b2  is  negative,  the  expression 
will  be  positive  when  k  =  0,  and  negative  when  ah  +  bk  is  zero. 
Consequently,  in  order  that  we  may  have  a  real  maximum  or 
minimum,  ac  must  be  greater  than  b2  ;  or  what  is  the  same  thing, 


~b2u      ~b2u 


*bx2      ~dy 


2  must  be  greater  than 


/  ~b2U  \2 

\Zx~dy)  ' 


(9) 


298  HIGHER  MATHEMATICS.  §  101. 

This  is  called  Lagrange's  criterion  for  maximum  and  minimum 
Yalues  of  a  function  of  two  variables.  When  this  criterion  is 
satisfied  f(x,  y)  will  either  be  a  maximum  or  a  minimum.  To 
summarize,  in  order  that  u  =  f(x,  y)  may  be  a  maximum  or  a 
minimum,  we  must  have 

(2)  ^-g-  negative,  if  u  is  a  maximum ;  positive,  if  u  is  a  minimum. 

(3)  5P"  x  ^2  "  ^J   m™*  n«t  be  negative. 

If  ^"X  vi8lessthan  W/'       '      '    (10) 

or  *d2ufbx2  and  ~d2u/?)y2  have  different  signs,  the  function  is  neither 
a  maximum  nor  a  minimum.  If  a  man  were  travelling  across  a 
mountain  pass  he  might  reach  a  maximum  height  in  the  direction 
in  which  he  was  travelling,  yet  if  he  were  to  diverge  on  either  side 
of  the  path  he  would  ascend  to  higher  ground.  This  is  not  there- 
fore a  true  maximum.  A  similar  thing  might  be  said  of  a  "  bar" 
across  a  valley  for  a  minimum.     If 

i«M        WU  _  /  l2U  \2 

there  will  probably  be  neither  a  maximum  nor  a  minimum,  but 
the  higher  derivatives  must  be  examined  before  we  can  definitely 
decide  this  question. 

Examples. — (1)  Show  that  the  velocity  of  a  bimoleoular  chemical  re- 
action V=  k(a  -  x)  (b  -  x)  is  greatest  when  a  =  b.  Here  'dVj'da  =  -  k(b  -  x); 
'dV/'db  =  -  k(a  -  x).    Hence  if  k(b  -  x)  =  0 ;  and  k(a  -  x)  —  0,  a  =  b}  etc 

(2)  Test  the  function  u  —  x3  +  yz  -  3axy  for  maxima  or  minima,  Here 
d^ildx=Sxi-day=0,r.y=xila;dujdy=Syi-3a^=01r.yi-a^=xila'i-ax=0; 
,\  a?=0,  a?3  -  a3=0,  or  x=a.  The  other  roots,  being  imaginary,  are  neglected ; 
.  •.  y  =  a;2/a  =  a,  or  y  =  0 ; 


•W*  =  ex;  dxdy  =  'Sa; 


#2 


Call  these  derivatives  (a),  (6),  and  (c)  respectively,  then  if  x  =  0,  (a)  =  0, 
(6)  =  -  3a,  (c)  =  0 ;  if  x  =  a,  (a)  =  6a,  (6)  =  -  3a,  (c)  =  6a ; 

.._,_=36o2;(__)  =9a2. 

This  means  that  x  =  y  =  a  will  make  the  function  a  minimum  because 
'dhtl'dx2  is  positive ;  x  =  0  will  give  neither  a  maximum  nor  a  minimum. 

(3)  Find  the  condition  that  the  rectangular  parallelopiped  whose  edges 
are  x,  y,  and  z  shall  have  a  minimum  surface  u  when  its  volume  is  vs.    Since 


§101. 


INFINITE  SERIES  AND  THEIR  USES. 


299 


v3  =  xyz,  u  =  xy  +  yz  +  zx  =  xy  +  v3jx  +  vP/y.  When  du/dx  =  0,  x*y  =  v8 ; 
when  du/dy=0,  xy2=v3.  The  only  real  roots  of  these  equations  are  x=y=v, 
therefore  *  =  v.     The  sides  of  the  box  are,  therefore,  equal  to  each  other. 

(4)  Show  that  u  =  x3y2(l  -  x  -  y)  is  a  maximum  when  x  —  £,  y  =  £. 

(5)  Find  the  maximum  value  of  u  in  u= x-'>  -  Sax2  -  lay2,  'dufdx  =  3x(x  -  2a); 
du/dy=  -  8ay;  d^/dx2 =6{x  -  a) ;  Vhifdx'dy^Q ;  d'iuldy*=  -  8a.  Condition  (5) 
is  satisfied  by  x  «=  0,  y  =  0  and  by  x=2a,  y  =  0. 
The  former  alone  satisfies  Lagrange's  condi- 
tion (9),  the  latter  comes  under  (10). 

(6)  In  Fig.  129,  let  P1  be  a  luminous 
point ;  OMx,  OM2  are  mirrors  at  right  angles 
to  each  other.  The  image  of  P1  is  reflected 
at  Nt  and  N2  in  such  a  way  that  (i)  the  angles 
of  incidence  and  reflection  are  equal,  (ii)  the 
length  of  the  path  P^N^  is  the  shortest 
possible.  (Fermat'8  principle :  "  a  ray  of  light 
passes  from  one  point  to  another  by  the  path 
which  makes  the  time  of  transit  a  mini- 
mum ".)  Let  i1  =  rlyi2  =  r2  be  the  angles  of  FlG'  129* 
incidence  and  reflection  as  shown  in  the  figure.  To  find  the  position  of  Nj 
and  N2 : 

Let  ON2  =  x;ON1  =  y;  OM2=  ^ ;  JkfjP,  =  a^ ;  M^  =  62;  OMx  =  bv    Let 

5  =  P^  +  NXN2  +  NtP2  =  s/a\  +  (&j  -  y)2  +  six2  +  y2  +  Jfa  -  xf  +  b22. 
Fiivd  'ds/'dx  and  'ds/'dy.     Equate  to  zero,  etc.    The  final  result  is 
x  =  (aA  -  aAM&i  +  h) ;  y  =  (o^  -  a^bJlfa  +  a?). 
Note  that  x\y  =  (Oj  +  ^/(Oj  +  bj.    Work  out  the  same  problem  when  the 
angle  M2OM1  =  a. 

(7)  Required  the  volume  of  the  greatest  rectangular  box  that  oan  be  sent 
by  "Parcel  Post"  in  accord  with  the  regulation:  "length  plus  girth  must 
not  exceed  six  feet ".     Ansr.  1  ft.  x  1  ft.  x  2  ft.  =  2  eft.     Hint.  F=  xyz  is  to 


x  +  2(y  +  z)  =  6.     But  obviously  y  =  b,  .•.  V  =  xy2 


be  a  maximum  when  V  -■ 
is  to  be  a  maximum,  etc. 

(8)  Required  the  greatest  cylindrical  case  that  can  be  sent  under  the  same 
regulation.  Ansr.  Length  2  ft.,  diameter  4/ir  ft.,  capacity  2*55  eft.  Hint. 
Volume  of  cylinder  =  area  of  base  x  height,  or,  &rW2  is  to  be  a  maximum 
when  the  length  +  the  perimeter  of  the  cylinder  =  6,  i.e.,  I  +  wD  =  6.  Ob- 
viously I  and  D  denote  the  respective  length  and  diameter  of  the  cylinder. 

(9)  Prove  that  the  sum  of  three  positive  quantities,  x,  y,  z,  whose  product 
is  constant,  is  greatest  when  those  quantities  are  equal.  Hint.  Let  xyz  =  a; 
x  +  y  +  z  =  u.  Hence  u  m  ajyz  +  y  +  g ;  .-.  T^u/dy  =  -  ajy2z  +  1=0; 
dufdz  =  -  ajyz2  +  1  =  0;  .-.  y  =  x,  u=*x\  .viayxf.  si  a.  To  show  that 
u  is  a  minimum,  note  'dhil'dx2  =  +  Sa/x*. 

III.  Functions  of  three  variables. 

Without  going  into  details  I  shall  simply  state  that  if  we  are 
dealing  with  three  variables  xy  y,  and  z,  such  that 


300  HIGHER  MATHEMATICS.  §  101. 

u=f{x,y,z),  ....  (12) 
there  will  be  a  maximum  or  a  minimum  if  the  first  partial  deriva- 
tives are  each  equal  to  zero;  and  Lagrange's  criterion,  uxxuyy>(uX3,)2, 
is  satisfied ;  and  if 

%*K*V**  +  2uvtuxguxy  -  unu\  -  u^u2^  -  uji2xy)>0.  (13) 
For  a  maximum  u^  will  be  negative,  and  positive  for  a  minimum. 
The  meaning  of  the  notation  used  will  be  understood  from  page 
19.     u^  =  ~dht/~dx2 ;  Uxy  =  ~d2u/bx'oy. 

Examples.— (1)  If  u  =  x2  +  y2  +  z2  +  x  -  2z  -  xy,  ux  =  2x  -  y  +  1  =  0 ; 
uv  =  2y  -  x\Va  =  2*  -  2  =  0 ;  .-.  aj  =  -|;y  =  -£;*  =  l;tt  =  -$.  uxx  =  2; 
uyy  =  2 ;  uu  =  2 ;  uxy  —  -  1 ;  uxz  =  0 ;  uyz  =  0.  Hence,  Lagrange's  criterion 
furnishes  +  3 ;  and  criterion  (13)  furnishes  2(8  +  0  -  0  -  0  -  2)  =  12.  Hence, 
since  u^  is  positive,  -  |  is  a  minimum  value  of  u. 

(2)  If  we  have  an  implicit  function  of  three  variables,  and  seek  the  maxi- 
mum value  of  say  z  in  u  =  2x2  +  by2  +  z2  -  ±xy  -  2x  -  fy  -  £  =  0,  we  proceed 
as  follows :  ux  =  4oj  -  4y  -  2  =  0 ;  wy  =  10y  -  4cc  -  4  =  0  ;  .*.  a:  =  f ;  2/  =  1 ; 
e  =  +  2.  tfe  =  2z  =  +  4  ;  n**  =  4 ;  wyi,  =  10 ;  uxy  =  -  4.  Lagrange's  criterion 
furnishes  the  value  40  -  16  =  24.  z  is  therefore  a  maximum  when  x  =  -% 
and  y  =  1. 

IF.  Conditional  Maxima  and  Minima. 

If  the  variables 

•    u-f(x,y,e)  =  0,        .         .         .         (H) 
are  also  connected  by  the  condition 

v  =  <£(z,  y,  z)  =  0,        .         .         .         (15) 
we  must  also  have,  for  a  maximum  or  a  minimum, 

'bu.         bu  ~ou 

-dx  +  ^dy  +  Tzdz  =  0.  .         .         (16) 

From  (15),  we  have  by  partial  differentiation 

7)v  bv  7)v 

Txdx  +  rydy  +  s*  - °-      •     •     <17> 

Multiply  (17)  by  an  arbitrary  constant  A,  called  an  undetermined 
multiplier,  and  add  the  result  to  (16). 

fin       x<toA_         fbu       Jbv\n         ftu       .M,        n       «„ 

(s +  As>fa  +  fe + v^ +  (»5 +  ^r - °-  <l8> 

But  X  is  arbitrary,  and  it  can  be  so  chosen  that 

bu   _    ^v 

~dx 
Substitute  the  result  in  (18),  and  we  obtain 


ox 


fill         x  "bv\  ,  /dw         .  c)t?\  ,  . 


§  102.  INFINITE  SERIES  AND  THEIR  USES.  301 

But  if  y  and  z  are  independent,  we  also  have 

Hence,  we  have  the  three  equations 

ox         ox  oy         oy  oz         oz 

together  with  </>(#,  y,  z)  =  0,  for  evaluating  x,  y,  z,  and  X.     This  is 
called  Lagrange's  method  of  undetermined  multipliers.    To 

illustrate  the  application  of  these  facts  in  the  determination   of 
maxima  and  minima,  let  us  turn  to  the  following  examples  : — 

Examples. — (1)  Find  the  greatest  value  of  7  =  8xyz,  subject  to  the  con- 
dition that  x2  +  y2  +  z2  =*  1.     By  differentiation, 

xdx  +  ydy  +  zdz  =  0 ;  and  yzdx  +  xzdy  +  xydx  =  0. 
For  a  maximum,  we  must  have 

ye  +  \x  =  0 ;  xz  +  \y  =  0 ;  xy  +  \z  =  0. 
Multiply  these  equations  respectively  by  x,  y,  and  z,  so  that 

xyz  +  \x2  -  0 ;  xyz  +  \y2  =  0 ;  xyz  +  \z*  =  0.     .        .        (19) 
By  addition 

Sxyz  +  \{x*  +  y2  +  s2)  =  0.    .-.  |F  +  \  =  0;  or,  \  =  -  f  7. 
Substitute  this  value  of  \  in  equation  (19),  and  we  get 

x  =  n/|;  y  =  n/|;  z  =  s/f;  .-.  7=  |  Jj. 

(2)  Find  the  rectangular  parallelopiped  of  maximum  surface  which  can 
be  inscribed  in  a  sphere  whose  equation  is  x2  +  y1  +  z2  =  r2.  The  surface  of 
the  parallelopiped  is  s  =  8(xy  +  xs  +  yz),  where  2x,  2y,  and  2z  are  the  lengths 
of  its  three  coterminous  edges.  By  differentiation,  xdx  +  ydy  +  zdz  =  0 ; 
(y  +  z)dx  +  (x  +  z)dy  +  (y  +  x)dz = 0.  For  a  maximum,  therefore,  y  +  z  +  xa? = 0 ; 
x+z+\y  =  0',  x  +  y  +  \e=0.  Proceed  as  before,  and  we  get  finally  x= y  =  z. 
Ansr.  Cube  with  edges  2x  =  1r  \/J.  • 

(3)  Find  the  dimensions  of  a  cistern  of  maximum  capacity  that  can  be 
formed  out  of  300  sq.  ft.  of  sheet  iron,  when  there  is  no  lid.  Let  x,  y,  z, 
respectively  =  length,  breadth  and  depth.  Then,  xy  +  2xz  +  2yz  =  300 ;  and, 
u  =  xyz,  is  to  be  a  maximum.  Proceed  as  before,  and  we  get  x  =  y  —  2z. 
Substitute  in  the  first  equation,  and  we  get  x  =  y  —  10,  *  =  5.  Hence  the 
cistern  must  be  10  ft.  long,  10  ft.  broad,  and  5  ft.  deep. 

§  102.    Lagrange's  Theorem. 

Just  as  Maclaurin's  theorem  is  a  special  case  -oi  Taylor's,  so 
the  latter  is  a  special  form  of  the  more  general  Lagrange's  theorem, 
and  the  latter,  in  turn,  a  special  form  of  Laplace's  theorem.  There 
is  no  need  for  me  to  enter  into  extended  details,  but  I  shall  have 
something  to  say  about  Lagrange's  theorem. 

If  we  have  an  implicit  function  of  three  variables, 

z  =  y  +  x<f>(z),     ....         (1) 


302  HIGHER  MATHEMATICS.  §  102. 

such  that  x  and  y  have  no  other  relation  than  is  given  by  the 
equation  (1),  each  may  vary  independently  of  the  other.  It  is 
required  to  develop  another  function  of  z,  say  f(z),  in  ascending 
powers  of  x.     Let 

then,  by  Maclaurin's  theorem, 


u  =  un  + 


(du\  x      (<Pu\x*_      (dhAa^ 
\dx)0l  +  \dafl)02  !  +  \dafi)0S  !  +  *  *  * 


Without  going  into  details,  it  is  found  that  after  evaluating  the 
respective  differential  coefficients  indicated  in  this  series  from  (1), 
we  get  as  a  final  result 

/(^)=/(,)  +  ^(,)^|[f-V)}2K-••.    m 

which  is  known  as  Lagrange's  theorem.  The  application  of  this 
series  to  specific  problems  is  illustrated  by  the  following  set  of 
examples : — 

Examples.— (1)  Given  a  -  by  +  cy*  =  0,  find  y. 
Rearranging  the  given  equation,  we  get 

a      c  » 

v  =  i  +  iy*-> (3) 

and  on  comparing  this  with  the  typical  equations  (1)  and  (2),  we  have 

/(*)  =  v*  •••  f(y)  =  * ;  *(*)  =  y2,  <t>(y)  =  *2 ;  *  =  «/& ;  «  =  «/&. 

From  (2),  df(y)/dy  =  1 ;  de/dg  =  1,  etc.,  *  of  (1)  is  y  of  (3),  therefore, 

y        +*1+    dz     2t  +    dz*     3!  +  •••' 
...  2/  =  s  +  *^  +  4^_  +  6.  5«*gf  +  ... ; 

_a4.i8   .cxi   i8    ^      6.5    a4    c3 
•'*  y~b  +  &2'6  +  2!'  b*'b*  +    3!   'o4'&8  +  ,,,; 
_a(        ac       4    a2c2      6 . 5    ^c3  \ 

•'•  y  -  b\l  +  P  +  21*  "W  +  "ST  "W  +  ' ' ')' 
a  series  which  is  identical  with  that  which  arises  when  the  least  of  the  two 
roots  of  equation  (3)  is  expanded  by  Taylor's  theorem. 

(2)  Given  yz  -  ay  +  b  =  0,  find  y".    On  comparing  the  given  equation 

y=a+a^ 
with  the  typical  forms,  we  see  that 

f{*)  =  y",  .-./(y)  =  «M;  *(*)  =  0*.  •*•<*>&)  =  *s;  *  =  &/a ;  «  =  i/a. 

,    «      ^(tt^-1*6)    a;2 
...J.-I-  +  W-VJ  + ^i ^ '  21  +  •  •  •  ; 


hl    1      ^(n  +  5)    6^    1 
1  +  V '  a  +       21       '  a*'  a*  + 


§  102.  INFINITE  SEKIES  AND  THEIK  USES.  303 

(3)  In  solving  the  velocity  equations 

dx  dt 

■jt  =  k.ia  -  x)  (a  -  x  -  f) ;  ^  =  k2(a  -  x)  {a  -  x  -  |), 

for  the  reaction  between  propyl  iodide  and  sodium  ethylate,  W.  Hecht,  M. 
Conrad,  and  0.  Bruckner  (Zeit.  phys.  Ghem.,  1,  273,  1889)  found  that  by 
division  of  the  two  equations,  and  integration, 


H1-!)' 


where  a  denotes  the  amount  of  substance  at  the  beginning  of  the  reaction ; 
x  and  |  are  the  amounts  decomposed  at  the  time  t ;  K  =  J^fk^ ;  when  t  =  0, 
x  =  0,  and  £  =  0.  If  K  is  small,  Maclaurin's  theorem  furnishes  the  expres- 
sion 

If  we  put  x  +  £  =  y,  we  can  get  a  straightforward  relation  between  y  and  t ; 
for  obviously, 

Zfc  +  t-VJ  (l  +  *)t-V;  .'.  (1  +  K)dl  =  dy; 
dt  dy 

*  '*  dt  =  k^a  ~  ®  ^a  "  x  ~  ® '  becomes  di  =  k\{aK  +  a  -  ?/)  (a  -  y), 

which  can  be  integrated  in  the  ordinary  way.  But  K  was  usually  too. large 
to  allow  of  the  approximation  (4).  We  have  therefore  to  solve  the  problem : 
Given 

l-?=l-«'  +  i  =  ('l-lV,flndi. 
a  a     a      \        a)  a 

For  the  sake  of  brevity  write  this : 

1-3  +  3=  (1-  e)*,  .%  1  -  x  +  M  -1  ■*  Km  +  %K(K  -  l)z*  -  . . . ; 

,:x=(K+  l)z  -  E{K-  l)*2+...=/i(*).         .        .        (5) 

.-.**  =  */!(*);  .:M  =  Xj^j  =  x<p{z).    ...        (6) 

On  referring  to  the  fundamental  types  (1)  and  (2),  we  see  that 

f{z)  =  zj(y)  =  y;  ${z)  -  <f>{z),  <p{y)  =  <p(y) ;  y  =  0,a?  =  2/; 

We  must  now  evaluate  the  separate  terms. 

|™  =  1  =  1 <«> 

From  (5)  and  (6), 

since  y  =  0;  again,  from  (5),  (6),  (7),  and  (8), 

|C1^W]=  §WW  =|{(z  +  i)-^-i)y  +  ...}2 

_  2K(K-1) K(K-1) 

-{(K+l)-$K(K-l)y +  ...}*       (K+l)*>      '        '        <10> 

sinoe  y  is  zero.     Hence,  the  required  development,  from  (7),  is 


304  HIGHER  MATHEMATICS.  §  103. 

-      1        x  ,1    K(K  ~  *)    (*\* 
Z~  K+l' 1+ 2'   (K+l)s'\2\)   + (U) 

We  have  put  z  for  |/a,  and  x  for  y/a.  On  restoring  the  proper  values  of  z  and 
x  into  the  given  velocity  equations,  we  can  get,  by  integration,  a  relation 
between  if,  t,  and  constants. 

§  103.    Functions  requiring  special  Treatment  before 
Substituting  Numbers. 

In  discussing  the  velocity  of  reactions  of  the  second  order,  we 
found  that  if  the  concentration  of  the  two  species  of  reacting 
molecules  is  the  same,  the  expression 

,'  1     ,     a  -  x   a 

assumes  the  indeterminate  form 

kt  =  oo  x  0, 
by  substituting  a  =  b.  We  are  constantly  meeting  with  the  same 
sort  of  thing  when  dealing  with  other  functions,  which  may  re- 
duce to  one  or  other  of  the  forms :  £,  -gj,  oo  -  go,  l00,  oo°,  0°  . . . 
We  can  say  nothing  at  all  about  the  value  of  any  one  of  these 
expressions,  and,  consequently,  we  must  be  prepared  to  deal  with 
them  another  way  so  that  they  may  represent  something  instead 
of  nothing.  They  have  been  termed  illusory,  indeterminate  and 
singular  forms.  In  one  sense,  the  word  ''indeterminate"  is  a 
misnomer,  because  it  is  the  object  of  this  section  to  show  how 
values  of  such  functions  may  be  determined. 

Sometimes  a  simple  substitution  will  make  the  value  apparent 
at  a  glance.     For  instance,  the  fraction  (x  -f  a)/(x  +  b)  is  inde- 
terminate when  x  is  infinite.     Now  substitute  x  =  y  ~  1  and  it  is 
easy  to  see  that  when  x  is  infinite,  y  is  zero  and  consequently, 
x  +  a  1  +  ay  _ 

U*=°>x~Tb  =  -^oiTty  "  l' 

Fractious  which  assume  the  form  $  are  called  vanishing  frac- 
tions, thus,  (x2  -  &x  +  S)/(x2  -  1)  reduces  to  g,  when  x  =  1.  The 
trouble  is  due  to  the  fact  that  the  numerator  and  denominator 
contain  the  common  factor  (x  -  1).  If  this  be  eliminated  before 
the  substitution,  the  true  value  of  the  fraction  for  x  =  1  can  be 
obtained.     Thus, 

x\  -  ±x  +  3  _  (x  -  1)  (x  -  3)  =  x  -  3       _  2      _ 
'  x2  -  1       ~  (x  -  1)  (x  +  1)  ~  x  +  1  ~      2  " 


§  103.  INFINITE  SERIES  AND  THEIR  USES.  305 

These  indeterminate  functions  may  often  be  evalued  by  alge- 
braic or  trigonometrical  methods,  but  not  always.  Taylor's  theorem 
furnishes  a  convenient  means  of  dealing  with  many  of  these  func- 
tions. The  most  important  case  for  discussion  is  "  $,"  since  this 
form  most  frequently  occurs  and  most  of  the  other  forms  can  be 
referred  to  it  by  some  special  artifice. 

I.  The  function  assumes  the  form  J. 

This  form  is  the  so-called  vanishing  fraction.  As  already 
pointed  out,  the  numerator  and  denominator  here  contain  some 
common  factor  which  vanishes  for  some  particular  value  of  x,  say. 
These  factors  must  be  got  rid  of.  One  of  the  best  ways  of  doing 
this,  short  of  factorizing  at  sight,  is  to  substitute  a  +  h  for  x  in  the 
numerator  and  denominator  of  the  fraction  and  then  reduce  the 
fraction  to  its  simplest  form.  In  this  way,  some  power  of  h  will 
appear  as  a  common  factor  of  each.  After  reducing  the  fraction 
to  its  simplest  form,  put  h  =  0  so  that  a  =  x.  The  true  value  of 
the  fraction  for  this  particular  value  of  the  variable  x  will  then  be 
apparent. 

For  oases  in  which  x  is  to  be  made  equal  to  zero,  the  numerator 
and  denominator  may  be  expanded  at  once  by  Maclaurin's  theorem 
without  any  preliminary  substitution  for  x.  For  instance,  the  trig- 
onometrical function  (sin  x)/x  approaches  unity  when  x  converges 
towards  zero.  This  is  seen  directly.  Develop  sin  x  in  ascending 
powers  of  x  by  Taylor's  or  Maclaurin's  theorems.     We  thus  obtain 


x      x%      x5      x 


sins      VI      3!  +  5!       71  +  '")  a?     !*_*?_ 

x  x  -  1  ~  3!  +  5!      7!  +  '*' 

The  terms  to  the  right  of  unity  all  vanish  when  x  =  0,  therefore, 

smx 
lA  =  o— £~  -  !• 

Examples. — (1)  Show  Ltxa,0(o?  -  b^jx  =  log  a/b. 

(2)  Show  Lt*  =  0(1  -  cos  x)/x2  =  $. 

(3)  The  fraction  (xn  -  an)/(x  -  a)  becomes  #  when  x  =»  a.    Put  x  =  a  +  h 
and  expand  by  Taylor's  theorem  in  the  usual  way.     Thus, 

x»  -  an                  (a  +  h)n  -  an 
Lt*-1TI  =  Lt*  =  « % ■*-*. 

It  is  rarely  necessary  to  expand  more  than  two  or  three  of  the  lowest  powers 

of  h.    The  intermediate  steps  are 

_         {an  +  naP-^h  +  %n(n  -  l)an  -  2h*  +  ...)-  a1* 

Lt*  -  o  a+h-a * 

U 


306  HIGHER  MATHEMATICS.  §  103. 

Cancel  out  an  in  the  numerator,  and  a  in  the  denominator ;  divide  out  the  fe'e 
and  put  h  =  0. 

(4)  The  velocity,  V,  of  a  body  falling  in  a  resisting  medium  after  an 
interval  of  time  t,  is 

K~/B  fgftt  +  i*  "  v ~gt> 
when  the  coefficient  of  resistance,  fi,  is  made  zero.     Hint.  Expand  the  numer- 
ator only  before  substituting  £  =  0. 

(5)  Show  that  LtA  =  0^log  ( 1  +  -  J  =  -,  as  on  page  51. 

(6)  If  H  denotes  the  height  which  a  body  must  fall  in  order  to  acquire  a 
velocity,  V,  then 

where  k  is  the  coefficient  of  resistance.     If  k=0,  show  that  H  =  %V2lg. 
We  can  generalize  the  preceding  discussion.     Let 

Obviously, 

/i(«)=/2(«)  =  0.  ...        (2) 

Expand  the  two  given  functions  by  Taylor's  theorem, 

/i(*  +  h)     f1(*)+f1'(x)h  +  jfl"(x)W  +  ...  > 

f2(x  +  h)     f2(x)  +  f2'(x)h  +  W(x)h*  +  . . .'  •         W 
Now  substitute  x  =  a,  and  fx(a)  =  f2(a)  =  0  as  in  (2) ;  divide  by 
h;  and 

/i(«  +  *)  _  A'(a)  +  ¥i"m  +  ...  ... 

Ma  +  h)-fi(a)  +  if2"(a)h+...      '.  '         W 
remains. 

^(q  +  h)  _             f1,(a)  +  if1"(a)h  +  ...  ^ 

^4R-m    andLt,.^=Lt,  =  /^).  (6) 

In  words,  if  the  fraction  fi(x)/f2(x)  becomes  £,  when  a;  =  a,  the 
fraction  can  be  evaluated  by  dividing  the  first  derivative  of  the 
numerator  by  the  first  derivative  of  the  denominator,  and  sub- 
stituting x  =  a  in  the  result.  This  leaves  us  with  three  methods 
for  dealing  with  indeterminate  fractions. 

1.  Division  Method. — i.e.,  by  dividing  out  the  common  factors. 

2.  Expansion  Method. — i.e.,  by  substituting  x  +  h  f or  x,  etc. 

3.  Differentiation  Method. — i.e.,  by  the  method  just  indicated. 

fdx 
Examples. — (1)  Prove  that  /  — -  =  log  x,  by  means  of  the  general  formula 


§  103.  INFINITE  SERIES  AND  THEIR  USES.  307 


i- 


xndx  =  ^j.     Hint.  Show  that 

xn  +  1                  sen  +  1  log  x  .  dn 
Ltn=-%  +  i =  Lt«=-i dTi =  Lt«=-i*n  +  1log*  =  log*. 

by  differentiating  the  numerator  and  denominator  separately  with  regard  to 
n  and  substituting  n  =  -  1  in  the  result. 

(2)  Show  Lty«1^r(^rT  -  ^zi)  =  clog^.  See  (10),  p.  259,  (3), 
p.  264. 

II.  The  function  assumes  the  form  ~. 
Functions  of  this  typeoan  be  converted  into  the  preceding  "  J" 
case  by  interchanging  the  numerator  and  denominator,  but  it  is 
not  difficult  to  show  that  (6)  applies  to  both  £  and  to  ^  ;  and 
generally,  if  the  ratio  of  the  first  derivatives  vanishes,  use  the 
second  ;  if  the  second  vanishes,  use  the  third,  etc.    Or,  symbolically, 

r,    fM    Tt    *M    n    f^-    rt     '*?&■_     m 

This  is  the  so-called  rule  of  l'Hopital. 

log  X  x~l 

Examples.— (1)   Show  that  Lt* m  0 -^rj  =  Lt^  =  0  _  x_2  =  Ltx  _  0  - x  =  0. 

(2)  The  nth  derivative  of  x"  is  n !  and  the  nth  derivative  of  ex  is  ex  by 
Leibnitz'  theorem,  when  n  is  positive.     Hence  show  that 

e*  e* 

Lt*=  oo^,  =  Lt*  -  oo!  #  2  ...  n  =  CD' 

III.  The  function  assumes  the  form  oo  x  0. 
Obviously  ^  such  a  fraction  can  be  converted  into  the  "  J  "  form 
by  putting  the  infinite  expression  as  the  denominator  of  the  fraction ; 
or  into  the  ~  form  by  putting  the  zero  factor  as  the  denominator 
of  the  fraction  as  shown  in  the  subjoined  examples. 

Examples.— (1)  The  reader  has  already  encountered  the  problem:  what 
does  x  log  x  become  when  x  =  0  ?  We  are  evidently  dealing  with  the  0  x  oo 
case.    Obviously,  as  in  a  preceding  example, 

Ltx  =  0a;loga;  =  Lt^^o-y— =  0. 
x 
(2)  Show  Lta  =  b~  log  |j^-||  =  ^T^y  as  indicated  on  page  220. 
(8)  Show  Lt*  =  O30  -  *log  x  =  0  x  oo  =  0. 

IV.  The  function  assumes  the  form  oo  -  oo,  or  0  -  0. 

First  reduce  the  expression  to  a  single  fraction  and  treat  as 
above. 


308  HIGHER  MATHEMATICS.  §  104. 

Examples.— (1)  Show  by  differentiating  twice,  etc.,  that 

t  f  x  1         T.       a>  log  a  -  <e  +  1  x  1 

**-*£  -  i     iogaJ  =  ljt*  =  i  («-i)iog«    =  ^ST*""*  etc* 

(2)  Show  that  Lt*-ir-^ r—-  =  1. 

v  '  x~l\ogx      logo; 

(3)  W.  Hecht,  M.  Conrad  and  6.  Bruckner  {Zeit.  phys.  Chem.,  4,  273, 1889) 
wanted  the  limiting  value  of  the  following  expression  in  their  work  on 
chemical  kinetics: — 


Ltw 


1     /      na  -  x  \ 

l^li1°S'^7-^wJ.     Ansr. 


V.  The  function  assumes  one  of  the  forms  l00,  oo°,  0°. 

Take  logarithms  and  the  expression  reduces  to  one  of  the 
preceding  types. 

Examples.— (1)  Lt^  _  Qxx  =  0°*  Take  logs  and  noting  that  y  =  &*, 
l°g  V  =  &  l°g  3-  But  we  have  just  found  that  x  log  x  =  0  when  x  =  0 ; 
.♦.  log  y  =  0  when  x  =  0 ;  .-.  y  =  1.     Hence,  ~Ltx  =  0xx  =  1. 

(2)  Show  Ltx  =  0(l  +  mx)llx  =  1"  =  em.  Here  log  y  =  a;-1  log  (1  +  wkb). 
.  But  when  x  =  0,  log  2/  =  # ;  by  the  differential  method,  we  find  that 
log(l  +  mx)x  ~ x  becomes  m  when  x  =  0 ;  hence  log  y  =  m ;  or  y  =  em,  when 
a  =  0. 

§  104.    The  Calculus  of  Finite  Differences. 

The  calculus  of  finite  differences  deals  with  the  changes  which 
take  place  in  the  value  of  a  function  when  the  independent  variable 
suffers  a  finite  change.  Thus  if  x  is  increased  a  finite  quantity  h, 
the  function  x2  increases  to  (x  +  h)2,  and  there  is  an  increment  of 
(x  +  h)2  -  x2  =  2xh  +  h2  in  the  given  function.  The  independent 
variable  of  the  differential  calculus  is  only  supposed  to  suffer  in- 
finitesimally  small  changes.  I  shall  show  in  the  next  two  sections 
some  useful  results  which  have  been  obtained  in  this  subject ; 
meanwhile  let  us  look  at  the  notation  we  shall  employ. 

In  the  series 

1»,  23,  3»,  43,  53 

subtract  the  first  term  from  the  second,  the  second  from  the  third, 
the  third  from  the  fourth,  and  so  on.     The  result  is  a  new  series, 

7,  19,  37,  61,  91, . . . 
called  the  first  order  of  differences.     By  treating  this  new  series 
in  a  similar  way,  we  get  a  third  series, 

12,  18,  24,  30, ... , 
called  the  second  order  of  differences.     This  may  be  repeated 


§104. 


INFINITE  SERIES  AND  THEIR  USES. 


309 


as  long  as  we  please,  unless  the  series  terminates  or  the  differ- 
ences become  very  irregular. 

The  different  orders  of  differences  are  usually  arranged  in  the 
form  of  a  "  table  of  differences".  To  construct  such  a  table,  we 
can  begin  with  the  first  member  of  series  of  corresponding  values 
of  the  two  variables.  Let  the  different  values  of  one  variable, 
say,  x0,  xv  x2,  . .  .  correspond  with  y0,  yv  y2, .  . .  The  differences 
between  the  dependent  variables  are  denoted  by  the  symbol 
"  A,"  with  a  superscript  to  denote  the  order  of  difference,  and  a 
subscript  to  show  the  relation  between  it  and  the  independent 
variable.     Thus,  in  general  symbols  : — 


Argument. 

Function. 

Orders  of  Differences. 

First. 

Second. 

Third. 

Fourth. 

Xa  +  2h 
Xa  +  ih 

Va 

Va  +  2h 

A1, 

A  a  +  2* 
*'.  +  » 

A*a 

A2a  +  2ft 

A3a 
A3a  +  A 

a*. 

If  we  apply  this  to  the  function  y 
differences : 

x.  y.  Ai.  A». 


x3  we  get  the  table  of 


A3. 


A«. 


(6)o 

(6)! 


0 


(8).        (19°        (12)° 
(27),         J?1        (18), 

(64)3         'J2        (2*), 
(125)4        (bl)s 

Such  a  table  will  often  furnish  a  good  idea  of  any  sudden  change 
which  might  occur  in  the  relative  values  of  the  variables  with  a 
view  to  expressing  the  experimental  results  in  terms  of  an  em- 
pirical or  interpolation  formula.  It  is  not  uncommon  to  find 
faulty  measurements,  and  other  mistakes  in  observation  or  calcula- 
tion, shown  up  in  an  unmistakable  manner  by  the  appearance  of 
a  marked  irregularity  in  a  member  of  one  of  the  difference  columns. 
It  is,  of  course,  quite  possible  that  these  irregularities  are  due  to 
something  of  the  nature  of  a  discontinuity,  in  the  phenomenon 
under  consideration. 


310 


HIGHER  MATHEMATICS. 


§105. 


§  103.    Interpolation. 

In  one  method  of  fixing  the  order  of  a  chemical  reaction  it  is 
necessary  to  find  the  time  which  is  required  for  the  transformation 
of  equal  fractional  parts  of  a  given  substance  in  two  separate 
systems.  Let  x  denote  the  concentration  of  the  reacting  sub- 
I tance  at  the  time  t ;  a,  the  initial  concentration  of  the  reacting  sub- 
stance ;  and  suppose  that  the  following  numbers  were  obtained : — 


System  I.,  a  =  0-1. 

System  11.,  a  =  0-0625. 

t. 

X. 

t. 

X. 

1-0 
1-5 
2-5 
4-0 

0-0581 
0-0490 
0-0382 
0-0300 

1 

3 

7 

17 

0-04816 
0-03564 
0-02638 
0-01748 

In  one  system,  T3^ths  of  the  substance  were  transformed  in  four 
minutes ;  it  is  required  to  find  in  what  time  y^ths  of  the  reacting 
substance  were  transformed  in  the  second  system. 

Expressed  in  more  general  terms,  given  a  definite  number  of 
observations,  to  calculate  any  required  intermediate  values.  This 
kind  of  problem  frequently  confronts  the  practical  worker,  par- 
ticularly in  dealing  with  isolated  and  discontinuous  observations, 
and  measurements  in  which  time  is  one  of  the  variables.  The 
process  of  computation  of  the  numerical  values  of  two  variables 
intermediate  between  those  actually  determined  by  observation  and 
measurement,  is  called  interpolation.  When  we  attempt  to 
obtain  values  lying  beyond  the  limits  of  those  actually  measured, 
the  process  is  called  extrapolation.  The  term  "interpolation" 
is  also  applied  to  both  processes.     See  page  92. 

It  is  apparent  that  the  correct  formula  connecting  the  two  vari- 
ables must  be  known  before  exact  interpolation  can  be  performed. 
If  the  relation  between  the  mass,  m,  and  volume,  v,  of  a  substance 
be  represented  by  the  expression  m  =  %v,  we  can  readily  calculate 
the  value  of  m  for  any  value  of  v  we  might  desire.  Moreover,  the 
method  of  testing  a  supposed  formula  is  to  compare  the  experi- 
mental values  with  those  furnished  by  interpolation.  Interpolation, 
therefore,  is  based  on  the  assumption  that  when  a  law  is  known 


§  105.  INFINITE  SERIES  AND  THEIR  USES.  311 

with  fair  exactness,  we  can,  by  the  principle  of  continuity,  antici- 
pate the  results  of  any  future  measurements. 

When  the  form  of  the  function  connecting  the  two  variables  is 
known,  the  determination  of  the  value  of  one  variable  correspond- 
ing with  any  assigned  value  of  the  other  is  simple  arithmetic. 
When  the  form  of  the  function  is  quite  unknown,  and  the  definite 
values  set  out  in  the  table  alone  are  known,  the  problem  loses  its 
determinate  character,  and  we  must  then  resort  to  the  methods  now 
to  be  described. 

I.   Interpolation  by  proportional  parts. 

If  the  differences  between  the  succeeding  pairs  oi  values  are 
small  and  regular,  any  intermediate  value  can  be  calculated  by 
simple  proportion  on  the  assumption  that  the  change  in  the  value 
of  the  function  is  proportional  to  that  of  the  variable.  This  is 
obviously  nothing  more  than  the  rule  of  proportional  parts  illus- 
trated on  page  289,  by  the  interpolation  of  log  (n  +  h)  when  log  n 
and  log  (n  +  1)  are  known.  The  rule  is  in  very  common  use.  For 
example,  weighing  by  the  method  of  vibrations  is  an  example  of 
interpolation.  Let  x  denote  the  zero  point  of  the  balance,  let  wQ 
be  the  true  weight  of  the  body  in  question.  This  is  to  be  measured 
by  finding  the  weight  required  to  bring  the  index  of  the  balance  to 
zero  point.  Let  x1  be  the  position  of  rest  when  a  weight  wx  is 
added  and  x2  the  position  of  rest  when  a  weight  w2  is  added.  As- 
suming that  for  small  deflections  of  the  beam  the  difference  in  the 
two  positions  of  rest  will  be  proportional  to  the  difference  of  the 
weights,  the  weight,  tvQ,  necessary  to  bring  the  pointer  to  zero  will 
be  given  by  the  simple  proportion  : 

K  -  v>x)  :  (x0  -  xY)  =  (w2  -  wj  :  (x2  -  xx). 

When  the  intervals  between  the  two  terms  are  large,  or  the 
differences  between  the  various  members  of  the  series  decrease 
rapidly,  simple  proportion  cannot  be  used  with  confidence.  To  take 
away  any  arbitrary  choice  in  the  determination  of  the  intermediate 
values,  it  is  commonly  assumed  that  the  function  can  be  expressed 
by  a  limited  series  of  powers  of  one  of  the  variables.  Thus  we  have 
the  interpolation  formulae  of  Newton,  Bessel,  Stirling,  Lagrange, 
and  Gauss. 

II.  Newton's  interpolation  formula. 
Let  us  now  return  to  fundamentals.     If  yx  denotes  a  function  of 
x,  say 


312  HIGHER  MATHEMATICS.  §  105. 

y*  =  /0»)> 

then,  if  x  be  increased  by  h, 

yx  +  h=f(x  +  h), 
and  consequently, 

Increment  yx  =  yx  +  h  -  yx  =  f(x  +  h)  -  f{x)  =  A1,.  (1) 

Similarly,  the  increment 

^yx^A\  +  h-  A*.  -  Ai(Ay  -  A»„  .  .  (2) 
where  A1*  is  the  first  difference  in  the  value  of  yx,  when  x  is  in- 
creased to  x  +  h ;  A2X  is  the  first  difference  of  the  first  difference 
of  yx,  that  is,  the  second  difference  of  yx  when  x  is  increased  to 
x  +  In.  It  will  now  be  obvious  that  A1  is  the  symbol  of  an  opera- 
tion— the  taking  of  the  increment  in  the  value  of  f{x)  when  the 
variable  is  increased  to  x  +  h.  For  the  sake  of  brevity,  we  gener- 
ally write  Ax  for  Ayx.     From  (1)  and  (2),  it  follows  that 

yx  +  k  =  yx  +  ^1yx;        ...       (3) 
y*+»  -  y«+»  +  A1y*+*  -  y«  +  Aty,  +  a1^,  +  a1^)  ; 

.-.  &+*  ->*  +  SA1^  +  A*y„  .         .         (4) 

Similarly, 

?.  +  »  =  S/*  +  3Aiy,  +  3A*y,  +  A«y„  .  .  (5) 
We  see  that  the  numerical  coefficients  of  the  successive  orders  of 
differences  follow  the  binomial  law  of  page  36.  This  must  also  be 
true  of  yx  +  „*  if  n  is  a  positive  integer,  consequently, 

?«  +  .»  =  yx  +  nAyx  +  %n{n  -  1)A2^  +  . . . 
This  is  Newton's  interpolation  formula  (Newton's  Principia,  3, 
lem.  5,  1687)  employed  in  finding  or  interpolating  one  or  more 
terms  when  n  particular  values  of  the  function  are  known.     Let  us 
write  y0  in  place  of  yx  for  the  first  term,  then 

n(n  -  1)   .        n(n  -  1)  (n  -  2)   _ 
**  -  y,  +  *A»,  +  -^-Jf—  A2o  +  ^ if 1*\  +■■■    (6) 

continued  until  the  differences  become  negligibly  small  or  irregular. 
If  we  write  nh  =  x,  n  —  x/h,  and  (6)  assumes  the  form 

«      «  x*    ^.^-^   A»0    a?(a?-A)(g-afc)    A3 

^=^o  +  ^x  +  — F"~'  2T+  P  •"3l+---    ") 

where  h  denotes  the  increment  in  the  successive  values  of  the  inde- 
pendent variable ;  and  x  is  the  total  increment  of  the  interpolated 
term.     The  application  is  best  illustrated  by  example. 


§105. 


INFINITE  SERIES  AND  THEIR  USES. 


313 


Examples.— (1)  If  y0  =  2,844  ;  yl  =  2,705 ;  y2  =  2,501 ;  ys  =  2,236,  find  yx 
(Inst,  of  Actuaries  Exam.,  1889).  First  set  up  the  difference  table,  paying 
particular  attention  to  the  algebraic  signs  of  the  differences. 


X. 

y- 

Al. 

A«. 

A». 

0 
1 
2 
8 

2,844 
2,705 
2,501 
2,236 

(-  139)0 

-  204K 

(-  265)2 

(-65)0 
(-61), 

(+4)o 

Now  substitute  these  values  in  (6)  or  (7),  h  =  1,  and  if  n  =  £,  we  have  x  »•  \. 

...  nmv,  +  w,  +  tJ,,  +  Hi  -  i)  tt  -  \v 

.%  y^  =  2844  -  *  x  189  +  ^   x  65  +  ^  x  4  =  2821-592. 

(2)  The  amount  of  £1  in  50  years  at  2£  °/0«  3-4371090 ;  at  3  °/0  =  4-3839061 ; 
at  3£  °/0  =  5-5849264 ;  at  4  %  =  7-1066845  (Inst.  Actuaries  Exam.,  1888). 
Find  the  amount  at  3f°/0.  Here  A\  =  0-9467971 ;  A20  =  0-2542232 ; 
A80  =  0-0665146 ;  y0  =  8-4371090 ;  let  y0,  y2,  yAt  and  yQ  denote  the  respective 
values  of  y  here  given  ;  h'=  2.     Required  y6. 

m  .        5.3   „       5.3.1   , 
.'.  Vs  =  2/o  +  i*\  +  -Q- Aao  +  — is- A*o. 

.•.  y  =  3-4371090  +  2-3669928  +  0-4766685  +  0-0207858  =  6-3015561. 
The  correct  value  is  6*80094.  The  discrepancy  is  due  to  the  fact  that  the 
order  of  difference  above  the  third  ought  not  to  be  neglected.  But  we  can 
only  get  n  -  1  orders  of  differences  from  n  consecutive  terms  and  equidistant 
terms  values  of  a  function.  If  more  terms  had  been  given  we  could  have 
got  a  more  exact  result. 

(3)  Given  yQ  =  89,685 ;  y1  =  88,994 ;  y2  =  88,294 ;  ys  =  87,585,  find  y» 
(Inst/ Actuaries  Exam.,  1902).  Here  a1,,  =  -  691 ;  A20  =  -  9  ;  the  succeeding 
differences  A8,  A4,  . . .  are  all  zero.     Here  (6)  becomes 

y»  =  Vo  +  9^1o  +  36A2o  =  89,685  -  6,219  -  324  =  83,142. 

(4)  Given  log  4-22  =  0-6253125  ;  log  4-23=0-6263404 ;  log  4-24=0-6273659; 
log  4-25  m  0-6283889,  find  log  4-21684.  Here  y0t  ylt  ya,  y%  denotes  the  given 
quantities,  we  want  yx  =  2/_<)-oo3i6  ;  ^  =  0-01.     Hence,  from  (7), 


V  -  0*00316  —  Vo 


0-00316   A*0      0-00316(- 0-00316  -  1)    A2^ 
0-0001  '  2  I  ' 

06249872. 


O'Ol        1 
-  06253125  -  0-0003248  -  0-0000005 
(5)  What  is  the  cube  root  of  60-25,  given  the  cube  root  of  60  =  3-914868  ; 
3-957891 ;    63   =  3-979057 ;    64  =  4-000000  ?    Here, 


61  =  3-936497;    62 


A1,,  =  +  0-021629  ;  A20  =  -  0-000235.     Substitute  x  =  J  ;  fc  =  1. 

•*•  y  =  Vo+  $*\  -  ^5A20  =  3-914868  +  0*005407  +  0-000022  =  3-920297. 
The  number  obtained  by  simple  proportion  is  3*920295.     The  correct  number 
is  a  little  greater  than  3-920297. 


314  HIGHER  MATHEMATICS.  §  105. 

III.  Lagrange's  interpolation  formula. 

We  have  assumed  that  the  n  given  values  are  all  equidistant. 
This  need  not  be.  A  new  problem  is  now  presented :  Given  n 
consecutive  values  of  a  function,  which  are  not  equidistant  from 
one  another,  to  find  any  other  intermediate  value. 

Let  y  become  ya,  yb,  yc1 . . .  yn  when  x  becomes  a,  b,  c, . . .  n. 
Lagrange  has  shown  that  the  value  of  y  corresponding  with  any 
given  value  of  x,  can  be  determined  from  the  formula 

_ (x - b)  (x - c) . . .  (x - n)        (x-a)(x-c). .  .(x-ri) 
Vx~ (a-b)  (a  -o) . .  .(a-nfa  +  (b-a){b-c) . .  .(b-nfb  + ' '  "     <8' 

where  each  term  is  of  the  nth.  degree  in  x.  This  is  generally  known 
as  Lagrange's  interpolation  formula,  although  it  is  said  to  be 
really  due  to  Euler. 

Examples. — (1)  Find  the  probability  that  a  person  aged  53  will  live  a 
year  having  given  the  probability  that  a  person  aged  50  will  live  a  year 
=  0-98428 ;  for  a  person  aged  51  =  0-98335 ;  54,  0-98008 ;  55,  0;97877  (Inst. 
Actuaries  Exam.,  1890).  Here,  ya  =  0*98428,  a  =  0 ;  yb  =  0-98335,  6  =  1; 
yc  =  0-98008,  c  =  4  ;  ya  =  0-97877,  d  =  5  ;  .*.  x  =  3. 

(a  -  b){x  -  c)  {x  -  d)  =  (3  -  1)  (3  -  4)  (3  -  5)  =  +  4 ; 

(x  -  a)  {x  -  b)  (x  -  d)  =  (3  -  0)  (3  -  4)  (3  -  5)  =  +  6 ; 

(x  -  a)  (x  -  b)  {x  -  d)  =  (3  -  0)  (3  -  1)  (3  -  5)  =  -  12 ; 

(x  -  a)  (x  -  b)  {x  -  c)  =  (3  -  0)  (3  -  1)  (3  -  4)  =  -  6 ; 

(a  _  b)  {a  -  c)  (a  -  d)  =  (0  -  1)  (0  -  4)  (0  -  5)  =  -  20 ; 

(5  _  a)  (b  -  c)  (b  -  d)  =  (1  -  0)  (1  -  4)  (1  -  5)  =  +  12 ; 

(c  -  a)  (c  -  b)  (c  -  d)  =  (4  -  0)  (4  -  1)  (4  -  5)  =  -  .12 ; 

{d  -  a)  {d  -  b)  (d  -  c)  =  (5  -  0)  (5  -  1)  (5  -  4)  =  +  20. 

4          6         12         6             0-98428    098335    0-98008    3x0-97877 
V*=  -20ya  +  12yb  +  T2lJe  ~  20^=  ~       5~~~ + 2~  +  "~1 10^ ; 

yx  =  -  0-196856  +  0-491675  +  0-98008  -  0-29361 =0-98127. 

(2)  Given  log 280  =  2-4472 ;  log 281  =  2-4487 ;  log 283  =  2-4518 ;  log 286  =  2-4564, 
find  log  282  by  Lagrange's  formula  (Inst.  Actuaries  Exam.,  1890).  x  =  2, 
a  =  0,  b  —  1,  c  =  3,  d  =  6.    Hence  show  that 

yx  =  -  &/«  +  iyb  +  iye  -  ?\ya  =  2-4502. 

(3)  Find  by  Lagrange's  formula  log  x  =  2£,  given  log  200  =  2*30103 ; 
log  210  =  2-32222 ;  log  220  =  2-34242 ;  log  230  =  2-36173  (Inst.  Actuaries 
Exam.,  1891).  Here  a  =  0,  b  =  1,  c  =  2,  d  =  3.  Substitute  in  the  interpo- 
lation formula  and  we  get 

(a;  -  1)  (a;  -  2)  (a  -  3)  (a;  -  0)  (a?  -  2)  (g  -  3) 

•     V*  ~  (o  _  1)  (0  -  2)  (0  -  3)Va  +  (1  -  0)  (1  -  2)  (1  -  3)Vb  +  *  *  * ; 

x3  -  6x2  +  11a*  -  6          a*3  -  5a*2  +  6a* 
.-.  yx  = g ya  + g Vb  +  ' ' ' ; 


§  105.  INFINITE  SERIES  AND  THEIR  USES.  315 

the  student  must  fill  in  the  other  terms  himself.     Collect  together  the  different 
terms  in  x,  x2,  x3t  etc.,  and 

2-33333  =  2-30103  +  0*02171a;  -  0*00055a2  +  O'OOOOlaj3. 

When  this  equation  is  solved  by  the  approximation  methods  described  in  a 
later  chapter,  we  get  x  =  215*462  (nearly). 

(4)  Ammonium  sulphate  has  the  electrical  conductivities :  552, 1010, 1779 
units  at  the  respective  concentrations  :  0*778,  1*601,  3*377  grm.  molecules  per 
litre.  Calculate  the  conductivity  of  a  solution  containing  one  grm.  molecule 
of  the  salt  per  litre.  Ansr.  684*5  units  nearly.  Hint.  By  Lagrange's 
formula,  (8), 

_         (1-1-601)  (1 "  3-377)  (1-0*778)  (1-8-377) 

(0*778  - 1*601)  (0*778  -  3*377)        T  (1*601  -  0*778)  (1*601  -  3*377)1UiU  +  *  * ' 
0*601.2*377  0*222.2*377  0*222.0*601 

0*823.  2*599552  +  0*823. 1*776101°  "  2599.  lW7™* 
Simple  proportion  gives  680  units.  But  we  have  only  selected  three  observa- 
tions ;  if  we  used  all  the  known  data  in  working  out  the  conductivity  there 
would  be  a  wider  difference  between  the  results  furnished  by  proportion  and 
by  Lagrange's  formula.  The  above  has  been  selected  to  illustrate  the  use  of 
the  formula. 

(5)  From  certain  measurements  it  is  found  that  if  x  =  618,  y  =  3*927 ; 
x  =  588,  y  =  3*1416  ;  x  =  452,  y  =  1*5708.  Apply  Lagrange's  formula,  in  order 
to  find  the  best  value  to  represent  y  when  x  m  617.     Ansr.  3*898. 

If  the  function  is  periodic,  Gauss'  interpolation  formula  may 

be  used.     This  has  a  close  formal  analogy  with  Lagrange's.1 
sin  \(x  -  b) .  sin  \{x  -  o) . . .  sin  \{x  -  n) 
y*      sinj(a  -  6).sinJ(a  -  c) . . .  sin  \{a  -  nfa  +  "'      {  } 

IV.  Interpolation  by  central  differences. 

A  comparison  of  the  difference  table,  page  309,  with  Newton's 
formula  will  show  that  the  interpolated  term  yx  is  built  up  by 
taking  the  algebraic  sum  of  certain  proportions  of  each  of  the 
terms  employed.  The  greatest  proportions  are  taken  from  those 
terms  nearest  the  interpolated  term.  Consequently  we  should 
expect  more  accurate  results  when  the  interpolated  term  occupies  a 
central  position  among  the  terms  employed  rather  than  if  it  were 
nearer  the  beginning  or  end  of  the  given  series  of  terms. 

Let  us  take  the  series  yQ,  yv  y2,  y3,  y±  so  that  the  term,  yxi  to  be 
interpolated  lies  nearest  to  the  central  term  y2.  Hence,  with  our 
former  notation,  Newton's  expression  assumes  the  form 

1  For  the  theoretical  bases  of  these  reference  interpolation  formulae  th,e  reader 
must  consult  Boole's  work,  A  Treatise  on  the  Calculus  of  Finite  Differences^  London, 
38,  1880. 


316  HIGHER  MATHEMATICS.  §  105. 

Vt + .  -  Vo  +  (2  +  x)^y0  +  (2  +  %  f  +  "W  +  ■■■     (10) 

It  will  be  found  convenient  to  replace  the  suffixes  of  y0,  ylf  y2,  y%,  y^ 
respectively  by  y  _  2,  y  _  v  yQ,  yv  y2.  The  table  of  differences  then 
assumes  the  form 

x_x         y.,         £-■         A2_2         ^ 

*o  JKo  A1  A2_!  A3  A4_2 

*%      y  y%  l  , 

Equation  (10)  must  now  be  written 

%  =  y_i  +  {2  +  x)*_i  +  V  +  %f+%*^  +  ...    (11) 

Let  us  now  try  to  convert  this  formula  into  one  in  which 
only  the  central  differences,  blackened  in  the  above  table,  appear. 
It  will  be  good  practice  in  the  manipulation  of  difference  columns. 
First  assume  that 

Ai  =  J(Ai_1  +  Ai0);  A3  =  J(A8_2  +  A3_1).     .       (12) 
.*.  A3_3  =  2A3- A3_r      .         .         .       (13) 
Again  from  the  table  of  differences 

A>—  -A*-!-  A*_2;  .-.  A«_2  =  A3_x-  A*_2.       (14) 
By  adding  together  (13)  and  (14), 

A«_2  =  A«-iA*_2.        .         .         .       (15) 
In  a  similar  manner,  from  the  table,  and  (15),  we  have 

A«_4-A*_1-A»_J--A«_l-A*  +  JA*_>    .      (16) 
And   also   from   the   first   of   equations  (12),    and   the   fact  that 
A2_j  =  A1,,  -  A1.^  A1.!  =  A1  -  JA2_1}  it  follows  that 
A1-,  =  A*.!  -  A2_2  =  (A*  -  JA2^)  -  (A2^  -  A*  +  1A*_2); 

.-.  Ai_2  =  Ai-|A2_1  +  A3-JA4_2.  .       (17) 

Still  further,  from  the  table  of  differences,  (16),  and  the  fact  that 
A1.!  =  2/0  ~  V-i>  we  Set 

V-2  =  y-i  -  A1  _ 2  =  (t/0  -  A1.^  -  Ai.gj 
=  (2/o  -  ^-i)  "  (^  -  IA1-!  +  A3  -  JA*J2). 
But  A1  _  j  is  equal  to  A1  -  -JA2  _  lf  as  just  shown,  therefore, 

y-2  =  y0-  2A1  +  2A2.!-  a»  +  *A*_2.  .  (18) 
Now  substitute  these  values  of  y_2,  A1^,  A2_2,  A3_2,  from  (15), 
(16),  (17),  and  (18),  in  (11) ;  rearrange  terms  and  we  get  a  new 
formula  (19). 


§  105.  INFINITE  SERIES  AND  THEIR  USES.  317 

x    A1*  +  A1    ,     x2  x(x2  -  1)    A3    ,  +  A3     „ 

v-y.+r-    °  a      +2iA2-1+    3i-        a       +  ---  <19> 

which  is  called  Stirling's  interpolation  formula  (J.  Stirling, 
Methodus  Differentialis,  London,  1730),  when  we  are  given  a  set 
of  corresponding  values  of  x  and  y,  we  can  calculate  the  value  y 
corresponding  to  any  assigned  value  x,  lying  between  x0  and  xv 
Stirling's  interpolation  formula  supposes  that  the  intervals  xx  -  x0, 
%o  ~  x  - 1>  •  •  • are  unity.  If,  however,  h  denotes  the  equal  incre- 
ments in  the  values  xx  -  x0,  x0  -  x  _  x .  .  . ,  Stirling's  formula  becomes 

£  Alo  +  Al-i       x2     2        (x  +  h)x(x-h)  A3_1  +  A3. 
y-y«  +  h'         2         +2TPA-1+         3I/i8         '  2 


\m 


(x  +  h)x2(x-h) 
+  U¥         A  "2 

(x  +  2h)  (x  +  h)x(x -h)(x-  2h)    A5_2  +  A5_8 
+  _  5\h*  '  2  +  "" 

where  y  is  written  in  place  of  yx. 

Example. — The  3  °/0  annuities  on  lives  aged  21,  25,  29,  33,  and  37  are 
respectively  21-857,  21-025,  20-132,  19-145,  and  18-057.  Find  the  annuity  for 
age  30.     Set  up  the  table  of  differences,  h  =  4. 

x.             y.  Al.                         A*.                            A».                          A*. 

21  (21-857) _2  (_  0-832)  _2 

26  (21-025).,  V(_  o-agslx  ("O^1)-.     /-<M>33)_2 

29  (20-132)o  _0.987!  (-0-094). x      _          '          (+0'026)_2 

33  (19-145),  .i-oss!  (-0-101)o 

37  (18-057)2 

o^non      0"940  0-094        15x0-02      15x0026 

...yao»i8a--T--?nri5+  3!x43  -    4!x44  - 

m  20-132  -  0-235  -  0-003  +  0-0008  -  0-0001  =  19895. 
By  Newton's  formula,  we  get  19-895. 

The  central  difference  formula  of  Stirling  thus  furnishes  the 
same  result  as  the  ordinary  difference  formula  of  Newton.  We 
get  different  results  when  the  higher  orders  of  differences  are  neg- 
lected. For  instance,  if  we  neglect  differences  of  the  second  order 
in  formulae  (7)  and  (20),  Stirling's  formula  would  furnish  more 
accurate  results,  because,  in  virtue  of  the  substitution  A1  =  A1  -  JA2_  v 
we  have  really  retained  a  portion  of  the  second  order  of  differences. 
If,  therefore,  we  take  the  difference  formula  as  far  as  the  first, 
third,  or  some  odd  order  of  differences,  we  get  the  same  results 
with  the  central  and  the  ordinary  difference  formulae.  One  more 
term  is  required  to  get  an  odd  order  of  differences  when  central 
differences  are  employed.     Thus,  five  terms  are  required  to  get 


318 


HIGHER  MATHEMATICS. 


§106. 


third  order  differences  in  the  one  case,  and  four  terms  in  the  other. 
For  practical  purposes  I  do  not  see  that  any  advantage  is  to  be 
gained  by  the  use  of  central  differences. 

V.  Graphic  interpolation. 

Intermediate  values  may  be  obtained  from  the  graphic  curve 
by  measuring  the  ordinate  corresponding  to  a  given  abscissa  or 
vice  versd. 

In  measuring  high  temperatures  by  means  of  the  Le  Chatelier- 

Austin  pyrometer,  the  deflec- 
tion of  the  galvanometer  index 
on  a  millimetre  scale  is  caused 
by  the  electromotive  force  gen- 
erated by  the  heating  of  a 
thermo-couple  (Pt  -  Pt  with 
10  °/0  Ed)  in  circuit  with  the 
galvanometer.  The  displace- 
ment of  the  index  is  nearly 
proportional  to  the  tempera- 
ture. The  scale  is  calibrated 
by  heating  the  junction  to 
well-defined  temperatures  and 
plotting  the  temperatures  as 
ordinates,  the  scale  readings 
as  abscissae.  The  resulting 
graph  or  "  calibration  curve  " 
is  shown  in  Fig.  130.  The 
ordinate  to  the  curve  corre- 
sponding to  any  scale  reading,  gives  the  desired  temperature.  For 
example,  the  scale  reading 
ture  1300°. 


1 

PA  1 

1 

t  n  i  ii  i 

lunr 

iZOO' 

2000° 

/gold 

ULP 

UTE 

POT 

\*.\ 

600* 

Xelenium  boils 

r*  ALUMINIUM 

1      1      1 

*00° 

ZINC, 

Sulphur  boils' 

/if 

AO- 

200' 

<r> 

/WATEfTB 

r     1      1      1    " 

ScaleJleading 

fO   60    SO  100  &0  1*0  160  180  200 

Fig.  130.— Calibration  Chart. 


160  "  corresponds  with  the  tempera- 


§  106.  Differential  Coefficients  from  Numerical  Observations. 

It  is  sometimes  necessary  to  calculate  the  value  of  dy/dx  and 
d2y/dx2  from  the  relation  y  =*  f(x).     Three  methods  are  available : — 

J.  Differentiation  of  a  known  function. 

If  corresponding  values  of  two  variables  can  be  represented  in 
the  form  of  a  mathematical  equation,  the  differential  coefficient  of 
the  one  variable  with  respect  to  the  other  can  be  easily  obtained. 


§  106.  INFINITE  SERIES  AND  THEIR  USES.  319 

In  illustration,  A.  Horstmann  (Liebig's  Ann.  Ergbd.,  8,  112,  1872), 
wished  to  compare  the  experimental  values  of  the  heats  of  vaporiz- 
ation, Q,  of  ammonium  chloride  with  those  calculated  from  the 
expression :  Q  =  T(dp/dT)dv,  which  had  been  deduced  from  the 
principles  of  thermodynamics.  He  found  that  the  observed  vapour 
pressure,  p,  at  different  temperatures,  0,  could  be  represented  well 
enough  by  Biot's  formula :  log10^?  =  a  +  bae  ~ 258'6.  Hence,  the 
value  of  dp/dO,  or  dp/dT,  for  the  vapour  pressure  at  any  particular 
temperature  could  be  obtained  by  differentiating  this  formula  and 
substituting  the  observed  values  of  p  and  t  in  the  result.  It  is 
assumed,  of  course,  that  the  numerical  values  of  a,  b  and  a  are 
known.  Following  Horstmann  a  =  5*15790,  b  =  -  3 -34598,  and 
log10a  m  0*9979266  -  1.  Suppose  it  be  required  to  find  the  value 
of  dp/dO  at  300°.  When  0  =  300,  6  -  2585  =  41-5  •  and  a415  =  0-819, 
because  log10a416  =  41-5  log10a  =41-5  x  -0-0020734  =  -  0-086046  ; 
consequently,  0086046  =  -  41'51og10a  =  log10a-41'5  =  log10l'221. 
Hence, a -415  =  1-221 ; .-.  a415  =  0-819;  and 6a416  =  -  3-34598  x  0-8192 
=  -  2-74036.  Hence,  log10p  =  51579  -  2-7403  ;  or,  log10^  =  2-4175 
«=  log10261*5 ;    or,  p  =  261.5.      By   differentiation   of  log1Qp  =  a 

+-  6a  »- 2585 

|£  -  ^a41'6log10a  -  261-5  x  -  2-74035  x  -  0-0020734  -  1-5. 

Examples. — (1)  Assuming  that  the  pressure,  p,  of  steam  at  0°  C.  in  lbs. 
per  square  foot  is  given  by  the  law  0  =  29'71pT  -  37*6,  show  that  when 
p  =  290,  dpjdd  =  15-<>7.  Hint,  de/dp  =  -  l(29'77)p  ' 4/5  ;  .\  dp/de  =  0'168p4/« ; 
.-.  dpjdd  =  0-168  x  2904/5  lbs.  per  square  foot  per  °  C. 

(2)  The  volume,  v„  of  a  cubic  foot  of  saturated  steam  at  T°  abs.  is  given 
by  the  formula  L  =  T(vt  +  v„)dp/dT,  where  vv  the  volume  of  one  pound  of 
water  which  may  be  taken  as  negligibly  small  in  comparison  with  v, ;  L  is 
the  latent  heat  of  one  pound  of  steam  in  mechanical  units,  i.e.,  740,710  ft. 
lbs.  Given  also  the  formula  of  the  preceding  example,  show  that  when 
6  -  127°  C.,2>  =  {(127  +  37-6)/29-77}6 ;  .-.  logp  =  3-71365  ;  .-.  dp/de,  or,  what 
is  the  same  thing,  dpjdT  =  0*168  x  935  =  157  lbs.  per  square  foot  per  degree 
absolute.    Hence,  740710  =  157  x«,x400;  .-.  v,  =  11-8. 

II.  Graphic  interpolation. 

In  the  above-quoted  investigation,  Horstmann  sought  the 
value  of  dp/dT  for  the  dissociation  pressure  of  aqueous  vapour 
from  crystalline  disodium  hydrogen  phosphate  at  different  tempera- 
tures. Here  the  form  of  p  =  f(T)  was  not  known,  and  it  became 
necessary  to  deal  directly  with  the  numerical  observations,  or  with 


320 


HIGHER  MATHEMATICS. 


§106. 


the  curve  expressing  these  measurements.  In  the  latter  case, 
the  tangent  to  the  "smoothed"  or  " faired"  curve  obtained  by- 
plotting  corresponding  values  of  p  and  T  on  squared  paper  will 
sometimes  allow  the  required  differential  coefficient  to  be  obtained. 
Suppose,  for  example,  we  seek  the  numerical  value  of  dp/dO  at 
150°  when  it  is  known  that  when 

p  =  8-698,    9-966,    11-380  lbs.  per  sq.  ft. ; 
0=145,       150,       155°  G. 
These  numbers  are  plotted  on  squared  paper  as  in  Fig.  131.     To 
find  dp/dO  at  the  point  P  corresponding 
with  150°,  and  9*966  lbs.  per  square  foot, 
first  draw  the  tangent  PA  ;  from  P  draw 
PB  parallel  with  the  0-axis.     If  now  the 
pressure  were  to  increase  throughout  5° 
from  150°  to  155°  at  the  same  rate  as  it  is 
increasing  at  P,  the  increase  in  pressure 
for  5°  rise  of  temperature  would  be  equal 
to  the  length   BA,   or  to   1300  lbs.  per 
square  foot.     Consequently,  the  increase 
of  pressure  per  degree  rise  of  temperature 
is  equal  to  1300  -r5  =  260  lbs.  per  sq.  ft. 
Hence  dp/dT  =  260. 
The  graphic  differentiation  of  an  experimental  curve  is  avoided 
if  very  accurate  results  are  wanted,  because  the  errors  of  the  ex- 
perimental curve  are  greatly  exaggerated  when  drawing  tangents. 
If   the   measurements  are  good  better  results  can   be  obtained, 
because    the    curve    does   not  then   want   smoothing.     Graphic 
interpolation  was  accurate  enough  for  Horstmann's  work.      See 
also  O.  W.  Bichardson,  J.  Nicol,  and  T.  Parnell,  Phil.  Mag.  [6], 
8,  1,  1904,  for  another  illustration. 

We  now  seek  a  more  exact  method  for  finding  the  differential 
coefficient  of  one  variable,  say  y,  with  respect  to  another,  say  x, 
from  a  set  of  corresponding  values  of  x  and  y  obtained  by  actual 
measurement. 

III.  From  the  difference  formulae. 

Let  us  now  return  to  Stirling's  interpolation  formula.  Differ- 
entiate (19),  page  317,  with  respect  to  x,  and  if  we  take  the 
difference  between  y0  and  yx  to  be  infinitely  small,  we  must  put 
x  =  0  in  the  result.     In  this  way,  we  find  that 


P 

1 

11000 

A 

10500 

/ 

/ 

10000 

/ 

7 

B 

9500 

/ 

/ 

9O00 

/ 

/ 

8500 

i 

/ 

M 

v" 

k 

5* 

1? 

0* 

IS 

5C 

9>C 

Fig.  181. 


8  106.  INFINITE  SERIES  AND  THEIR  USES.  321 

#=^/Ai0  +  Ai_1_l    A3_1  +  A3  1     AS_2  +  A5_3  \ 

dx      h\%  6*  2  ^30'  2  -•••J-V1; 

This  series  may  be  written  in  the  form 
^_1/Ai0  +  AL1     12A3_1  + A3         12.2*   A*_2  +  A5_8  \ 

daj     H        2  3!'  2  +    5!    '  2       ~7-'T    ^ 

To  illustrate  the  use  of  formula  (2),  let  the  first  two  columns 
of  the  following  table  represent  a  set  of  measurements  obtained 
in  the  laboratory.  It  is  required  to  find  the  value  of  dy/dx  cor- 
responding to  x  =  5*2.     First  set  up  the  table  of  central  differences. 

X.  y.  Al.  A2.  A'.  A*.  A5. 

4-9  (134-290)  _  3 

50  (148-413)  _2 

51  (164-022)  _x 
5-2  (181'272)o 
53  (200337)! 
5  4  (221-406)2 
5-5  (244-692)3 

Make  the  proper  substitutions  in  (2).     In  the  case  of  5*2  only  the 
block  figures  in  the  above  table  are  required.     Thus, 


(14123)  _3 
(15-609)  _  2 

(1-486)  _3 

(0-155).  8 

(17-250)  _i 

(1641)  _  2 

(0-174)  _  2 

<°™>-»     (-0-004)_, 

(19-065)o 

(1-815)  _i 

(0-189)  _i 

(0-015  -2     (  +  0.009)2 

(21-069)! 

(2-004)0 

(0-218)0 

(0-024)  _  x 

(23-286)2 

(2-217)! 

dx     01  \ 


17-250  +  19-065  _  1    Q-174  +  Q-189      1     Q-Q09  -  Q-Q04\ 
2  6'  2  +  30'  2  J* 

.-.  ^  =  181-273. 
dx 

The  student  may  now  show  that  by  differentiating  Stirling's 
formula  twice,  and  putting  x  =  0  in  the  result,  we  obtain  the 
second   differential  coefficient 

which  may  also  be  writtten 

d*y      1/2    o  2    A  2.22fl  2.22.32  Q  \      /JX 

d  =  Pl2!A2-i  "  4!A4-2  +  "6TA6-3  -  -8!-a8-  +  ' '  }  W 
The  difference  columns  should  not  be  carried  further  than  is 
consistent  with  the  accuracy  of  the  data,  otherwise  the  higher 
approximations  will  be  less  accurate  than  the  first.  Do  not  carry 
the  differences  further  than  the  point  at  which  they  begin  to  ex- 
hibit marked  irregularities.  The  A5  differences  in  the  above  table, 
for  instance,  are  "  out  of  bounds  ".  The  first  two  terms  generally 
suffice  for  all  practical  requirements. 

Examples. — (1)  From  Horstmann's  observations  on  the  dissociation 
pressure,  p,  of  the  ammonio-chlorides  of  silver  at  different  temperatures,  0 : 

X 


322  HIGHER  MATHEMATICS.  §  107. 

0=8,  12,  16,  f . .    °C. 

p  =  43-2,        52-0,        65-3, ...cm.  Hg. 
show  that  at  12°,  dp/de  =  2-76. 

(2)  Show  that  dsjde  =  -,4-7  x  10 -6,  at  0°  C,  from  the  following  data  :— 
0  =  1,  0-5,  0,  -  0-5,         -  1-0, . . . ; 

106  x  s  =  1288-3,         1290-7,         1293*1,         1295-4,         1297*8, . .  . 

(3)  Find  the  value  of  d^y/dx2  for  y=5'2  from  the  above  table.    Ansr.  181-4. 

(4)  The  variation  in  the  pressure  of  saturated  steam,  p,  with  temperature 
0  has  been  found  to  be  as  follows  : — 

a  =  90,  95,  100,  105,  110,  115,  120,...; 

p  =  1463,        1765,        2116,        2524,        2994,        3534,  4152, . . . 

Hence  show  that  at  105°  dp/d0   =   87-58,   d*pld02  =  2-48.  Hint,    dyjdx 
=  ift(408  +  470)  -  tM5  +  8)}  =  i  (437-917)  =  87-583. 

§  107.  How  to  Represent  a  Set  of  Observations  by  Means  of 

a  Formula. 

After  a  set  of  measurements  of  two  independent  variables  has 
been  made  in  the  laboratory,  it  is  necessary  to  find  if  there  is  any 
simple  relation  between  them,  in  other  words,  to  find  if  a  general 
expression  of  the  one  variable  can  be  obtained  in  terms  of  the 
other  so  as  to  abbreviate  in  one  simple  formula  the  whole  set  of 
observations,  as  well  as  intermediate  values  not  actually  measured. 

The  most  satisfactory  method  of  finding  a  formula  to  express 
the  relation  between  the  two  variables  in  any  set  of  measurements, 
is  to  deduce  a  mathematical  expression  in  terms  of  variables  and 
constants,  from  known  principles  or  laws,  and  then  determine  the 
value  of  the  constants  from  the  experimental  results  themselves. 
Such  expressions  are  said  to  be  theoretical  formulae  as  distinct 
from  empirical  formulas,  which  have  no  well-defined  relation 
with  known  principles  or  laws. 

The  terms  "formula"  and  "function"  are  by  no  means 
synonymous.  The  function  is  the  relation  or  law  involved  in  the 
process.  The  relation  may  be  represented  in  a  formula  by  symbols 
which  stand  for  numbers.  The  formula  is  not  the  function,  it  is 
only  its  dress.  The  fit  may  or  may  not  be  a  good  one.  This 
must  be  borne  in  mind  when  the  formal  relations  of  the  symbols 
are  made  to  represent  some  physical  process  or  concrete  thing. 

It  is,  of  course,  impossible  to  determine  the  correct  form  of  a 
function  from  the  experimental  data  alone.  An  infinite  number 
of  formulas  might  satisfy  the  numerical  data,  in  the  same  sense 
that  an  infinite  number  of  curves  might  be  drawn  through  a  series 


§  107.  INFINITE  SEKIES  AND  THEIR  USES.  323 

of  points.  For  instance,  over  thirty  empirical  formulae  have  been 
proposed  to  express  the  unknown  relation  between  the  pressure 
and  temperature  of  saturated  steam. 

As  a  matter  of  fact,  empirical  formulae  frequently  originate 
from  a  lucky  guess.  Good  guessing  is  a  fine  art.  A  hint  as  to 
the  most  plausible  form  of  the  function  is  sometimes  obtained  by 
plotting  the  experimental  results.  It  is  useful  to  remember  that 
if  the  curve  increases  or  decreases  regularly,  the  equation  is  prob- 
ably algebraic  ;  if  it  alternately  increases  and  decreases,  the  curve 
is  probably  expressed  by  some  trigonometrical  function. 

If  the  curve  is  a  straight  line,  the  equation  will  be  of  the  form, 
y  =  mx  +  b.  If  not,  try  y  =  axn,  or  y  =  ax/(l  +  bx).  If  the  rate 
of  increase  (or  decrease)  of  the  function  is  proportional  to  itself  we 
have  the  compound  interest  law.  In  other  words,  if  dy/dx  varies 
proportionally  with  y,  y  =  be~ax  or  be**.  If  dy/dx  varies  pro- 
portionally with  x/y,  try  y  «=  bxa.  If  dy/dx  varies  as  x,  try 
y  =  a  +  bx2.  Other  general  formulae  may  be  tried  when  the 
above  prove  unsatisfactory,  thus, 

y  =  ^r^;  V  =  10a  +  bx;  y  =  a  +  blogx;  y  =  a  +  btf, 

Otherwise  we  may  fall  back  upon  Maclaurin's  expansion  in  ascend- 
ing powers  of  x,  the  constants  being  positive,  negative  or  zero. 
This  series  is  particularly  useful  when  the  terms  converge  rapidly.* 

When  the  results  exhibit  a  periodicity,  as  in  the  ebb  and  flow 
of  tides ;  annual  variations  of  temperature  and  pressure  of  the  at- 
mosphere ;  cyclic  variations  in  magnetic  declination,  etc.,  we 
refer  the  results  to  a  trigonometrical  series  as  indicated  in  the 
chapter  on  Fourier's  SBries. 

Empirical  formulae,  however  closely  they  agree  with  facts,  do 
not  pretend  to  represent  the  true  relation  between  the  variables 
under  consideration.  They  do  little  more  than  follow,  more  or 
less  closely,  the  course  of  the  graphic  curve  representing  the  re- 
lation between  the  variables  within  a  more  or  less  restricted  range. 
Thus,  Eegnault  employed  three  interpolation  formulae  for  the  vapour 
pressure  of  water  between  -  32°  F.  and  230°  F.1  For  example, 
from  -  32°  F.  to  0°F.,  he  used  p  =  a  +  b*\  from  0°  to  100°  F., 

1  Rankine  was  afterwards  lucky  enough  to  find  that  logp  =  a  -  fid  ~  ?  -  yd  ~  2, 
represented  Regnault's  results  for  the  vapour  pressure  of  water  throughout  the  whole 
range  -  32°  F.  to  230°  F. 


etc. 


324  HIGHER  MATHEMATICS.  .    §  108. 

logp  =  a  +  bae  +  c/3* ;  from  100°  to  230°  F.,  logp  =  a  +  bae  -  c/3e. 
Kopp  required  four  formulae  to  represent  his  measurements  of  the 
thermal  expansion  of  water  between  0°  and  100°  C.  Each  of  Kopp's 
formulas  was  only  applicable  within  the  limited  range  of  25°  C. 

If  all  attempts  to  deduce  or  guess  a  satisfactory  formula  are 
unsuccessful,  the  results  are  simply  tabulated,  or  preferably  plotted 
on  squared  paper,  because  then  "  it  is  the  thing  itself  that  is  before 
the  mind  instead  of  a  numerical  symbol  of  the  thing  ". 

§  108.    To  Evaluate  the  Constants  in  Empirical  or 
Theoretical  Formulae. 

Before  a*  formula  containing  constants  can  be  adapted  to  any 
particular  process,  the  numerical  values  of  the  constants  must  be 
accurately  known.  For  instance,  the  volume,  v,  to  which  unit 
volume  of  any  gas  expands  when  heated  to  6°,  may  be  represented 

by 

V  =  1  +  aO, 
where  a  is  a  constant.  The  law  embodied  in  this  equation  can 
only  be  applied  to  a  particular  gas  when  a  assumes  the  numerical 
value  characteristic  of  that  gas.  Ii  we  are  dealing  with  hydrogen, 
a  =  0-00366  ;  if  carbon  dioxide,  a  =  0-00371  ;  and  if  sulphur 
dioxide  a  =  0-00385. 

Again,  if  we  want  to  apply  the  law  of  definite  proportions,  we 
must  know  exactly  what  the  definite  proportions  are  before  it  can 
be  decided  whether  any  particular  combination  is  comprised  under 
the  law.  In  other  words,  we  must  not  only  know  the  general  law, 
but  also  particular  numbers  appropriate  to  particular  elements. 
In  mathematical  language  this  means  that  before  a  function  can  be 
used  practically,  we  must  know  : 

(i)  The  form  of  the  function.; 
(ii)  The  numerical  values  of  the  constants. 
The  determination  of  the  form  of  the  function  has  been  discussed 
in  the  preceding  section,  the  evaluation  of  the  constants  remains 
to  be  considered. 

Is  it  legitimate  to  deduce  the  numerical  values  of  the  constants 
from  the  experiments  themselves?  The  answer  is  that  the  nu- 
merical data  are  determined  from  experiments  purposely  made  by 
different  methods  under  different  conditions.  When  all  indepen- 
dently  furnish   the   same  result  it   is   fair  to   assume   that  the 


§  108.  INFINITE  SERIES  AND  THEIR,  USES.  325 

experimental  numbers  includes  the  values  of  the  constants  under 
discussion.1 

In  some  determinations  of  the  volume,  v,  of  carbon  dioxide 
dissolved  in  one  volume  of  water  at  different  temperatures,  0,  the 
following  pairs  of  values  were  obtained  : — 

(9=     0,  5,  10,  15; 

v  =  1-80,         145,         1-18,         1-00. 

As  Herschel  has  remarked,  in  all  cases  of  "direct  unimpeded 
action,"  we  may  expect  the  two  quantities  to  vary  in  a  simple 
proportion,  so  as  to  obey  the  linear  equation, 

y  =  a  +  bx;  wt  have,  v  =  a  +  b6,  .  .  (1) 
which,  be  it  observed,  is  obtained  from  Maclaurin's  series  by  the 
rejection  of  all  but  the  first  two  terms. 

It  is  required  to  find  from  these  observations  the  values  of  the 
constants,  a  and  b,  which  will  represent  the  experimental  data  in 
the  best  possible  manner.     The  above  results  can  be  written, 

(i)  1-80  =  a, 

(ii)  1-45  =  a  +  5b, 

(iii)  1-18  =  a  +  106, 

(iv)  1-00  =  a  +  15b, 

which  is  called  a  set  of  observation  equations.     We  infer,  from 

(i)  and   (ii)  a  =  1-80,  b  =  -  0-07, 
(ii)  and  (iii)  a  =  1'62,  b  =  -  0-054, 
(iii)  and  (iv)  a  =  1-54,  b  =  -  0*036,  etc. 

The  want  of  agreement  between  the  values  of  the  constants 
obtained  from  different  sets  of  equations  is  due  to  errors  of 
observation,  and,  of  course,  to  the  fact  that  the  particular  form  of 
the  function  chosen  does  not  fit  the  experimental  results.  It 
nearly  always  occurs  when  the  attempt  is  made  to  calculate  the 
constants  in  this  manner. 

The  numerical  values  of  the  constants  deduced  from  any  arbi- 
trary set  of  observation  equations  can  only  be  absolutely  correct 
when  the  measurements  are  perfectly  accurate.  The  problem  here 
presented  is  to  pick  the  best  representative  values  of  the  constants 
from  the  experimental  numbers.     Several  methods  are  available. 

1  J.  F.  W.  Herschel' a  A  Preliminary  Discourse  on  the  Study  of  Natural  Phil- 
osophy, London,  1831,  is  worth  reading  in  this  connexion. 


(2) 


326  HIGHER  MATHEMATICS.  §  108. 

I.  Solving  the  equations  by  algebraic  methods. 

Pick  out  as  many  observation  equations  as  there  are  unknowns 
and  solve  for  a,  b,  c  by  ordinary  algebraic  methods.  The  different 
values  of  the  unknown  corresponding  with  the  different  sets  of 
observation  arbitrarily  selected  are  thus  ignored. 

Example.— Corresponding  values  of  the  variables  x  and  y  are  known,  say, 
xv  Vi 5  x2>  V% J  xv  Vs'*  '••  Calculate  the  constants  a,  b,  c,  in  the  interpolation 
formula 

bx 

y  =  a.lOl  +  cx. 
When  xx  =  0,yx  =  a.    Thus  b  and  c  remain  to  be  determined.    Take  logarithms 
of  the  two  equations  in  a?2,  y2  and  x3,  y3  and  show  that, 

This  method  may  be  used  with  any  of  the  above  formulae  when 
an  exact  determination  of  the  constants  is  of  no  particular  interest, 
or  when  the  errors  of  observation  are  relatively  small.  V.  H.  Reg- 
nault  used  it  in  his  celebrated  "  Memoire  sur  les  forces  elastiques 
de  la  vapeur  d'eau "  (Ann.  Chim.  Phys.,  [3],  11,  273,  1844)  to 
evaluate  the  constants  mentioned  in  the  formula,  page  323  ;  so  did 
G.  C.  Schmidt  (Zeit.  phys.  Chem.,  7,  433,  1891);  and  A.  Horst- 
mann  (Liebig's  Ann.  Ergbd.,  8,  112,  1872). 

II.  Method  of  Least  Squares. 

The  constants  must  satisfy  the  following  criterion  :  The  differ- 
ences between  the  observed  and  the  calculated  results  must  be  the 
smallest  possible  with  small  positive  and  negative  differences. 
One  of  the  best  ways  of  fixing  the  numerical  values  of  the  con- 
stants in  any  formula  is  to  use  what  is  known  as  the  method  of 
least  squares.  This  rule  proceeds  from  the  assumption  that  the 
most  probable  values  of  the  constants  are  those  for  which  the 
sum  of  the  squares  of  the  differences  between  the  observed  and  the 
calculated  results  are  the  smallest  possible.  We  employ  the  rule 
for  computing  the  maximum  or  minimum  values  of  a  function. 

In  this  work  we  usually  pass  from  the  special  to  the  general. 
Here  we  can  reverse  this  procedure  and  take  the  general  case  first. 
Let  the  observed  magnitude  y  depend  on  x  in  such  a  way  that 

y  =  a  +  bx (3) 

It  is  required  to  determine  the  most  probable  values  of  a  and  b.    For 


§108. 


INFINITE  SERIES  AND  THEIR  USES. 


32? 


perfect  accuracy,we  should  have  the  following  observation  equations : 

a  +  bx1  -  y1  =  0 ;  a  +  bx2  -  y2  =  0 ;  . . .  a  +  bxn  -  yn  =  0. 
In  practice  this  is  unattainable.     Let  vv  v2,  . . .  vn  denote  the  actual 
deviations  so  that 

a  +  bx1-y1  =  v1;  a  +  bx2  -  y2  =  v2;  . . .  a  +  bxn  -  yn  =  vn. 
It  is  required  to  determine  the  constants  so  that, 

2,(y2)  =  V-f  +  V22  +  ...  +  V2  is  a  minimum. 

With  observations  affected  with  errors  the  smallest  value  of  v2 
will  generally  differ  from  zero ;  and  the  sum  of  the  squares  will 
therefore  always  be  a  positive  number.  We  must  therefore  choose 
such  values  of  a  and  b  as  will  make 

s;:>  +  &*„  -  y„f  ' 

the  smallest  possible.  This  condition  is  fulfilled,  page  156,  by 
equating  the  partial  derivatives  of  %(v2)  with  respect  to  a  and  b 
to  zero.     In  this  way,  we  obtain, 

Ya^{a  +  bx  -  y)2  =  0;  hence,  2,(a  +  bx  -  y)  =  0 ; 

ft 

>-,2(a  +  bx.-  y)2  =  0 ;  hence,  %x(a  +  bx  -  y)  =  0. 

If  there  are  n  observation  equations,  there  are  n  a's  and  2(a)  =  na, 
therefore, 

na  +  bl{x)  -  S(y)  =  0;  a%(x)  +  bl(x2)  -  %(xy)  =  0. 
Now  solve  these  two  simultaneous  equations  for  a  and  6, 

_  %(x) .  3(sy)  -  3(*) .  3(y) .       _  3(s)3(y)  -  nSQcy) 

[2(a;)]2  -  rc2(z2)         '  °  "     [2(z)]2  -  n2(z2)  ■       W 
which  determines  the  values  of  the  constants. 

Returning  to  the  special  case  at  the  commencement  of  this 
section,  to  find  the  best  representative  value  of  the  constants  a  and 
b  in  formula  (1).  Previous  to  substitution  in  (4),  it  is  best  to 
arrange  the  data  according  to  the  following  scheme : — 


e. 

v. 

0». 

ev. 

0 

5 

10 

15 

1-80 
1-45 
1-18 
VOO 

0 

25 

100 

225 

0 

7'25 
11-80 
15-00 

S(fl)  =  30 

%(*)  =  5-43 

2(6>2)  =  350 

2  (flu)  =  34-05 

328 


HIGHER  MATHEMATICS. 


§108. 


Substitute   these   values   in   equation  (4),   n,  the  number  of 
observations,  =  4,  hence  we  get 

a  =  1-758;  b  =  -  0-0534. 
The  amount  of  gas  dissolved  at  0°  is  therefore  obtained  from  the 
interpolation  formula, 

v  =  1-758  -  0-0534(9. 
To  show  that  this  is  the  best  possible  formula  to  employ,  in 
spite  of  1*758  volumes  obtained  at  0°,  proceed  in  the  following 
manner : — 


Temp.  =  e. 

Volume  of  gas  =  v.  > 

Difference  between 

Calculated  and 

Observed. 

Square  of  Difference 

between  Calculated 

and  Observed. 

Calculated. 

Observed. 

0 
5 

10 
15 

1-758 
1-491 
1-224 
0-957 

1-80 
1-45 
1-18 
1-00 

-  0-042 
+  0-041 
+  0-044 

-  0-043 

0-00176 
0-00168 
0-00194 
0-00185 

0-00723 

The  number  0 -00723,  the  sum  of  the  squares  of  the  differences 
between  the  observed  and  the  calculated  results,  is  a  minimum. 
Any  alteration  in  the  value  of  either  a  or  b  will  cause  this  term  to 
increase.  This  can  easily  be  verified.  For  example,  if  we  try  the 
very  natural  a  =  1-80,  &  =  -  0065,  we  get  0-039;  if  a  =  1-772, 
b  =  -  0056  we  get  0-0082,  etc. 

Examples. — (1)  Find  the  law  connecting  the  length,  I,  of  a  rod  with 
temperature,  8,  when  the  length  of  a  metre  bar  at  0°  elongates  with  rise  of 
temperature  according  to  the  following  scheme : — 

6=      20°,  48°,  50°,  60°  C; 

2  =  1000-22,         1000-65,         1000-90,         1001-05  mm. 

(F.  Kohlrausch's  Leitfaden  der praktischen  Physik,  Leipzig,  12, 1896.)    During 

the  calculation,  for  the  sake  of  brevity,  use  I  =  -22,  -65,  "9  and  1-05.     Assume 

I  =  a  +  bd,  and  show  that  a  =  999-804,  b  =  0-0212,  or  I  =  999-804  +  0012120. 

(2)  According  to  G.  J.  W.  Bremer's  measurements  (Zeit.  phys.  Chem.,  3, 
423,  1889),  aqueous  solutions  of  sodium  carbonate  containing  p  °/0  of  the  salt 
expand  by  an  amount  v  as  indicated  in  the  following  table : — 
p  =  3-2420        4-8122         7'4587         10-1400; 
104  x  v  =  1-766,         2-046,         2-342,  2-732. 

Hence  show  that  v  =  0-0001354  +  0'00001360.p. 

Suppose  that  instead  of  the  general  formula  (3),  we  had. 
started  with 

y  =  a  +  bx  +  ex2,   .         .         .         .         (5  J 


§  108.  INFINITE  SERIES  AND  THEIR  USES.  329 

where  a,  b  and  c  are  constants  to  be  determined.     The  resulting 
formulae  for  b  and  c  (omitting  a),  analogous  to  (4),  are, 

.  S(s*) .  %{xy)  -  S(s») .  S(qfy) .  r  __  SQk2) •  S(a%)  -  S(s3)  ■  S(sy)  (6x 
b  =  S(a?).:§(a*)-[2(a?)Ja  '  ~  2(0?) .  S(fl?*)  -  [S(aj»)]a  '  V 
These  two  formulaB  have  been  deduced  by  a  similar  method  to 
that  employed  in  the  preceding  case,  a  is  a  constant  to  be 
determined  separately  by  arranging  the  experiment  so  that  when 
x  =  0,  a  =  y0. 

Examples. — (1)  The  following  series  of  measurements  of  the  tempera- 
ture, 6,  at  different  depths,  x,  in  an  artesian  well,  were  made  at  Grenelle 
(France) : — 

x  =    28,  66,  173,         248,  298,         400,         505,        548; 

0  =  11-71,      12-90,       16-40,       20-00,       22-20,       23-75,      26-45,     27*70. 
The  mean  temperature  at  the  surface,  where  x  =  0,  was  10*6°.     Hence  show 
that  at  a  depth  of  x  metres,  0  =  10-6  +  0-042096a  -  0-000020558a;2. 
(2)  If,  when  x  =  0,  y  =  1  and  when 
mm   8-97,  2056,  36-10,  49-96,  62-38,  83-73; 

y  =  1-0078,  1-0184,  1-0317,  1-0443,  1-0563,  10759. 

Hence  show  that  y  -  1  +  0-00084<e  +  0'0000009a;2. 

Thomson  (Wied.  Ann.,  44,  553,  1891)  employed  the  general 
formulae  for  a,  b,  c,  when  still  another  correction  .term  is  included, 
namely, 

y  =  ax  +  bx2  +  ex3.        ...         (7) 
Illustrations  will  be  found  in  the  original  paper. 

If  three  variables  are  to  be  investigated,  we  may  use  the 
general  formula 

z  =  ax  +  by.  .         .         .         (8) 

The  reader  may  be  able  to  prove,  on  the  above  lines,  that 

S(s2) .  S(o»)  -  %xy) .  S(yg) , ,  _  S(a?) .  S(y*)  -  S(sy) .  3(a»)  ,q* 
a  ~      S(a2) .  W)  -  [Z(xy)Y     '  %(x*) .  :%*)  -  [l(xy)f     '       ^  > 

M.  Centnerszwer  (Zeit.  phys.  Chem.,  26,  1,  1896)  referred  his  ob- 
servations on  the  partial  pressure  of  oxygen  during  the  oxidation 
of  phosphorus  in  the  presence  of  different  gases  and  vapours  to  the 
empirical  formula 

px  =  p0  -  a  log  (1  +  bx) ;  or  to  p0  -  px  =  a  log  (1  +  bx), 

where  p0  denotes  the  pressure  of  pure  oxygen,  px  the  partial  pres- 
sure of  oxygen  mixed  with  x  °/   of  foreign  gas  or  vapour.     Show 


330  ttlGHEtt  MATHEMATICS.  §  108. 

with  Centnerszwer,  that  if  y  =  p0  -  px 

%) .  3(s*)  -  3(afy) .  S(g») . ,  _  3(sy)  ■  S(s3)  -  2(3%) .  %{x*)  (m 
a  ~      S(s*) .  S(s*)  -  P(oj8)]»      '  °  .       2(z2) .  S(s*)  -  [2(*3)]2     ' l    j 

Example.— Show,  with  Centnerszwer,  that  a  —  184,  &  =  113  for  chlor- 
benzene  when  it  is  known  that  when 

i>x  =  561,        549,  536,  523,  509,  485; 

«=    0,        0-054,         0-108,         0-215,         0-430,        0-858. 

The  method  of  least  squares  assumes  that  the  observations  are 
all  equally  reliable.  The  reader  will  notice  that  we  have  assumed 
that  one  variable  is  quite  free  from  error,  and  very  often  we  can 
do  so  with  safety,  especially  when  the  one  variable,  can  be 
measured  with  a  much  greater  degree  of  accuracy  than  the  other. 
We  shall  see  later  on  what  to  do  when  this  is  not  the  case. 

III.  Graphic  methods. 

Eeturning  to  the  solubility  determinations  at  the  beginning  of 
this  section,  prick  points  corresponding  to  pairs  of  values  of  v  and 
6  on  squared  paper.  The  points  lie  approximately  on  a  straight 
line.  Stretch  a  black  thread  so  as  to  get  the  straight  line  which 
lies  most  evenly  among  the  points.  Two  points  lying  on  the 
black  thread  are  v  —  l'O,  6  =  14-5,  and  v  =  1*7,  6  =  1*5. 

.-.  a  +  14-56  =  1 ;  a  +  1-56  =  17. 

By  subtraction,  b  =  -  0-54,  .-.  a  =  1*78.  It  is  here  assumed  that 
the  curve  which  goes  most  evenly  among  the  points  represents  the 
correct  law,  see  page  148.  But  the  number  of  observations  is, 
perhaps,  too  small  to  show  the  method  to  advantage.  Try 
these : 

p  =    2,        4,        6,        8,       10,      20,      25,      30,      35,      40, 
s  =  1-02,  103,  1-06,  1-07,  1-09,  1-18,  1-23,  1-29,  1*34,  1-40, 

where  s  denotes  the  density  of  aqueous  solutions  containing  p  °/0 
of  calcium  chloride  at  15°  0.  The  selection  of  the  best  "  black 
thread  "  line  is,  in  general,  more  uncertain  the  greater  the  mag- 
nitude of  the  errors  of  observation  affecting  the  measurements. 
The  values  deduced  for  the  constants  will  differ  slightly  with 
different  workers  or  even  with  the  same  worker  at  different  times. 
With  care,  and  accurately  ruled  paper,  the  results  are  sufficiently 
exact  for  most  practical  requirements. 


§108. 


INFINITE  SERIES  AND  THEIR  USES. 


331 


When  the  "  best  "  curve  has  to  be  drawn  freehand,  the 
results  are  still  more  uncertain.  For  example,  the  amount  of 
"active"  oxygen,  y,  contained  in  a  solution  of  hydrogen  dioxide 
in  dilute  sulphuric  acid  was  found,  after  the  lapse  of  t  days, 
to  be: 

*  =  .6,       9,       10,     14,     18,     27,     34,     38,     41,     54,     87, 
y  =  3-4,    3-1,    3-1,    2-6,    2-2,    1-3,    0-9,    0-7,    0-6,    0-4,    0-2, 

where  2/  =  3*9  when  t  =  0.     We  leave  these  measurements  with 
the  reader  as  an  exercise. 

In  J.  Perry's  Practical  Mathematics,  London,  1899,  a  trial 
plotting  on  "logarithmic  paper  "  is  recommended  in  certain  cases. 
On  squared  paper,  the  distances  between  the  horizontal  and  vertical 
lines  are  in  fractions  of  a  metre  or  of  a  foot.  On  logarithmic 
paper  (Fig.  132),  the  distances  between  the  lines,  like  the  divisions 
on  the  slide  rale,  are  proportional  to  the 
logarithms  of  the  numbers.  If,  therefore,  50 
the  experimental  numbers  follow  a  law  M 
like  log10o*  +  alog1(#  =  constant,  the  func- 
tion can  be  plotted  as  easily  as  on  squared  20 
paper.  If  the  resulting  graph  is  a 
straight  line,  we  may  be  sure  that  we 
are  dealing  with  some  such  law  as 
xya  =  constant ;  or,  (x  +  a)  (y  +  b)a  = 
constant. 


30  40 


Log.  Paper. 


Example.— The  pressure,  pt  of  saturated  steam  in  pounds  per  square 
inch  when  the  volume  is  v  cubic  feet  per  pound  is 

p  =    10,         20,  30,  40,         50,         60, 

v  =  37-80,     19-72,     13-48,     10-29,     8-34,     6'62. 
Hence,  by  plotting  corresponding  values  of  p  and  v 
on  logarithmic  paper,  we  get  the  straight  line  : 

logioP  +  7log10v  =  log106 ;  hence,  pv1'0™  =  382, 
since  log10o  =  2-5811,  .-.  b  =  382  and  y  =  1-065.     The 
graph  is  shown  on  log  paper  in  Fig.  132,  and  on 
ordinary  squared  paper  in  Pig.  134. 


50 


40 


30 


\ 

v 

v 

V 

\ 

y. 

0    10 


30    40    50 


A  semi-logarithmic  paper  (Fig.  133)  may 
be  made  with  distances  between  say  the  hori- 
zontal columns  in  fractions  of  a  metre,  while 
the  distances  between  the  vertical  columns 
are  proportional  to  the  logarithms  of  the  numbers.  Functions 
obeying  the  compound  interest  law  will  plot,  on  such  paper,  as  a 


Fig.  133.— Semi-log. 
Paper. 


332 


HIGHER  MATHEMATICS. 


§108. 


straight  line. 


One  advantage  of  logarithmic  paper  is  that  the 
skill  required  for  drawing  an  accurate  free- 
hand curve  is  not  required.  The  stretched 
black  thread  will  be  found  sufficient.  With 
semi-logarithmic  paper,  either  x  +  log10?/  = 
constant ;  or,  y  +  log10#  =  constant  will  give 
a  straight  line. 

According  to  C.  Eunge  and  Paschen's  law, 
if  the  logarithms  of  the  atomic  weights  are 
20  30  40  5a       plotted  as  ordinates  with  the  distances  be- 
Fig.  134.  tween   the   brightest   spectral   lines   in    the 

magnetic  field  as  abscissae,  chemically  allied  elements  lie  on  the 
same  straight  line.  This,  for  example,  is  the  case  with  magnesium, 
calcium,  strontium,  and  barium.  Eadium,  too,  lies  on  the  same 
line,  hence  C.  Runge  and  J.  Precht  (Ber.  deut.  phys.  Ges.,  313, 
1903)  infer  the  atomic  weight  of  radium  to  be  257*8.  Obviously 
we  can  plot  atomic  weights  and  the  other  data  directly  on  the 
logarithmic  paper.  Another  example  will  be  found  in  W.  N. 
Hartley  and  E.  P.  Hedley's  study  (Journ.  Chem.  Soc,  91,  1010, 
1907),  of  the  absorption  spectra  solutions  of  certain  organic  com- 
pounds where  the  oscillation  frequencies  were  plotted  against  the 
logarithms  of  the  thicknesses  of  the  solutions. 

Examples. — (1)  Plot  on  semi-logarithmic  paper  Harcourt  and  Esson's 
numbers  (I.e.) : 

t=    2,  5,  8,        11,        14,        17,        27,        31,        35,        44, 

y  =  94-8,    87-9,     81-3,     74-9,    68*7,    64-0,     49-3,     44-0,     39-1,     316, 

for  the  amount  of  substance  y  remaining  in  a  reacting  system  after  the  elapse 

of  an  interval  of  time  t.     Hence  determine  values  for  the  constants  a  and  b  in 

y  =  ae  -« ;  i.e.,in  \og10y  +  bt  =  log10a, 
a  straight  line  on  "  semi-log  "  paper.     The  graph  is  shown  in  Fig.  133  on 
"  semi-log  "  paper  and  in  Fig.  134  on  ordinary  paper. 

(2)  What  "  law  "  can  you  find  in  J.  Perry's  numbers  (Proc.  Roy.  Soc,  23, 
472,  1875), 

6  =  58,       86,  148,        166,        188,       202,        210, 

G  =  0,      -004,       -018,      -029,       -051,       -073,       -090, 
for  the  electrical  conductivity  C  of  glass  at  a  temperature  of  6°  F.  ? 

(3)  Evaluate  the  constant  a  in  S.  Arrhenius'  formula,  77  =  ax,  for  the  vis- 
cosity 7}  of  an  aqueous  solution  of  sodium  benzoate  of  concentration  x,  given 

77  =  1-6498,       1-2780,       1-1303,       1-0623, 
x=       1,  h  h  h 

Several  other  methods  have  been  proposed.     Gauss'  method, 
for  example,  will  be  taken  up  later   on.     See  also  Hopkinson, 


§  109.  INFINITE  SERIES  AND  THEIR  USES.  333 

Messenger  of  Mathematics,  2,  65,  1872 ;  or  S.  Lupton's  Notes  on 
Observations,  London,  104,  1898. 

§  109.    Substitutes  for  Integration. 

It  may  not  always  be  convenient,  or  even  possible,  to  integrate 
the  differential  equation  ;  in  that  case  a  less  exact  method  of 
verifying  the  theory  embodied  in  the  equation  must  be  adopted. 
For  the  sake  of  illustration,  take  the  equation 

£  =  k{a-x);     .         .         .         .         (1) 

used  to  represent  the  velocity  of  a  chemical  reaction,  x  denotes  the 
amount  of  substance  transformed  at  the  time  t ;  and  a  denotes  the 
initial  concentration.  Let  dt  denote  unit  interval  of  time,  and  let 
Ax  denote  the  difference  between  the  initial  and  final  quantity  of 
substance  transformed  in  unit  interval  of  time,  then  \Ax  denotes 
the  average  amount  of  substance  transformed  during  the  same 
interval  of  time.     Hence,  for  the  first  interval,  we  write 

Ax  =  k^a  -  $Ax), 
which,  by  algebraic  transformation,  becomes 

*?-t?U'        •     •     •     (2) 

For  the  next  interval, 

Ax  =  kx(a  -  x  -  \Ax),  etc. 
These  expressions  may  be  used  in  place   of  the  integral  of  (1), 
namely 

»-?**•?>    ;...    •    •    (3) 

for  the  verification  of  (1). 

With  equations  of  the  second  order 

dx 

df  =  h(a  -  XY>  ...         (4) 

we  get,  in  the  same  way, 

k2a*  h(a  -  xf 

Aa?  =  Tv^a ;  Ax  =  r+ tga^xy etc"     •     (5) 

by  putting,  as  before,  Ax  in  place  of  dx,  dt  =  1,  x  =  \Ax,  and 
remembering  that  the  second  power  of  Ax  is  negligibly  small. 
The  regular  integral  of  (4)  is 

h*h'tt  ...        (6) 


334 


HIGHER  MATHEMATICS. 


§109. 


Numerical  Illustration. — Let  us  suppose  that  hx  and  k2  are  both  equal 
to  01,  and  that  a  =  100.     From  (2) 

0-1  x  100      -  en 

9-52;  .-.  a  -  x 


Ax  = 


Ax 


1-05 
0*1  x  90-48 


100  -  9-52  =  90-48  ; 


8-62 ;  .-.a-  x  =  90-48  -  8-62  =  81-87. 


1-05 

Again  from  (5),  for  reactions  of  the  second  order 

0-1  x  10,000 

x  =  100  -  90-09  = 


Ax  = 


AX  = 


=  90-99  ;  .-.  a 
4-33;  .-.a 


9-09  -  4-33  =  4-76. 


1  +  0-1  x  100 

(9-09)2  x  0-1 

1  +  0-1  x  909 

The  following  table  shows  that  the  results  obtained  by  this  method  of 
approximation  compare  very  favourably  with  those  obtained  from  the  regular 
integrals  (3)  and  (6).  There  is,  of  course,  a  slight  error,  but  that  is  usually 
within  the  limits  of  experimental  error. 


First  Order. 

Second  Order. 

t. 

a  -  x. 

a  - 

■  X. 

by  (2*. 

by  (3). 

by  (5). 

by  (6). 

0 

100 

100 

0 

100 

100 

1 

90-48 

90-48 

1 

9-09 

9-09 

2 

81-86 

81-86 

2 

4-76 

4-76 

3 

74-08 

74-06 

3 

3-23 

3-23 

4 

67-03 

67-01 

4 

2-44 

2-44 

5 

60-65 

60-63 

5 

1-96 

1-96 

6 

54-88 

54-86 

6 

1-64 

1-64 

7 

49-66 

49-64 

7 

1-41 

1-41 

8 

44-93 

44-91 

8 

1-23 

1-24 

This  method  of  integration  was  used  by  W.  Federlin  (Zeit. 
phys.  Ghem.,  51,  565,  1902)  in  his  study  of  the  reaction  between 
phosphorous  acid,  potassium  iodide,  and  potassium  persulphate  ; 
and  by  E.  Wegscheider  {Zeit.  phys.  Chem.,  51,  52,  1902)  for  the 
saponification  of  the  sulphonic  esters. 

The  student  should  always  be  on  the  lookout  for  short  cuts 
and  simplifications.  Thus,  it  may  be  possible  to  transform  the 
integral  into  a  simpler  form  before  evaluating  by  the  methods  of 
approximation.     For  example,  let 

u  =  jy .  dx 

be  the  integral.  In  one  investigation  (E.  A.  Lehfeldt,  Phil.  Mag., 
[5],  56,  42,  1898;  [6],  1,  377,  403,  1901),  y  represented  the  con- 
centration, x  the  electromotive  force,  and  u  the  osmotic  pressure 


§  110.  INFINITE  SERIES  AND  THEIR  USES.  335 

of  a  solution,  y  was  a  known  function,  hence,  dy  could  be  readily- 
calculated.     Integrate  by  parts,  and  we  get 

u  =  xy  -  jx .  dy, 
which  can  be  evaluated  by  the  planimeter,  or  any  other  means. 
Again,  to  calculate  the  vapour  pressure,  p2,  in  the  expression 

where  px  and  p2  denote  the  vapour  pressure  of  two  components  of 
a  mixture ;  x  is  the  fractional  composition  of  the  mixture.  Sup- 
pose that  pl  and  x  are  known,  it  is  required  to  calculate  p2.  Here 
also,  on  integration  by  parts, 

The  second  setting  is  much  better  adapted  for  numerical  com- 
putation. 

§  110.    Approximate  Integration. 

We  have  seen  that  the  area  enclosed  by  a  curve  can  be 
estimated  by  finding  the  value  of  a  definite  integral.  This  may 
be  reversed.  The  numerical  value  of  a  definite  integral  can  be 
determined  from  measurements  of  the  area  enclosed  by  the  curve. 
For  instance,  if  the  integral  jf(x) .  dx  is  unknown,  the  value  of 

I  f(x) .  dx  can  be  found  by  plotting  the  curve  y  =  f(x) ;  erecting 

ordinates  to  the  curve  on  the  points  x  =  a  and  x  =  b  ;  and  then 
measuring  the  surface  bounded  by  the  #-axis,  the  two  ordinates 
just  drawn  and  the  curve  itself. 

This  area  may  be  measured  by  means  of  the  planimeter,  an 
instrument  which  automatically  registers  the  area  of  any  plane 
figure  when  a  tracer  is  passed  round  the  boundary  lines.  A  good 
description  of  these  instruments  by  O.  Henrici  will  be  found  in  the 
British  Association's  Beports,  496,  1894. 

Another  way  is  to  cut  the  figure  out  of  a  sheet  of  paper,  or 
other  uniform  material.  Let  Wj  be  the  weight  of  a  known  area  a^ 
and  w  the  weight  of  the  piece  cut  out.  The  desired  area  x  can 
then  be  obtained  by  simple  proportion, 

w1  :  a  =  w  :  x. 

Other  methods  may  be  used  for  the  finding  the  approximate 
value  of  an  integral  between  certain  limits.     First  plot  the  curve. 


336  HIGHER  MATHEMATICS.  §  110. 

Divide  the  curve  into  n  portions  bounded  by  n  +  1  equidistant 
ordinates  y0,  yv  y2, .  . .,  yn,  whose  magnitude  and  common  distance 
apart  is  known,  it  is  required  to  find  an  approximate  expression  for 
the  area  so  divided,  that  is  to  say,  to  evaluate  the  integral 

f(x).dx. 
Jo 
Assuming  Newton's  interpolation  formula 

f(x)  =  y0+  xA\  +  2i  x(x  -  1) A*0  +  . . .,       *.  (1) 

we  may  write, 

.*.      f(x) .  dx  =  y0\  dx  +  aO  x.dx  +     ~%\x(x  r  1)^  +  . ..,    (2) 

which  is  known  as  the  Newton -Cotes  integration  formula.    We 

may  now  apply  this  to  special  cases,  such  as  calculating  the  value 
of  a  definite  integral  from  a  set  of  experimental  measurements,  etc. 

I.  Parabolic  Formula. 
Take  three  ordinates.      There  are  two  intervals.      Eeject  all 
terms  after  A20.    Eemember  that  Ax0  =  yx  -  y0  and  A20  =  y2  -  2y1  +  y0. 
Let  the  common  difference  be  unity, 

?M .  dx  -  %0  +  2A\,  +  ia»0  =  fy,  +  iVl  +  y2).        (3) 

Jo 

If  h  represents  the  common  distance  of  the  ordinates  apart,  we 
bay,e  the  familiar  result  known  as  Simpson's  one-third  rule,  thus, 


l 


0  A'      6 


Vz  ,V    \ 


f(x) .  dx  =  lh{y,  +  4^  +  y2).        .        .         (4) 

A  graphic  representation  will  perhaps  make  the  assumptions  in- 
volved in  this  formula  more  apparent.    Make 
b — c^p  the  construction  shown  in  Fig.  135.      We 

seek  the  area  of  the  portion  ANNA'  cor- 
responding to  the  integral  f{x) .  dx  between 
the  limits  x  =  x0  and  x  =  xn,  where  f(x) 
represents  the  equation  of  the  curve  ABGDN. 
Assume  that  each  strip  is  bounded  on  one 
Fig.  135.  g^e  ^j  a  parabolic  curve.     The  area  of  the 

portion 

ABCC'A'  =  Area  trapezium  ACG'A'  +  Area  parabolic  segment  ABCA, 
From  well-known  mensuration  formulae  (16),  page  601,  the  area 
of  the  portion 

ABCC'A'  =  A'C[i(A'A  +  C'C)  +  %{B'B  -  $(A'A  +  C'C)}]; 
=  ZhQA'A  +  $BB  +  ICC)  -  \h(A'A  +  ±B'B  +  Q'C).      (5) 


$110.  INFINITE  SERIES  AND  THEIR  USES.  337 

Extend  this  discussion  to  include  the  whole  figure, 

Area  ANITA'  =  Jfc(l  +  4  +  2  +  4  +  ...  +2  +  4  +  1),  (6) 
where  the  successive  coefficients  of  the  perpendiculars  A  A' ',  BB', . . . 
alone  are  stated  ;  h  represents  the  distance  of  the  strips  apart.  The 
greater  the  number  of  equal  parts  into  which  the  area  is  divided, 
the  more  closely  will  the  calculated  correspond  with  true  area. 

Put  OA'  =  x0 ;  ON  —  xn ;  A'N  =  xn  -  x0  and  divide  the  area 
into  n  parts  ;  h  =  (xn  -  x0)/n.  Let  y0,  yv  yv . . .  yn  denote  the 
successive  ordinates  erected  upon  Ox,  then  equation  (6)  may  be 
written  in  the  form, 

J*/(rC)  * dx  =  *htty°  +  *J  +  4(^  +  2/3  +  •  -  •  +  2/»  - 1)  I         (7) 
+  %2  +  2/4  +  ---  +  2/n-2).  .    J 

In  practical  work  a  great  deal  of  trouble  is  avoided  by  making 
the  measurements  at  equal  intervals  x1  -  x0,  x2  -  xv  . . .,  xn  -  xn  _  v 
E.  Wegscheider  (Zeit.  phys.  Chem.,  41,  52,  1902)  employed  Simp- 
son's rule  for  integrating  the  velocity  equations  for  the  speed  of 
hydrolysis  of  sulphonic  esters;  and  G.  Bredig  and  ~F.  Epstein 
(Zeit.  anorg.  Chem.,  42,  341,  1904)  in  their  study  of  the  velocity  of 
adiabatic  reactions. 

Examples. — (1)  Evaluate  the  integral  ja? .  dx  between  the  limits  1  and  11 
by  the  aid  of  formula  (6),  given  h  =  1  and  y0,  yv  y2,  y3, . . .  y8,  y9,  yw  are  re- 
spectively  1,  8,  27,  64, . . .  1000,  1331.  Compare  the  result  with  the  absolutely 
correct  value.     From  (6),' 

J    xs .  dx  =  i(10980)  =  -3G60  ;  andf  x  .  dx  =  £(11)4  -  i(l)4  =  3660, 

is  the  perfect  result  obtained  by  actual  integration. 

(2)  In  measuring  the  magnitude  of  an  electric  current  by  means  of  the 
hydrogen  voltameter,  let  C0,  Clt  C2, .  . .  denote  the  currents  passing  through 
the  galvanometer  at  the  times  tQt  t^,  t2, . . .  minutes.  The  volume  of  hydrogen 
liberated,  v,  will  be  equal  to  the  product  of  the  M  intensity  "  of  the  current,  C 
amperes,  the  time,  t,  and  the  electrochemical  equivalent  of  the  hydrogen,  x ; 
.*.  v  =  xCt.  Arrange  the  observations  so  that  the  galvanometer  is  read  after 
the  elapse  of  equal  intervals  of  time.  Hence  ^-^0  =  ^-^  =  ^-^2=.  *.«&. 
Prom  (7), 


/! 


tC.dt=ih{(C0  +  C»)  +  4(C1  +  C9+...  +C„_1)  +  2(C2  +  C4+...+C„_2)}. 

In  an  experiment,  v  =  0-22  when  t  =  3,  and 

t  =  1-0,  1-5,  20,  2-5,  3-0,    . . . ; 

C  =  1*53,        1-03,        0-90,        0-84,       0-57, . . . 
f4  0-5 

''•jG.dt  =  -y{(l*53  +  0-57)  +  4(1-03  +  0'84)  +  2  x  0-90}  =  1-897. 

•••a5  =  r8§  =  °-1159- 

Y 


338  HIGHER  MATHEMATICS.  §  110. 

This  example  also  illustrates  how  the  value  of  an  integral  can  be  obtained 
from  a  table  of  numerical  measurements.  The  result  01159,  is  better  than  if 
we  had  simply  proceeded  by  what  appears,  at  first  sight,  the  more  correct 
method,  namely, 

y .  dt  =  {tx  -  g^o^i  +  {t2  _  t$±±£*  + . . .  =  i-9i, 

0*22 
for  then  x  =  — — ■  =  0-1152.     The  correct  value  is  0-116  nearly. 

(3)  If  jdz  =  jealx(b  -  x)~1dx,  where  b  is  the  end  value  of  x,  then,  in  the 


/: 


J     exi  eUxi  +  «2>  e*2    \ 

\b^x7  +  H  _  a/-   +  •  )  +  V=lJ' 


r 

Jo 


2       A  6       Kb-x^^b-  l{xx  +  x.2) 

between  the  limits  xx  and  x2.     Hint.  Use  (4) ;  h  =  ^(xx  +  x2). 

If  we  take  four  ordinates  and  three  integrals,  (4)  assumes  the  form 

i 

f{x) .  dx  =  §h(y0  +  S(yi  +  y2)  +  y3) ; .         .         (8) 

) 

where  h  denotes  the  distance  of  the  ordinates  apart,  y0,  y^  . . .  the 
ordinates  of  the  successive  perpendiculars,  in  the  preceding  diagram. 
This  formula  is  known  as  Simpson's  three-eighths  rule.  If  we 
take  seven  ordinates  and  neglect  certain  small  differences,  we  get 

f(x) .  dx  =  T3<5-%0  +  6yY  +  y2  +  6y2  +  y±  +  5y5  +  y6)  (9) 
Jo 
which  is  known  as  Weddle's  rule  (Math.  Joum.,  11,  79,  1854). 
J.  E.  H.  Gordon  {Proc.  Roy.  Soc,  25,  144,  1876  ;  or  Phil.  Trans., 
167,  i.,  1,  1877)  employed  Weddle's  rule  to  find  the  intensity  of 
the  magnetic  field  in  the  axis  of  a  helix  of  wire  through  which  an 
electric  current  was  flowing.  The  intensity  of  the  field  was 
measured  at  seven  equidistant  points  along  the  axis  by  means  of 
a  dynamometer,  and  the  total  force  was  computed  from  (9). 

Examples. — (1)  Compare  Simpson's  one-third  rule  and  the  three-eighths 
rule  when  h  =  1,  with  the  result  of  the  integration  of 

f    x*dx.     Ansr.  i{(+  3)5  -  (-  3)5}  =  97-2, 

by  actual  integration  ;  for  Simpson's  one-third  rule, 

\U+  3)4  +  (-  3)4  +  4{(+  2)4  +  04  +  (-  2)4}  +2(1+1)3-  98. 
The  three-eighths  rule  gives 

J_J(x)dx  =  f%0  +  3yx  +  3y2  +  2yz  +  By,  +  Sy5  +  y6), 

f[(+  3)4  +  (-  3)4  +  3{(+  2)4  +  l4  +  (-  l)4  +  (-  2)4}  +  2  x  0]  =  99. 
The  errors  are  thus  as  8  :  18,  or  as  4  :  9.  A  great  number  of  cases  has  been 
tried  and  it  is  generally  agreed  that  the  parabolic  rule  with  an  odd  number  of 
ordinates  always  gives  a  better  arithmetical  result  than  if  one  more  ordinate  is 
employed.  Thus,  Simpson's  rule  with  five  ordinates  gives  a  better  result  than 
if  six  ordinates  are  used. 


8  HO. 


INFINITE  SERIES  AND  THEIR  USES. 


339 


(2)  On  plotting /(x)  in  jf(x)dx,  it  is  found  that  the  lengths  of  the  ordinates 
3  cm.  apart  were:  14-2,  14-9,  15-3,  16*1,  145,  14-1,  13-7  cm.  Find  the 
numerical  value  of  the  integral.  Ansr.  263'9  sq.  cm.  by  Simpson's  one-third 
rule.    Hint.  From  (7), 


/; 


f(x)dx  =  (14-2  +  13-7)  +  4(14-9  +  15-1  +  14-1)  +  2(15-3  +  14-5). 


An  objection  to  these  rules  is  that  more  weight  is  attached  to 
some  of  the  measurements  than  to  others.  E.g.,  more  weight  is 
attached  to  y}i  yv  and  yb  than  to  y,2  and  yi  in  applying  Weddle's  rule. 

II.  Trapezoidal  Formula. 

Instead  of  assuming  each  strip  to  be  the  sum  of  a  trapezium 
and  a  parabolic  segment,  we  may  suppose 
that  each  strip  is  a  complete  trapezium.  In 
Fig.  136,  let  AN  be  a  curve  whose  equation 
19  V  =  f(x)  ;  AA't  BB\  . .  .  perpendiculars 
drawn  from  the  ic-axis.  The  area  of  the 
portion  ANN' A'  is  to  be  determined.  Let 
OB'  -  OA'  =  OC  -  OB'  =  ...  =  h.  It  fol- 
lows from  known  mensuration  formulae,  (11), 
page  604, 

*rea  ANN9 A'-  =  tfi{AA'  +  BB')  +  ih(B'B  +  C'G)  +  . . . ; 

=  ih(AA'  +  2BB'  +  2CC  +  ...  +  2MM'  +  NN) ; 
=  h(i  +  1  +  1  +  . . .  +  1  +  1  +  J),     .         .       (10) 
where  the  coefficients  of  the  successive  ordinates  alone  are  written. 
The  result  is  known  as  the  trapezoidal  rule.  ■      „ 

Let  x0,  xlf  x2,  ... ,  xn,  be  the  values  of  the  abscissae  correspond- 
ing with  the  ordinates  y0,  yv  y2,  . . . ,  yn>  then, 


0  A1 


f{x).dx  =  \{Xl-xQ){y, 


If  ^ 


Xn     Xn     —     X-i      = 


yl)  +  ..-  +  Wn-xn_l){yn_1  +  yn).      (11) 
=  h,  we  get,  by  multiplying  out, 


f(x) .  dx  =  h{i(y0  +  yn)  +  y2  +  y,  +  . . .  +  y^.  (12) 

The  trapezoidal  rule,  though  more  easily  manipulated,  is  not 
quite  so  accurate  as  those  rules  based  on  the  parabolic  formula  of 
Newton  and  Cotes. 

The  expression, 

A.rea  ANNA'  =  h(&  +  if  +  1  +  1  +  .  .  .   +  1  +  1  +  If  +  T*_),    (13) 

or, 


340 


HIGHER  MATHEMATICS. 


§110. 


is  said  to  combine  the  accuracy  of  the  parabolic  rule  with  the 
simplicity  of  the  trapezoidal.     It  is  called  Durand's  rule. 

/wdx 
—-,  by  the  approximation  form- 
2     x 

ulas  (7),  (10)  and  (13),  assuming  h  =  1,  n  =  8.  Find  the  absolute  value  of  the 
result  and  show  that  these  approximation  formulae  give  more  accurate 
results  when  the  interval  h  is  made  smaller.  Ansr.  (7)  gives  1*611,  (10) 
gives  1*629,  (13)  gives  1*616.     The  correct  result  is  1*610. 

(2)  Now  try  what  the  trapezoidal  formula  would  give  for  the  integration 
of  Ex.  (2),  page  339.     Ansr.  263*55.    Hint.  From  (12) 

3{£(14*2  +  13*7)  +  14*9  +  15*3  +  15*1  +  14*5  +  14*1}. 

G.  Lemoine  (Ann.  Chim.  Phys.,  [4],  27,  289,  1872)  encountered 
some  non-integrable  equations  during  his  study  of  the  action  of 
heat  on  red  phosphorus.  In  consequence,  he  adopted  these 
methods  of  approximation.  The  resulting  tables  "calculated" 
and  "  observed"  were  very  satisfactory. 

Double  integrals  for  the  calculation  of  volumes  can  be  evaluated 
by  a  double  application  of  the  formula.  For  illustrations,  see  C.  W. 
Merrifield's  report  "  On  the  present  state  of  our  knowledge  of  the 
application  of  quadratures  and  interpolation  to  actual  calculation," 
B.  A.  Reports,  321,  1880. 

III.  Mid-section  Formula. 
A  shorter  method  is  sometimes  used.     Suppose  the  indicator 


diagram  (Fig.  137)  to  be  under  investigation.  Drop  perpendiculars 
PM  and  QN  on  to  the  "Atmospheric  line"  MN;  divide  MN  into 
n  equal  parts.     In  the  diagram  n  =  6.     Then  measure  the  average 


§  111.             INFINITE  SERIES  AND  THEIR  USES.  341 

length  ab,  cd,  ef,...  of  each  strip ;  add,  and  divide  by  n.  Alge- 
braically, if  the  length  of  ab  =  y1 ;  cd  =  y&;  ef  =  y5;  ... 

Total  area  =    -fa  +  yz  +  yb  +  . .  .).             .  (15) 


§  111.    Integration  by  Infinite  Series. 

Some  integrations  tax,  and  even  baffle,  the  resources  of  the 
most  expert.  It  is,  indeed,  a  common  thing  to  find  expressions 
which  cannot  be  integrated  by  the  methods  at  our  disposal.  We 
may  then  resort  to  the  methods  of  the  two  preceding  sections,  or, 
if  the  integral  can  be  expanded  in  the  form  of  a  converging  series 
of  ascending  or  descending  powers  of  x,  we  can  integrate  each 
term  of  the  expanded  series  separately  and  thus  obtain  any  desired 
degree  of  accuracy  by  summing  up  a  finite  number  of  these  terms. 

If  f{x)  can  be  developed  in  a  converging  series  of  ascending 
powers  of  x,  that  is  to  say,  if 

f(x)  =  a0  +  axx  +  a^c2  +  a3x*  + (1) 

By  integration,  it  follows  that 

if(x)dx  =  j(a0  +  aYx  +  a2x2  +  . .  .)dx ; 
=  ja0dx  +  fa^dx  +  ja<fl2dx  + 
=  a0x  +  \aYx2  +  \a$?  +  . . . ; 
=  x(a0  +  \axx  +  \a<p2  +  ...)  +  0.    .         (2) 
Again,  if  f(x)  is  a  converging  series,  jf(x) .  dx  is  also  convergent. 
Thus,  if 

f(x)  =  1  +  x  +  x2  +  x2  +  . . .  +  x"-1  +  xn  +  . . . ,  (3) 

11  11 

f(x) .  dx  =  x  +  yX2  +  3#3  +  . . .  +  -an  +  ^jna;n+1  +  •  •  •     (4) 

Series  (3)  is  convergent  when  x  is  less  than  unity,  for  all  values  of 
n.  Series  (4)  is  convergent  when  ^TiX,  and  therefore  when  x  is 
less  than  unity.  The  convergency  of  the  two  series  thus  depends 
on  the  same  condition,  a?>l.  If  the  one  is  convergent,  the  other 
must  be  the  same. 

If  the  reader  is  able  to  develop  a  function  in  terms  of  Taylor's 
series,  this  method  of  integration  will  require  but  few  words  of 
explanation.  One  illustration  will  suffice.  By  division,  or  by 
Taylor's  theorem, 

(1  +  a2)"1  =  1  -  x2  +  x*  -  x*  +  ... 
Consequently. 


f 


342  HIGHER  MATHEMATICS.  §111. 

1 1 2  =  \dx  -  \x2  .dx  +  p4  .  dx  -  I x6 .  dx  +  . . -. ; 

.-.  1(1  +  x2)-Hx  -  x  -  lx*  +  \xb  -  . . .  =  tan  ~lx  +  C, 
from  (6),  page  284. 

/dx  a;3        1.3  a;' 
-t  =  x  +  g— g  +  2'4       +  . . .  =  sin  -  *x  +  C 

i«x   «,         f    f7a;  o    /- — f^       !    sin2aj      1-3    siu4aj  \ 

<2)  show i T^Tx  =  WsmaV  +  a*  —  +  O-  -T-  +  -•)  +  c- 

/o  a;3  a;5  a;7 

*-*»<*«>  =  ^  -  He  +  17275  -  1727^7  +  -..  +  & 

The  two  following  integrals  will  be  required  later  on.     k2  is  less  than  unity. 

(6)  How  would  you  propose  to  integrate  [  (1  -  a;)-1  log  x  .  dx  in  series? 

o 
Hint.    Develop  (1  -  a?)-1  in  series.      Multiply  through  with  log  x.dx.     Then 
integrate  term  by  term.     The  quickest  plan  for  the  latter  operation  will  be 
to  first  integrate  Ja^log  x .  dx  by  parts,  and  show  that 

f  xn  +  1 (  1     \ 

ja-logz.^^^loga;-— XJ. 

f1  loga;  _  /l       1       1       1  \ 

ins   n,        Jl  \      1  2        1     x       UfxY 

(7)  ShowthatsI^  =  -+31.2  +  ^-2j    +... 

Then,  remembering  that  2  sin2£a:  =  1  -  cos  xt  (35)  page  612,  show  that 
f     x.dx  1  /  a;3         lx>  \      _ 

JVl-cosa;=72V2g  +  3~6  +  lM00  +  ---J  +  C- 

(8)  Show  j  un-g-dB-Lv  3  "  6(a)  7  +  120(2)  IT  "  •"  J0 ' 

=  0-5236  -  0-0875  +  0-0069  -  0-0003  +  . . .  =  0-446. 

We  often  integrate  a  function  in  series  when  it  is  a  compara- 
tively simple  matter  to  express  the  integral  in  a  finite  form.  The 
finite  integral  may  be  unfitted  for  numerical  computations.  Thus, 
instead  of 

dG       mm,     ^         7/      T       dG  ^i       °  +  *~T        ,Cx 

S.  Arrhenius  (Zeit.  phys.  Chem.,  1,  110,  1887)  used 

,      1     1     xf  a      1  \ 

kt "  a  "  ci  ~  Aw  ~c?)      •     •     (6) 

because  a;  being  small  in  comparison  with  C,  (5)  would  not  give 


§  111.  INFINITE  SEKIES  AND  THEIR  USES.  343 

accurate  results  in  numerical  work,  on  account  of  the  factor  x~l, 
and  in  (6)  the  higher  terms  are  negligibly  small.  Again,  the 
ordinary  integral  of 

dx      . ,  .  „         vo     7  1       ((a  -  b)x       .      a(b  -  x)} 

£-*(«-  x)  (b  -  xf;  ht  =  jj-jpl^-l  +  log  jj^J, 

from  (9),  page  221,  does  not  give  accurate  results  when  a  is  nearly 
equal  to  b,  for  the  factor  (a  -  b)~2  then  becomes  very  great.  We 
can  get  rid  of  the  difficulty  by  integration  in  series.  Add  and  sub- 
tract (b  -  x)-3  dx  to  the  denominator  of 

dx  V     1  1  l_]dx, 

(a-x)(b-xf     [_{b-xf^{a-x)(b-xf     (b-xfA      ' 

_  rj *-6f         1         }ldx . 

~|_(&-z)3     b-x[(a-x)(b-xfjj      ' 

_rj ^/_j <±z±      l     )idx. 

l(b-xf    b-x\(b-xf    b-x'(a-x)(b-xffju"L, 

f     1  *-&      (a-bf    (a-J>Y  ldx 

l(b-xf    (b-xf^^-xf    (b-xf  J 

This  is  a  geometrical  series  with  a  quotient  (a  -  b)/(b  -  x)  and 
convergent  when  (a  -  b)  <  (b  -  x) ;  that  is  when  a  <  b,  or  when 
a  is  only  a  little  greater  than  b.  Now  integrate  term  by  term ; 
evaluate  the  constant  when  x  =  0  and  t  =  0 ;  wo  get 

*  =  ~t  [_2{{b  -  xf  ~  PJ  3~{(b  -  xf  ~  &)  +  "  ']' 

The  first  term  is  independent  of  a  -  b,  and  it  will  be  sufficiently 
exact  for  practical  work. 
Integrals  of  the  form 

e~x2dx;  or,      e~x2dx      ...        (7) 
Jo  Jo 

are  extensively  employed  in  the  solution  of  physical  problems. 

E.g.,  in  the  investigation  of  the  path  of  a  ray  of  light  through 

the  atmosphere  (Kramp) ;  the  conduction  of  heat  (Fourier) ;  the 

secular  cooling  of  the  earth  (Kelvin),  etc.     One  solution  of  the 

important  differential  equation 

IV       d2F 

is  represented  by  this  integral.  Errors  of  observation  may  also  be 
represented  by  similar  integrals.  Glaisher  calls  the  first  of  equa- 
tions (7)  the  error  function  complement,  and  writes  it,  "  erfc  a?  "  ; 


344  HIGHER  MATHEMATICS.  §  111. 

and  the  second,  he  calls  the  error  function,  and  writes  it,  "  erf  x  ". 
J.  W.  L.  Glaisher  (Phil.  Mag.  [4],  42,  294,  421,  1871)  and  E. 
Pendlebury  (ib.,  p.  437)  have  given  a  list  of  integrals  expressible 
in  terms  of  the  error  function.  The  numerical  value  of  any  in- 
tegral which  can  be  reduced  to  the  error  function,  may  then  be 
read  off  directly  from  known  tables.  See  also  J.  Burgess,  Trans. 
Boy.  Soc.  Edin.,  39,  257,  1898. 

We  have  deduced  the  fact,  on  page  240,  that  functions  of  the 
same  form,  when  integrated  between  the  same  limits,  have  the  same 
value.     Hence,  we  may  write 

e~x2dx=\   e-v2dy; 
Jo  Jo 

.-.  [  e~x2dx\  e-*2dy=[   f  e-^  +  ^dxdy=\[  e~'dx\.     (8) 
Jo  Jo  JoJo  Do  J 

Now  put 

y  =  vx ;  i.e.,  dy  =  xdv. 

Our  integral  becomes 


n 

Jo  Jc 


xe-*n  +  *>dxdv.       ...        (9) 

o 

It  is  a  common  device  when  integrating  exponential  functions  to 
first  differentiate  a  similar  one.  Thus,  to  integrate  jxe  -"^dx,  first 
differentiate  e~ax2,  and  we  have  d(e~ax2)  =  -  2axe~ax2dx.  From 
this  we  infer  that 

[d{e  ~  **2)  -  -  2a  \xe  -  ax2dx  ;  or,  \xe  -  ax2dx  =  -  ~e  ~  ax2  +  G. 
Applying  this  result  to  the  "  dx  "  integration  of  (9),  we  get 

J  „ Xe  dX      L   2(1  +  «»)    Jo      W  +  «9 

since   the   function   vanishes   when  x  is    oo.      Again,  from  (13) 
page  193,  the  "  dv "  integration  becomes 


f,"s(IT^-|jtan"1,']0"- 


Consequently,  by  combining  the  two  last  results  with  (8)  and  (9), 
it  follows  that 

[£«-"**»]*-  |i «. £«-■**" '-  T-    •  (10) 

This  fact  seems  to  have  been  discovered  by  Euler  about  1730. 
There  is  another  ingenious  method  of  integration,  due  to  Gauss, 


§  111.  INFINITE  SERIES  AND  THEIR  USES.  345 

in  which  the  penultimate  integral  of  equations  (8)  is  transformed 
into  polar  coordinates  and  the  limits  are  made  so  as  to  just  cover 
one  quadrant. 

IT 

.*.  |°°[  e-^+^dxdy  =  Pf  e~r2r.d0.dr  =  gfV^r.  dr  =  J  etc. 

This  important  result  enables  us  to  solve  integrals  of  the 
form  je  ~  *^xndx,  for  by  successive  reduction 

JV-v .  dx  =  (W-1)2,^1;,3)-2JV^-  *».  •   <"> 

when  n  is  odd ;  and,  when  n  is  even 

JV-V . <te  =  (W-1)(^8)-1J^a-<to.  .        (12) 

All  these  integrals  are  of  considerable  importance  in  the  kinetic 
theory  of  gases,  and  in  the  theory  of  probability.  In  the  former 
we  shall  meet  integrals  like 

— t=-     e~*2x*.dx;  and,  -t-     e~x2x*.dx.       .         (13) 

S/TT     Jo  V7rj0 

From  (12),  the  first  one  may  be  written  %Nma2;  the  latter  2Na/  \Ar. 
If  the  limits  are  finite,  as,  for  instance,  in  the  probability  in- 
tegral, 

P  =  A (Vo^te) ;  .-.  P  =  -^Je-'^dt, 

by  putting  hx  =  t.  Develop  e  ■"  *2  into  a  series  by  Maclaurin's 
theorem,  as  just  done  in  Ex.  (3)  above.     The  result  is  that 

may  be  used  for  small  values  of  t  For  large  values,  integrate  by 
parts, 

•'•  ~\e"2dt  "  e"2(i  "  A  +  i  "  ^?+  •") 

By  the  decomposition  of  the  limits,  (4),  page  241,  we  get 

[  e~*dt -[' *~*dt -[*  *-*& 

Jo  Jo  J* 

The  first  integral  on  the  right-hand  side  =  \  J-rr.     Integrating  the 


346  HIGHER  MATHEMATICS.  §  112. 

second  between  the  limits  oo  and  t 

e~t2(         1        1.3       1.3.5  \ 

t  sfX         x*  +  (2^)2        (2t*y   +  ' '  •/  (15) 

This  series  converges  rapidly  for  large  values  of  t.  From  this  ex- 
pression the  value  of  P  can  be  found  with  any  desired  degree  of 
accuracy.     These  results  are  required  later  on. 

§  112.    The  Hyperbolic  Functions. 

I  shall  now  explain  the  origin  of  a  new  class  of  functions,  and 
show  how  they  are  to  be  used  as  tools  in  mathematical  reasoning. 
We  all  know  that  every  point  on  the  perimeter  of  a  circle  is  equi- 
distant from  the  centre;  and  that  the  radius  of  any  given  circle 
has  a  constant  magnitude,  whatever  portion  of  the  arc  be  taken. 
In  plane  trigonometry,  an  angle  is  conveniently  measured  as  a 
function  of  the  arc  of  a  circle.  Thus,  if  V  denotes  the  length  of 
an  arc  of  a  circle  subtending  an  angle  0  at  the  centre,  /  the  radius 
of  the  circle,  then 

-  Length  of  arc  V 

Length  of  radius        r' 

This  is  called  the  circular  measure  of  an  angle  and,  for  this  reason, 
trigonometrical  functions  are  sometimes  called  circular  functions. 
This  property  is  possessed  by  no  plane  curve  other  than  the  circle. 
For  instance,  the  hyperbola,  though  symmetrically  placed  with 
respect  to  its  centre,  is  not  at  all  points  equidistant  from  it.  The 
same  thing  is  true  of  the  ellipse.     The  parabola  has  no  centre. 

If  I  denotes  the  length  of  the  arc  of  any  hyperbola  which  cuts 
the  #-axis  at  a  distance  r  from  the  centre,  the  ratio 

I 
u  =  -, 
r 

is  called  an  hyperbolic  function  of  u,  just  as  the  ratio  V\r'  is  a 
circular  function  of  0.  If  the  reader  will  refer  to  Ex.  (5),  page  247, 
it  will  be  found  that  if  I  denotes  the  length  of  the  arc  of  the  rect- 
angular hyperbola 

x2  -  2/2  =  a*,      .         .         .         .         (1) 

between  the  ordinates  having  abscissas  a  and  x, 

But  this  relation  is  practically  that  developed  for  cos  x,  on 
page  286,  ix,  of   course,  being  written  for  u.      The  ratio  xja  is 


§112. 


INFINITE  SERIES  AND  THEIR  USES. 


347 


defined  as  the  hyperbolic  cosine  of  u.  It  is  usually  written 
cosh  u,  or  hycos  u,  and  pronounced  "coshw,"  or  "  h-cosine u'\ 
Hence, 


u~ 


u* 


coshw  =  \{eu  +  e~u)  =  1  +  ^  +  jj  + 


2!    '    4! 


(2) 


In  the  same  way,  proceeding  from  (1),  it  can  be  shown  that 


2  +  e~2u 


-v 


elu  -246" 


which  reduces  to 


-K^-a-X 


a  relation  previously  developed  for  i  sin  a?.     The  ratio  y/a  is  called 
the  hyperbolic  sine  of  u,  written  sinh  u,  or  hysin  u.     As  before 


sinh  u  =  \(eu  -  e~u)  =  w  +  3]  +  5]  + 


(3) 


The  remaining  four  hyperbolic  functions,  analogous  to  the 
remaining  four  trigonometrical  functions,  are  tanh  u,  cosech  u, 
sech  u  and  coth  u.  Values  for  each  of  these  functions  may  be 
deduced  from  their  relations  with  sinh  u  and  cosh  u.     Thus, 


sinh  u  - 

tanh  u  =  - — r—  ;  seen  u  = 


coth  u 


cosh^ 

1 
*  tanh  u 


;  cosech  u 


cosh  u ' 
1 


sinh  Uj 


(4) 


Unlike  the  circular  functions,  the  ratios  x/a,  y/a,  when  referred 
to  the  hyperbola,  do  not  represent 
angles.  An  hyperbolic  function  ex- 
presses a  certain  relation  between  the 
coordinates  of  a  given  portion  on  the 
arc  of  a  rectangular  hyperbola. 

Let  0  (Fig.  138)  be  the  centre  of 
the  hyperbola  APB,  described  about 

the  coordinate  axes  Ox,  Oy.     From  M      /A  M 

any  point  P(x,  y)  drop  a  perpen-  FlG-  138- 

dicular  PM  on  to  the  x-axis.     Let  OM  =  x,  MP  =  y,  OA  =  a. 

.'.  coshu  =  x/a;  sinhw  =  y/a. 

For  the  rectangular  hyperbola,  x2 


=    /7.2 


a^cosh^u  -  a2sinh2w  = 


y*  =  a".     Consequently, 
or   cosh%  -  sinh%  =  1. 


348 


HIGHER  MATHEMATICS. 


§112. 


The  last  formula  thus  resembles  the  well-known  trigonometrical 
relation :  cos2#  +  sin2#  =  1.  Draw  FM  a  tangent  to  the  circle 
AF  at  P.'  Drop  a  perpendicular  FM  on  to  the  sc-axis.  Let  the 
angle  MOF  =  6. 

.'.  x/a  =  sec  0  =  cosh u ;  y/a  =  tan  6  =  sinh u.     .         (6) 

I.  Conversion  Formula. — Corresponding  with  the  trigonometri- 
cal formulae  there  are  a  great  number  of  relations  among  the 
hyperbolic  functions,  such  as  (5)  above,  also 

cosh  2x  =  1  +  2  sinh2#  =  2  cosh2#  +  1.  .         (7) 

sinh  x  -  sinh  y  =  2  cosh  \{x  +  y) .  sinh  \{x  -  y),  .  (8) 
and  so  on.  These  have  been  summarized  in  the  Appendix,  "  Col- 
lection of  Reference  Formulae  ". 

II.  Graphic  representation  of  hyperbolic  functions. — We  have 

seen  that  the  trigonometrical  sine, 
cosine,  etc.,  are  periodic  functions. 
The  hyperbolic  functions  are  ex- 
ponential, not  periodic.  This  will 
be  evident  if  the  student  plots 
the  six  hyperbolic  functions  on 
squared  paper,  using  the  nu- 
merical values  of  x  and  y  given 
in  Tables  IV.  and  V.  I  have 
done   this   for   y  =  cosh  x,   and 

y  =  sechz    in   Fig.  139.       The 
graph  of  y  =  cosh  x,  is  known  in  statics  as  the  "  catenary  ". 

III.  Differentiation  of  the  hyperbolic  functions. — It  is  easy  to 
see  that 

#inh£)      d{\{f-e-')} 
~dx~'  =  Tx =*(«*  +  e~)  =  cosh  x- 

We  could  get  the  same  result  by  treating  sinh  u  exactly  as  we 
treated  sin  x  on  page  48,  using  the  reference  formulae  of  page  611. 
For  the  inverse  hyperbolic  functions,  let 

y  =  sinh_1#;  .*.  dx/dy  =  coshy. 
From  (5)  above,  it  follows  that 

cosh?/  =  \/smh2y  +  1 ;  .-.  cosh 3/  =   Jx2  +  1 ; 
and,  from  the  original  function,  it  follows  that 

dy  _       1 
*c      six2  +  T 


Fig.  139.- 


-Graphs  of  cosh  x  and 
seen  x. 


§112. 


INFINITE  SERIES  AND  THEIR  USES. 


349 


IV.  Integration  of  the  hyperbolic  functions. — A  standard  col- 
lection of  results  of  the  differentiation  and  integration  of  hyperbolic 
functions,  is  set  forth  in  the  following  table  : — 


Table  III. — Standard  Integrals. 


Function. 


y  =  sinh  x. 
y  =  cosh  x. 
y  =  tanh  x. 
y  =  coth  x. 
y  =  sech  x. 
y  =  cosech  x. 
y  =  sinh  -  lx. 
y  =  cosh  -  lx. 
y  =  tanh  -  xx. 
y  —  coth  -  lx. 
y  =  sech  -  lx. 
y  =  cosech  -  xx. 


Differential  Calculus. 


dx 


=  cosh  x. 


dx  =  smhx- 


=  sech2a\ 


-=-  =  —  cosech2x. 

dy  _       sinh  x 
dx  cos  h2  x 

dv  cosh  x 

dx  ~  ~  sinh2  x 
dy  1 

dx      six2  +  1 ' 
dy  1 


dx      s/x^^~l 
dx 


,  JB<1. 


1-x2 

dy  _     1      3,^-, 
dy_  =  _        1 

*»  Xn/x24-1' 


Integral  Calculus. 


/  cosh  x  dx    =  sinh  x. 

/  sinh  x  dx    =  cosh  #. 

sech2ic  dx    =  tanh  x.     , 

cosech2£E  dx  =  -  coth  a;  . 

— — —  da;  =  -  sech  x.  . 
cosh-  x 

/-— — —  dx  =  -  cosech  x. 
sinh2  x 

I 


dx 


sj. 


=  cosh  -  lx. . 


f  dx 
Jl-x2 

f  dx 
J  x*  -  1 

/dx 

f       dx 


tanh  -  lx. . 

coth  -  lx.  . 
=  -  sech  x.  . 
=  -  cosech  x. 


(9) 
(10) 

(11) 
(12) 
(13) 
(14) 

(15 
(16) 
(17) 
(18) 
(19) 
(20) 


Examples. — When  integrating  algebraic  expressions  involving  the  square 

root  of  a  quadratic,  hyperbolio  functions  may  frequently  be  substituted  in 

place  of  the  independent  variable.      Such  equations  are  very  common  in 

electrotechnics.    It  is  convenient  to  remember  that  x — a  tanh  u,  or  a;=tanhw 

may  be  put  in  place  of  a2  -  a2,  or  1  -  x2  ;  similarly,  x  =  a  cosh  u  may  be  tried 

in  place  of  six2  -  a2 ;  x  =  a  sinh  u,  for  six2  +  a2. 

(1)  Evaluate  j  six2  +  a2  .  dx.     Substitute  x  =  a  sinh  u  for  six2  +  a2,  dx= 

a  cosh  u  .  du.     From  (5),  above  ;  and  (22)  and  (24),  page  613, 

.'.  \  six2  +  a2  ,dx  =  j  sja2{l  +  sinh'%) .  a  cosh  u  .  du  =  a2jcosh2u  .  du 

=  £a2J(cosh2tt  +l).du; 

=  Ja2sinh2w  +  \a2u  =  £a  sinh  u .  a  cosh  u  +  \a2u. 

<m  lXsJ{x'2  +  a2)  +  ^a2  sinh-  lxja. 

And  since  we  are  given  sinh-J2/  =  \og(y  +  sly'2  +  1)  on  page  613,  (31), 

'    , 5     ,       xja2  +  x2  .  a:\„„x 

six'2  +  a2 .  dx= o 


■i- 


+  -log: 


+  n/ 


x*  +  a* 


+  C 


350  HIGHER  MATHEMATICS.  §  112. 

(2)  Now  try  and  show  that  J"  six'1  -  a2  .  dx  furnishes  the  result 
%x.sJx2  -  a2  -  £a2Jog  (x  +  six1  -  a?)Ja  +  C,  when  treated  in  a  similar  manner 
by  substituting  x  =  a  cosh  u. 

(3)  Find  the  area  of  the  segment  OPA  (Fig.  138)  of  the  rectangular 
hyperbola  x2  -  if  =  1.     Put  x  =  cosh u;  y  =  sinh u.     From  (6), 

.-.  Area  APM  =  I  y  .  dx  =   /  sinh2w  .du  =  W  (cosh  2u  -  1)  .  du. 

Area  4P.M"  =  J  sinh  2u  -  ^.  . \  Area  OPA  =  £  Area  PM  .OM-  Area  ^PM  =  \u. 
Note  the  area  of  the  circular  sector  OP' A  (same  figure)  =  %e,  where  0  is  the 
angle  AOP'. 

(4)  Rectify  the  catenary  curve  y  =  c  cosh  x\c  measured  from  its  lowest  point 
Ansr.  Z=csinh  (xjc).     Note  1  =  0  when  x  =  0,  .*.  C  =  0.     Hint.  (26),  page  613. 

(5)  Rectify  the  curve  y2  =  Aax  (see  Ex.  (1),  page  246).  The  expression 
sl(l-{-alx)dx has  to  be  integrated.  Hint.  Substitute  x  =  a sinh2^.  2ajcosh2u.du 
remains.  Ansr.  =  aj(l  +  cosh2u)dii,  or  a(u  +  £sinh2w).  At  vertex,  where 
x  =  0,  sinh^  =  0,  O  =  0.  Show  that  the  portion  bouDded  by  an  ordinate 
passing  through  the  focus  has  I  =  2-296<z.  Hint.  Diagrams  are  a  great  help 
in  fixing  limits.  Note  x  =  a,  .'.  sinhw  =  1,  cosh  u  =  si 2,  from  (5).  From 
(31),  page  613,  sinh  -  ll  =  u  =  log(l  +  \/2).  From  (20),  page  613,  sinh  2u  = 
2  sinh  u .  cosh  u. 

I  =  a  Yu  +  £  sinh  2wT  =a(u  +  sinh  u  .  cosh  u)  =  a  ("log(l  +  \/2)  +  J2~\. 

Use  Table  of  Natural  Logarithms,  Appendix  II.     Ansr.  2-296a. 

(6)  Show  that  y  =  A  cosh  mx  +  B  sinh  mx,  satisfies  the  equation  of 
d2yjdx2  =  m2y,  where  m,  .4  and  P  are  undetermined  constants.  Hint.  Dif- 
ferentiate twice,  etc.  Note  the  resemblance  of  this  result  with  y  =  A  cosnx  + 
B  sin  nx,  which  furnishes  d2y/dx2  =  -  n2y  when  treated  in  the  same  way. 

(7)  In  studying  the  rate  of  formation  of  carbon  monoxide  in  gas  producers, 
J.  K.  Clement  and  C.  N.  Haskins  (1909),  obtained  the  eauation 

with  the  initial  condition  that  x  =  0  when  t  =  0.     Integrating,  and 

log£±5=2«*. 
Solving  for  x,  we  get 

Qat  —  Q—ai 

V.  Numerical  Values  of  hyperbolic  furoctions. — Table  IV. 
(pages  616,  617,  and  618)  contains  numerical  values  of  the  hyper- 
bolic sines  and  cosines  for  values  of  x  from  0  to  5,  at  intervals  of 
OOl.  They  have  been  checked  by  comparison  with  Des  Ingenieurs 
Taschenbuch,  edited  by  the  Hiitte  Academy,  Berlin,  1877.  The 
tables  are  used  exactly  like  ordinary  logarithm  tables.  Numerical 
values  of  the  other  functions  can  be  easily  deduced  from  those  of 
sinh  x  and  cosh  a;  by  the  aid  of  equations  (4). 


§  112.  INFINITE  SERIES  AND  THEIR  USES.  351 

Example.— The  equation  I  =  x{e>i2x  -  e-*/2*),  represents  the  relation  be- 
tween the  length  I  of  the  string  hanging  from  two  points  at  a  distance  s  apart 
when  the  horizontal  tension  of  the  string  is  equal  to  a  length  x  of  the  string. 
Show  that  the  equation  may  be  written  in  the  form  11^  -  10  sinhw  =  0  by 
writing  u  =  10  jx  and  solved  by  the  aid  of  Table  IV.,  page  616.  Given  I  =  22, 
s  =  20.  .*.  x  =  13-16.  Hint.  Substitute  s  =  20,  I  =  22,  u  =  10/aj,  and  we  get 
22u  -  10(e«  -  e-M)  =  0;  .\  Hw  -  10  x  ${eu  -  e-»)  =  0;  etc.  u  is  found  by 
the  method  described  in  a  later  chapter ;  the  result  is  u  =  0-76.  But 
x  =  10/w,  etc. 

VI.  Dcmoivre's  theorem. — We  have  seen  that 

cosz  =  J(e<*  +  e"ix);  isina  =  i(elX  -  e~iX), 
e^x  =  cos x  +  i sin x ; . and  e~lX  =  cos x  -  t sin x. 
If  we  substitute  nx  for  x,  where  n  is  any  real  quantity,  positive  or 
negative,  integral  or  fractional, 

cosnx  =  \{einx  +  e~'nx);ism?ix=:  ^{eLnx  -  e~Lnx). 
By  addition  and  subtraction  and  a  comparison  with  the  preceding 
expressions ;  we  get 

cos  nx  +  t  sin  nx  =  eLnx     =  (cos  x  +  t  sin  x)n\  ^n 

cos  nx  -  i  sin  nx  =  e~ in*  =  (cos  x  -  t  sin x)nj 
which  is  known  as  DemoiYre's  theorem.     The  theorem  is  useful 
when  we  want  to  express  an  imaginary  exponential  in  the  form  of 
a  trigonometrical  series,  in  certain  integrations,   and  in  solving 
certain  equations. 

Examples. — (1)  Verify  the  following  result  and  compare  it  with  Demoivre's 
theorem :  (cos  x  +  i  sin  x)2  =  (cos2x  -  sin2x)  +  2t  sin  x  .  cos  x  =  cos  2x  +  i  sin  2x. 

(2)  Show  e*  +  '0  =  e*e#  =  e"(cos  fi  +  isin  j8). 

(3)  Show  Je«*(cos  fix  +  t  sin  fix)ax  =  eo*(cos  fix  +  i  sin  fix)  I  (a  +  ifi) ', 

_    „,.(cosfrc  +  i  sin  foe)  (q  -  ifi) , 
~CaX  a'  +  fi2 

(a  cos  fix  +  fi  sin  fix)  +  i(  -  fi  cos  fix  +  a  sin  fix) 

~  CaX a2 +  0* +  G' 

by  separating  the  real  and  imaginary  parts. 

For  a  fuller  discussion  on  the  properties  and  uses  of  hyperbolic 
functions,  consult  G.  Chrystal's  Algebra,  Part  ii.,  London,  1890 ; 
and  A.  G.  Greenhill's  A  Chapter  in  the  Integral  Calculus,  London, 
1888. 


CHAPTEE  VI. 

HOW  TO  SOLVE  NUMEEICAL  EQUATIONS. 

"The  object  of  all  arithmetical  operations  is  to  save  direct  enumeration. 
Having  done  a  sum  once,  we  seek  to  preserve  the  answer  foi 
future  use  ;  so  too  the  purpose  of  algebra,  which,  by  substituting 
relations  for  values,  symbolizes  and  definitely  fixes  all  numerical 
operations  which  follow  the  same  rule." — E.  Mach. 

§  113.  Some  General  Properties  of  the  Roots  of  Equations. 

The  mathematical  processes  culminating  in  the  integral  calculus 
furnish  us  with  a  relation  between  the  quantities  under  investiga- 
tion. For  example,  in  §  20,  we  found  a  relation  between  the 
temperature  of  a  body  and  the  time  the  body  has  been  cooling. 
This  relation  was  represented  symbolically  :  6  =  be'"*,  where  a  and 
b  are  constants.  I  have  also  shown  how  to  find  values  for  the 
constants  which  invariably  affect  formulae  representing  natural 
phenomena.  It  now  remains  to  compute  one  variable  when  the 
numerical  values  of  the  other  variable  and  of  the  constants  are 
known.  Given  b,  a,  and  0  to  find  t,  or  given  b,  a,  and  t  to  find  0. 
The  operation  of  finding  the  numerical  value  of  the  unknown 
quantity  is  called  solving  the  equation.  The  object  of  solving 
an  equation  is  to  find  what  value  or  values  of  the  unknown  will 
satisfy  the  equation,  or  will  make  one  side  of  the  equation  equal 
to  the  other.  Such  values  of  the  unknown  are  called  roots,  or 
solutions  of  the  equation. 

The  reader  must  distinguish  between  identical  equations  like 
(x  +  l)2  =  x2  +  2x  +  1, 
which  are  true  for  all  values  of  x,  and  conditional  equations  like 

#  +  1  =  8;  a;2  +  2x  +  1  =  0, 
which  are  only  true  when  x  has  some  particular  value  or  values, 
in  the  former  case,  when  x  =  7,  and  in  the  latter  when  x  =  -  1. 

352 


§  113.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS.  353 

An  equation  like 

a;2  +  2a>  +  2  ~  0, 
has  no  real  roots  because  no  real  values  of  x  will  satisfy  the  equa- 
tion. By  solving  as  if  the  equation  had  real  roots,  the  imaginary 
again  forces  itself  on  our  attention.  The  imaginary  roots  of  this 
equation  are  -  1  +  J  -  1,  or  -  1  +  t.  Imaginary  roots  in  an 
equation  with  real  coefficients  occur  in  pairs.  E.g.,  if  a  +  p  J  -  1 
is  one  root  of  the  equation,  a  -  f3  J  -  1  is  another. 
The  general  equation  of  the  nth  degree  is 

xn  +  axn  ~ 1  +  bxn  ~  2  +  . . .  +  qx  +  B  =  0.        .         (1) 

The  term  B  is  called  the  absolute  term.  If  n  «  2,  the  equation  is 
a  quadratic,  x2  +  ax  +  B  «  0 ;  ifn«3,  the  equation  is  said  to  be 
a  cubic  ;  if  n  =  4 ,  a  biquadratic,  etc.  If  xn  has  any  coefficient,  we 
can  divide  through  by  this  quantity,  and  so  reduce  the  equation  to 
the  above  form.  When  the  coefficients  a,  b,  . . .,  instead  of  being 
literal,  are  real  numbers,  the  given  relation  is  said  to  be  a  numer- 
ical equation.  Every  equation  of  the  nth  degree  has  n  equal  or 
unequal  roots  and  no  more — Gauss'  law.  E.g.,  xb  +  #4  +  x  + 1  =  0, 
has  five  roots  and  no  more. 

General  methods  for  the  solution  of  algebraic  equations  of  the 
first,  second  and  third  degree  are  treated  in  regular  algebraic  text- 
books; it  is,  therefore,  unnecessary  to  give  more  than  a  brief 
resume  of  their  most  salient  features.  We  nearly  always  resort  to 
the  approximation  methods  for  finding  the  roots  of  the  numerical 
equations  found  in  practical  calculations. 

After  suitable  reduction,  every  quadratic  may  be  written  in  the 
form : 

ax*  +  bx  +  c  =  0  ;  or,  x2  +  -x  +  -  =  0.        .        (2) 

a        a  v  ' 

If  a  and  &  represent  the  roots  of  this  equation,  x  must  be  equal  to 
a  or  /?,  where 

-  b  +  s/b2  -  lac             0      -6  -  n/6*  -  4ac         /OX 
a j- ;    and,0^ ^ (3) 

The  sum  and  produot  of  the  roots  in  (3)  are  therefore  so  related 
that  a  +  j3  —  -  b/a ;  a/3  —  c/a.  Hence  x2  -  (a  +  /3)x  +  a/3  =  0 ; 
or,  xs  -  (sum  of  roots)  x  +  product  of  roots  »  0 ;  (4)  if  one  of  the 
roots  is  known,  the  other  can  be  deduced  directly.  From  the 
second  of  equations  (2),  and  (4)  we  see  that  the  sum  of  the  roots 
is  equal  to  the  coefficient  of  the  second  term  with  its  sign  changed, 


354 


HIGHER  MATHEMATICS. 


§113. 


the  product  of  the  roots  is  equal  to  the  absolute  term.  If  a  is  a 
root  of  the  given  equation,  the  equation  can  be  divided  by  x  -  a 
without  remainder.  If  ft,  y,  . . .  are  roots  of  the  equation,  the 
equation  can  be  divided  by  (x  -  ft)  (x  -  y)  . . .  without  remainder. 
From  Gauss'  law,  therefore,  (2)  may  be  written 

(x  -  a)(x  -  ft)  =  0.  ...  (5) 
From  (3),  and  (4),  we  can  deduce  many  important  particulars  re- 
peoting  the  nature  of  the  roots l  of  the  quadratic.     These  are : 

Relations  between  the  Coefficients  of  Equations  and  their  Roots. 


Relation  between  the  Coefficients. 


The  Nature  of  the  Roots. 


/positive, 
zero, 
62  -  4oc  is  J  negative, 

perfect  square, 

I  not  a  perfect  square, 
a,  b,  c,  have  the  same  sign, 
a,  o,  differ  in  sign  from  c, 
a,  c,  differ  in  sign  from  b, 

a  =  0, 

6*0 

c  =  0 

c  =  0,  b  -  0 


real  and  unequal. 

real  and  equal. 

imaginary  and  unequal. 

rational  and  unequal.    . 

irrational  and  unequal. 

negative. 

opposite  sign.         .        . 

positive. 

one  root  infinite.    . 

equal  and  opposite  in  sign. 

one  root  zero. 

both  roots  zero. 


(6) 

(7) 
(8) 

(9) 
(10 

(11) 
(12) 
(13) 
(14) 
(15) 
(16) 
(17) 


On  account  of  the  important  rdle  played  by  the  expression 
b2  -  4ac,  in  fixing  the  character  of  the  roots,  "b2  -  4oc,"  is 
called  the  discriminant  of  the  equation. 

Examples. — (1)  In  the  familiar  equation  of  Guldberg  and  Waage 
K(a  -  x)  {b  -  x)  =  (c  +  x)  (d  +  x) 
found  in  most  text-books  of  theoretical  chemistry,  show  that 
K(a  +  b)  +  d  +  c 


V{ 


K(a  +  b)  +  d  +  c\  2      cd  +  Kab 


}' 


2(K  -  1)  ~    \  ^         2(K  -  1)         J      '      K-\ 

Hint.  Expand  the  given  equation ;  rearrange  terms  in  descending  powers  of 
aj;  and  substitute  in  the  above  equations  (2)  and  (3). 

(2)  If  v2  -  516*17t>  +  1852-6  =  0,  find  v.  This  equation  arises  in  Ex.  (4), 
page  362.  On  reference  to  equations  (2)  and  (3),  a  =  1 ;  6=  -  516*17 ;  c  =  1852*6. 
Hence  show  that  v  =  (516*17  ±  508 *94)2. 

(3)  The  thermal  value,  q,  of  the  reaction  between  hydrogen  and  carbon 
dioxide  is  represented  by  q  =  -  10232  +  0-168527  +  0-00101T2,  where  T  denotes 
the  absolute  temperature.  Show  T=  3100°  when  q =0.  Hint.  You  will  have  to 
reject  the  negative  root.  To  assist  the  calculation,  note  (0*1685) 2= 0*02839; 
4  x  10232  x  0*00101  m  41*33728 ;  n/41*36567  =  6*432. 


1  In  the  table,  the  words 
values  of  the  roots. 


"equal"  and    "unequal"  refer  to  the    numerical 


§  114.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS. 


355 


§  114.  Graphic  Methods  for  the  Approximate  Solution  of 
Numerical  Equations. 

In  practical  work,  it  is  generally  most  convenient  to  get  ap- 
proximate values  for  the  real  roots  of  equations  of  higher  degree 
than  the  second.  Cardan's  general  method — found  in  the  regular 
text- books — for  equations  of  the  third  degree,  is  generally  so  un- 
wieldy as  to  be  almost  useless.  Trigonometrical  methods  are 
better.  For  the  numerical  equations  pertaining  to  practical  work, 
one  of  the  most  instructive  methods  for  locating  the  real  roots,  is 
to  trace  the  graph  of  the  given  function.  Every  point  of  inter- 
section of  the  curve  with  the  x-axis,  represents  a  root  of  the 
equation.  The  location  of  the  roots  of  the  equation  thus  reduces 
itself  to  the  determination  of  the  points  of  intersection  of  the  graph 
of  the  equation  with  the  rc-axis.  The  accuracy  of  the  graphic  method 
depends  on  the  scale  of  the  diagram  and  the  skill  of  the  draughts- 
man.    The  larger  the  "  scale  "  the  more  accurate  the  results. 

Examples. — (1)  Find  the  root  of  the  equation  x  +  2  «=  0.  At  sight,  of 
course,  we  know  that  the  root  is  -  2.  But  plot  the  curve  y  »  x  +  2,  for 
values  of  y  when  -8,-2,-1,  0,  1,  2,  3,  are  suc- 
cessively assigned  to  x.  The  curve  (Fig.  140)  cuts 
the  x-axis  when  x  «=  -  2.  Hence,  x  =  -  2,  is  a  root 
of  the  equation. 

(2)  Locate  the  roots  of  x2  -  8*  +  9  -  0.  Pro- 
ceed as  before  by  assigning  successive  values  to  x. 
Roots  occur  between  6  and  7  and  1  and  2. 

(8)  Show  that  x3  -  6a;2  +  11a  -  6  -  0  has  roots 
in  the  neighbourhood  of- 1,  2,  and  8. 

(4)  Show,  by  plotting,  that  an  equation  of  an  odd  degree  with  real  co- 
efficients, has  either  one  or  an  odd  number  of  real  roots.  For  large  values  of 
x,  the  graph  must  lie  on  the  positive  side  of  the  x-axis,  and  on  the  opposite 
side  for  large  negative  values  of  x.  Therefore  the  graph  must  cut  the  x-axis 
at  least  once  ;  if  twice,  then  it  must  cut  the  axis 
a  third  time,  etc. 

(6)  Prove  by  plotting  if  the  results  obtained 
by  substituting  two  numbers  are  of  opposite  signs, 
at  least  one  root  lies  between  the  numbers  sub- 
stituted. 

(6)  Solve  x3  +  m  -  2  «  0.  Here  x8  «  -  X  +  2. 
Put  y — x3  and  y  —  -  x  +  2.  Plot  the  graph  of  each 
of  these  equations,  using  a  Table  of  Cubes, 
The  abscissa  of  the  point  of  intersection  of 
these  two  curves  is  one  root  of  the  given  equa- 
tion.    x—OM  (Fig.  141)  is  the  root  required. 


356 


HIGHER  MATHEMATICS. 


§114. 


(7)  Show,  by  plotting,  that  an  equation  of  an  even  degree  with  real 
coefficients,  has  either  2,  4,  ...  or  an  even  number  of  roots,  or  else  no  roots 
at  aU. 

(8)  Plots2 


+  1 


Fig.  142. 

(11)  If  *  +  «*  = 
for  e*. 


The  curve  touches  but  does  not  cut  the  oj-axis. 
This  means  that  the  point  of  contact  of 
the  curve  with  the  aj-axis,  corresponds  to 
two  points  infinitely  close  together.  That 
is  to  say,  that  there  are  at  least  two  equal 
roots. 

(9)  Solve  ac2+y2=l ;  x2  -  ix  =  y*  -  Sy. 
Plot  the  two  curves  as  shown  in  Fig. 
142  hence  x  =  ±  OM  are  the  roots  re- 
quired. 

The  graphic  method  can  also  be  em- 
ployed for  transcendental  equations. 

(10)  If  05  +  cos  x  =  0,  we  may  locate 
the  roots  by  finding  the  point  of  inter- 
section of  the  two  curves  y  =  -  x  and 
y  =  cos  x. 

0,   plot  y  =  e*  and  y  =  -  x.     Table  IV.,  page  616, 


In  his  Die  Thermodynamik  in  der  Ghemie  (Leipzig,  61,  1893), 
J.  J.  van  Laar  tabulates  the  values  of  b  calculated  from  the  expres- 
sion 


log 


-  2 


1-82 


v,  -  V 


for  corresponding   values   of   vY   and  v2 
table : 


Here   is   part   of  the 


Vi. 

v*- 

6. 

7754 
1688 
196-5 

1-0169 
1-0432 
1-1268 

0-804 
0-775 
0-700 

Operations  like  this  are  very  tedious.  There  are  no  general 
methods  for  solving  equations  containing  logarithms,  sines, 
cosines,  etc.  There  is  nothing  for  it  but  to  educe  the  required 
value  by  successive  approximations.  Thus,  substitute  for  v2  and 
vv  so  as  to  get 

log  (196-5  -  b)  -  log  (11268  -  b)  -  2  =  f^zj.        (1) 


§  114.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS. 


357 


Now  set  up  the  following  table  containing  values  of  b  computed 
on  the  right  and  on  the  left  sides  of  equation  (1) : — 


b. 

Bight  Side. 

Left  Side. 

1-0 
0-8 
0-6 
0-4 

14-5 
6-6 
8-4 
2-6 

5-8 
4-1 
3-9 
3-6 

In  the  first  pair,  b  is  greater  on  the  right  side  than  on  the  left ; 
in  the  second  pair,  b  is  greater  on  the  left  side  than  on  the  right. 
Hence,  we  see  that  the  desired  value  of  b  lies  between  0*8  and  0*6. 
Having  thus  located  the  root,  further  progress  depends  upon  the 
patience  of  the  computer.  Closer  approximations  are  got  by  pro- 
ceeding in  the  same  way  for  values  of  v  between  0*8  and  0*6.  By 
plotting  the  assigned  values  of  b,  as  ordinates,  with  the  computed 
values  on  the  right  and  left  sides  of  (1),  as  absciss®,  it  is  possible 
to  abbreviate  the  work  very  considerably.  Very  often  the  physical 
conditions  of  the  problem  furnish  us  with  an  approximate  idea  of 
the  magnitude  of  the  desired  root. 

Examples.— (1)  M.  Planck  (Wied.  Ann.,  40,  561,  1890)  in  his  study  of  the 
potential  difference  between  two  dilute  solutions  of  binary  eleotrolytes,  de- 
veloped the  equation 


xUj  -  Ux 

V2  -  xVx 


log  k  -  log  x  xC2  -  Cx 


log  k  +  log  x   C2  -  xCy" 
By  plotting  x  as  abscissa  and  y  as  ordinate  in  the  two  equations 


~  F„  -  xK 


log  k  -  log  x  a?C2  -  Gx 
log  k  +  log  x  '  C2  -  xCx 


the  point  of  intersection  of  the  two  curves  will  be  found  to  give  the  desired 
value  of  x.  In  one  experiment  the  constants  assumed  the  following  values : 
L7!  =  5-2  ;  U2  =  272 ;  Vx  =  5-4  ;  V2  =  54 ;  Cx  =  0-1 ;  02  =  10.  It  is  required 
to  find  the  corresponding  value  of  x.  The  alternative  method  just  described 
furnishes  x  =  0-1139. 

(2)  W.  Hecht,  M.  Conrad  and  0.  Bruokner  {Zeit.phys.  Chem.,  4, 273, 1889) 
in  their  study  of  "  affinity  constants  "  solved  the  equations 

nQKO„      .  25  ,  _      25         _ .  25  ,  25 

0*3537  =  log  gg~3 


log  20"^  ;  °*3537  =  l0S  20-317  ~  l08 


25 


"  with  an  accuracy  up  to  0*01  of  the  units  employed 
y  =  8-217. 


Ansrs.  x  =  36'78 : 


^ 


%  t 


V 


368  HIGHER  MATHEMATICS.  §  115. 

§  115.  Newton's  Method  for  the  Approximate  Solution  of 
Numerical  Equations. 

Aooording  to  the  above  method,  the  equation 

/(s)-y-0»-7*  +  7,  .  .  .  (1) 
has  a  root  lying  somewhere  between  -  3  and  -  4.  We  can  keep 
on  assigning  intermediate  values  to  x  until  we  get  as  near  to  the 
exact  value  of  the  root  as  our  patience  will  allow.  Thus,  if  x  =  -  3, 
y  =»  +  1,  ifa?  =  -  3*2,  y  =  -  3*3.  The  desired  root  thus  lies  •some- 
where between  -  3  and  -  3*2.  Assume  that  the  actual  value  of 
the  root  is  -  3*  1.  To  get  a  close  approximation  to  the  root  by 
plotting  is  a  somewhat  laborious  operation.  Newton's  method 
based  on  Taylor's  theorem,  allows  the  process  to  be  shortened. 

Let  a  be  the  desired  root,  then 

/(a)  -  a*  -  la  +  7.  .  .  .  (2) 

As  a  first  approximation,  assume  that  a  ==  -  3*1  +  h,  is  the  required 
root.     From  (1),  by  differentiation, 

dy      M      7.  ffy  d*y 

All  suooeeding  derivatives  are  zero.     By  Taylor's  theorem 

dy      h2   d*y      h*   d*y 

Put  v  —  -  31  and  a  —  v  +  h. 

dv       h*    dh)      h*   d*v 

Neglecting  the  higher  powers  of  h,  in  the  first  approximation, 

a»)  +  4'-6>V>--^   •     *     (4) 

where  f'(v)  =  dv/dx.  The  value  of  f(v)  is  found  by  substituting 
-  3*1,  in  (2),  and  the  value  of  f\v)  by  substituting  -  3*1,  in  the 
first  of  equations  (3),  thus,  from  (4), 

/(„)  _  1-091  _     Q4999 

Hence  the  first  approximation  to  the  root  is  -  3*05. 
As  a  seoond  approximation,  assume  that 

a  -  -  3*05  +  \  m  vx  +  hv 
As  before, 

/K)         0-022625      +n.nmoft„ 
1_  ~  7K)  209081  ~  +  °  °01082- 


§  116.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS.  359 

The  second  approximation,  therefore,  is  -  3*048918.  We  can,  in 
this  way,  obtain  third  and  higher  degrees  of  approximation.  The 
first  approximation  usually  gives  all  that  is  required  for  practical 
work. 

Examples. — (1)  In  the  same  way  show  that  the  first  approximation  to 
one  of  the  roots  of  a-8  -  4cc2  -  2x  +  4  =»  0,  is  a  =  4-2491  . . .  and  the  second 
«  -  4-2491405. . . . 

(2)  If  x*  +  2x*  +  3x  -  60  -  0;  x  =*  2-9022834. . . . 

(8)  The  method  can  sometimes  be  advantageously  varied  as  follows. 
Solve 

(??S)'-«« <6> 

Put  x  -  1,  and  the  left  side  becomes  0-3975 — a  number  very  nearly  0-398.  If, 
therefore,  we  put  1  +  a  for  x,  a  will  be  a  very  small  magnitude. 

/0-795\i  +  «      -    .  ,„ 

•HSTV       -/(a) (6) 

By  Maclaurin's  theorem, 

/(o)  =  /(0)  +  o/'(0)  +  remaining  terms.  .        .        (7) 

As  a  first  approximation,  omit  the  remaining  terms  since  they  include  higher 
powers  of  a  small  quantity  a.  If  /(0)  =  0*8975,  by  differentiation  of  the  left 
side  of  (6),  /(0)  -  -  0*5655.    Hence, 

/(a)  -  0*3975  -  0*5655a. 
But  by  hypothesis,  /(a)  *  0-398, 

.-.  0-398  =  0-3975  -  0'6655a ;  or,  a  =  -  0-0008842. 

Since,  *  —  1  +  a,  it  follows  that  *  =  0-991158.  By  substituting  this  value  of 
SB  in  the  left  side  of  (5),  the  expression  reduces  to  0-39801  whioh  is  sufficiently 
close  to  0-398  for  all  practical  requirements.  But,  if  not,  a  more  exact  result 
will  be  furnished  by  treating  0-9991158  +  p  *•  x  exaotly  as  we  have  done 
1  +  o  =  x, 

§  116.   How  to  Separate  Equal  Roots  from  an  Equation. 

This  is  a  preliminary  operation  to  the  determination  of  the 
roots  by  a  process,  perhaps  simpler  than  the  above.  From  (5), 
page  354,  we  see  that  if  a,  /?,  y,  . . .  are  the  roots  of  an  equation 
of  the  wth  degree, 

xn  +  aaf-1  +  . . .  +  sx  +  B  -  0, 
becomes 

(X  -  a)  (X  -  /?)  .  .  .  (X  -  rj)  -  0. 
If  two  of  the  roots  are  equal,  two  factors,  say  x  -  a  and  x  -  f3f 
will  be  identical  and  the  equation  will  be  divisible  by  (x  -  a)2 ;  if 
there  are  three  equal  roots,  the  equation  will  be  divisible  by  (x  -  a)s, 
etc.     If  there  are  n  equal  roots,  the  equation  will  contain  a  factor 


360  HIGHER  MATHEMATICS.  §  117. 

(x  -  a)n,  and  the  first  derivative  will  contain  a  factor  n(x  -  a)n  ~ l, 
or  x  -  a  will  occur  n  -  1  times.  The  highest  common  faotor  of 
the  original  equation  and  its  first  derivative  must,  therefore,  contain 
x  -  a,  repeated  once  less  than  in  the  original  equation.  If  there 
is  no  common  factor,  there  are  no  equal  roots. 

Examples.— (1)  x3  -  5a:2  -  Qx  +  48  =»  0  has  a  first  derivative  3a;a-  10aj-8. 
!The  common  factor  is  x  -  4.  This  shows  that  the  equation'  has  two  roots 
equal  to  x  +  4. 

(2)  x*  +  Ix*  -  3*2  -  56a?  +  60  »  0  has  two  roots  each  equal  to  x  -  5. 

§  117.  Sturm's  Method  of  Locating  the  Real  and  Unequal 
Roots  of  a  Numerical  Equation. 

Newton's  method  of  approximation  does  not  give  satisfactory 
results  when  the  two  roots  have  nearly  equal  values.  For  instance, 
the  curve 

y  =  x3  -  lx  +  7 

has  two  nearly  equal  roots  between  1  and  2,  which  do  not  appear 

if  we  draw  the  graph  for  the  corresponding  values  of  x  and  y,  viz.: 

x  =  0,        1,        2,        3,...; 

y-7,        1,        1,       13,... 

The  problem  of  separating  the  real  roots  of  a  numerical  equa- 
tion is,  however,  completely  solved  by  what  is  known  as  Sturm's 
theorem.  It  is  clear  that  if  x  assumes  every  possible  value  in 
succession  from  +  oo  to  -  oo,  every  change  of  sign  will  indicate 
the  proximity  of  a  real  root.  The  total  number  of  roots  is  known 
from  the  degree  of  the  equation,  therefore  the  number  of  imaginary 
roots  can  be  determined  by  difference. 
Number  of  real  roots  +  Number  of  imaginary  roots  m  Total  number  of  roots. 

Sturm's  theorem  enables  these  changes  of  sign  to  be  readily 
detected.     The  process  is  as  follows  : — 

First  remove  the  real  equal  roots,  as  indicated  in  the  preceding 

section,  let 

y  =  x3  -  lx  +  7,  .        .        .        (1) 

remain.     Find  the  first  differential  coefficient, 

y  =  Bx*  -  7 (2) 

Divide  the  primitive  (1)  by  the  first  derivative  (2),  thus, 

xs  -  lx  +  7 
3a;2  -  7    ' 
and  we  get  %x  with  the  remainder  -  i(14#  -  21).     Change  the 


§  117.       HOW  TO  SOLVE  NUMERICAL  EQUATIONS.  361 

sign  of  the  remainder  and  multiply  by  £,  the  result 

B  =  2x  -  3,      .         .         .         .         (3) 

is  now  to  be  divided  into  (2).     Change  the  sign  of  the  remainder 
and  we  obtain, 

-B-1 (4) 

The  right-hand  sides  of  equations  (1),  (2),  (3),  (4), 

&  -  7a; +  7;  3a;2-  7;  2a;-  3;  1, 

are  known  as  Sturm's  functions. 

Substitute  -  oo  for  x  in  (1),  the  sign  is  negative ; 

(2),  „        positive ; 

(3),  „        negative ; 

(4),  „         positive. 

Note  that  the  last  result  is  independent  of  x.     The  changes  of 

sign  may,  therefore,  be  written 

-  +  -  +. 
In  the  same  way, 


It 

tt 

ft 

tt 

tt 

n 

Value  of  x. 

Corresponding  Signs 

Number  of  Changes 

of  Sturm's  Functions. 

of  Sign. 

—  00 

-  +  -  + 

■      3 

•    -  4 

-  +  -  + 

3 

-  3 

+  +  -  + 

2 

-  2 

+  +  -  + 

2 

-  1 

+ + 

2 

+  0 

+ + 

2 

+  1 

+ + 

2 

+  2 

+  +  +  + 

0 

+  oo 

+  +  +  + 

0 

There  is,  therefore,  no  change  of  sign  caused  by  the  substitution 
of  any  value  of  x  less  than  -  4,  or  greater  than  +  2 ;  on  passing 
from  -  4  to  -  3,  there  is  one  change  of  sign  ;  on  passing  from 
1  to  2,  there  are  two  changes  of  sign.  The  equation  has,  there- 
fore, one  real  root  between  -  4  and  -  3,  and  two  between  1  and  2. 

It  now  remains  to  determine  a  sufficient  number  of  digits,  to 
distinguish  between  the  two  roots  lying  between  1  and  2.  First 
reduce  the  value  of  x  in  the  given  equation  by  1.  This  is  done  by 
substituting  u  + 1  in  place  of  x,  and  then  finding  Sturm's  functions 
for  the  resulting  equation.     These  are, 

u?  +  du2  -  4w  +  1 ;  3w2  +  6u  -  4 ;  8m— - 1;  1. 


362  HIGHER  MATHEMATICS. 

As  above,  noting  that  if  x  =  +  1*1,  u  =  +  0*1,  etc., 


§117. 


Value  of  x. 

Corresponding  Signs 
of  Sturm's  Functions. 

Number  of  Changes 
of  Sign. 

1-1 

1-2 
1-3 
1-4 
1-5 
1*6 
1-7 

+111 +++ 
+  +  1  1  1  1  1 
+  +  +  1  II  1 
+++++++ 

2 
2 

2 
1 
1 
1 
0 

The  second  digits  of  the  roots  between  1  and  2  are,  therefore, 
3  and  6,  and  three  real  roots  of  the  given  equation  are  approxi- 
mately -  3,  1-3,  1*6. 

Examples. — Locate  the  roots  in  the  following  equations : 

(1)  a*3- 3a?2 -4a; +13.    Ansr.  Between  -3  and  -2;  2  and  2-5;  2-5  and  3 

(2)  xz  -  4a>a  -  6a;  +  8.     Ansr.  Between  0  and  1 ;  5  and  6 ;  -  1  and  -  2. 

(3)  a*  +  a*3  -  x2  -  2x  +  4.  We  have  five  Sturm's  functions  for  this  equa- 
tion. Call  the  original  equation  (1),  the  first  derivative,  4a;3  +  3a;2  -  2a;  -  2,  (2) ; 
divide  (1)  by  (2)  and  x1  +  2a;  -  6  (3)  remains ;  divide  (2)  by  (3)  and  -  x  +  1  (4) 
remains  ;  divide  (3)  by  (4)  and  change  the  sign  of  the  result  for  +  1  (5).  Now 
let  x  a«  +  oo  and  -  oo,  we  get 

+  +  +  -  +  (2  variations  of  sign) ;  +  -  +  +  +  (2  variations). 
This  means  that  there  are  no  real  roots.     All  the  roots  are  imaginary. 

(4)  Calculate  the  volume,  v,  of  one  gram  of  carbon  dioxide  at  0°C.  and 
one  megadyne  pressure  per  sq.  cm.,  given  van  der  Waals'  equation 


(p  +  ^j}T^)  (*  -  0*9565)  -  1-8824T. 


0°C.  =  273,7°27;  p  =  1.  Expand  the  equation  and  arrange  terms  in  descend- 
ing powers  of  v.  Substitute  the  numerical  values  of  the  constants  and  reduce 
to 

v3  -  516-17«8  +  1852-6v  -  1772-0  =  0. 

The  only  admissible  root  of  this  cubic  is  512*5.  The  labour  of  solving  this 
equation  can  sometimes  be  reduced  by  neglecting  a/v2  when  it  is  small. 

(5)  The  equation,  xs  -  Srx2  +  4r*p  =  0,  is  obtained  in  problems  referring 
to  the  depth  to  which  a  floating  sphere  of  radius  r  and  density  p  sinks  in 
water.  Solve  this  equation  for  the  case  of  a  wooden  ball  of  unit  radius  and 
specific  gravity  0*65.  Hence,  Xs  -  3a;2  +  2*6  =  0.  The  three  roots,  by  Sturm's 
theorem,  are — a  negative  root,  a  positive  root  between  1  and  2,  and  one  over  2. 
The  depth  of  the  sphere  in  the  water  cannot  be  greater  than  its  diameter  2. 
A  negative  root  does  not  represent  a  physical  reality.  The  two  negative  roots 
must,  therefore,  be  excluded  from  the  solution.  The  other  root,  by  Newton's 
method  of  approximation,  is  z  =  1*204. . . . 


§  118.       HOW  TO  SOLVE  NUMEEICAL  EQUATIONS.  363 

In  this  last  example  we  have  rejected  two  roots  because  they 
were  inconsistent  with  the  physical  conditions  of  the  problem  under 
consideration.  This  is  a  very  common  thing  to  do.  Not  all  the 
solutions  to  which  an  equation  may  lead  are  solutions  of  the  prob- 
lem. Of  course,  every  solution  has  some  meaning,  but  this  may 
be  quite  outside  the  requirements  of  the  problem.  A  mathematical 
equation  often  expresses  more  than  Nature  allows.  In  the  physical 
world  only  changes  of  a  certain  kind  take  place.  If  the  velocity 
of  a  falling  body  is  represented  by  the  expression  v*  —  64s,  then, 
if  we  want  to  calculate  the  velocity  when  s  is  4,  we  get  v2  =  256, 
or,  v  —  ±  16.  In  other  words,  the  velocity  is  either  positive  or 
negative.  We  must  therefore  limit  the  generality  of  the  mathe- 
matical statement  by  rejecting  those  changes  which  are  physically 
inadmissible.  Thus  we  may  have  to  reject  imaginary  roots  when 
the  problem  requires  real  numbers ;  and  negative  or  fractional 
roots,  when  the  problem  requires  positive  or  whole  numbers. 
Sometimes,  indeed,  none  of  the  solutions  will  satisfy  the  condi- 
tions imposed  by  the  problem,  in  this  case  the  problem  is  inde- 
terminate. The  restrictions  which  may  be  imposed  by  the 
application  of  mathematical  equations  to  specific  problems,  intro- 
duces us  to  the  idea  of  limiting  conditions,  which  is  of  great 
importance  in  higher  mathematics.  The  ultimate  test  of  every 
solution  is  that  it  shall  satisfy  the  equation  when  substituted  in 
place  of  the  variable.     If  not  it  is  no  solution. 

Examples. — (1)  A  is  40  years,  B  20  years  old.  In  how  many  years  will 
A  be  three  times  as  old  as  B ?    Let  x  denote  the  required  number  of  years. 

.-.  40  +  x  »  3(20  +  x) ;  or  x  =  -  10.     • 
But  the  problem  requires  a  positive  number.    The  answer,  therefore,  is  that 
A  will  never  be  three  times  as  old  as  B.     (The  negative  sign  means  that  A 
was  three  times  as  old  as  B,  10  years  ago.) 

(2)  A  number  x  is  squared;  subtract  7;  extract  the  square  root  of  the 
result;  add  twice  the  number,  6  remains.     What  was  the  number  x? 

.-.  2x  +  J{x2  -  7)  =  5.       

Solve  in  the  usual  way,  namely,  square  5  -  2x  m  s/x2  -  7 ;  rearrange  terms 
and  use  (2),  §  113.  Hence  x  =  4  or  f .  On  trial  both  solutions,  x  —  4  and 
x  =  2|,  fail  to  satisfy  the  test.  These  extraneous  solutions  have  been  intro- 
duced during  rationalization  (by  squaring). 

§  118.   Horner's  Method  for  Approximating  to  the  Real 
Roots  of  Numerical  Equations. 

When  the  first  significant  digit  or  digits  of  a  root  have  been 
obtained,   by,  say,   Sturm's  theorem,    so  that  one   root  may  be 


364  HIGHER  MATHEMATICS.  §  118. 

distinguished  from  all  the  other  roots  nearly  equal  to  it,  Horner's 
method  is  one  of  the  simplest  and  best  ways  of  carrying  the 
approximation  as  far  as  may  be  necessary.  So  far  as  practical 
requirements  are  concerned,  Horner's  process  is  perfection.  The 
arithmetical  methods  for  the  extraction  of  square  and  cube  roots 
are  special  cases  of  Horner's  method,  because  to  extract  i/9,  or 
v^9,  is  equivalent  to  finding  the  roots  of  the  equation  x2  -  9  =  0, 
or  a8  -  9  =  0. 

••  Considering  the  remarkable  elegance,  generality,  and  simplicity  of  the 
method,  it  is  not  a  little  surprising  that  it  has  not  taken  a  more  prominent  place 
in  current  mathematical  text-books.  Although  it  has  been  well  expounded 
by  several  English  writers,  ...  it  has  scarcely  as  yet  found  a  place  in  English 
curricula.  Out  of  five  standard  Continental  text-books  where  one  would  have 
expected  to  find  it  we  found  it  mentioned  in  only  one,  and  there  it  was  ex- 
pounded in  a  way  which  showed  little  insight  into  its  true  character.  This 
probably  arises  from  the  mistaken  notion  that  there  is  in  the  method  some 
algebraic  profundity.  As  a  matter  of  fact,  its  spirit  is  purely  arithmetical ; 
and  its  beauty,  whioh  can  only  be  appreciated  after  one  has  used  it  in 
particular  cases,  is  of  that  indescribably  simple  kind  which  distinguishes 
the  use  of  position  in  the  decimal  notation  and  the  arrangement  of  the  simple 
rules  of  arithmetic.  It  is,  in  short,  one  of  those  things  whose  invention  was 
the  creation  of  a  commonplace." — Q.  Chrystal,  Text-book  of  Algebra  (London, 
i.,  346, 1898). 

In  outline,  the  method  is  as  follows  :  Find  by  means  of  Sturm's 
theorem,  or  otherwise,  the  integral  part  of  a  root,  and  transform 
the  equation  into  another  whose  roots  are  less  than  those  of  the 
original  equation  by  the  number  so  found.  Suppose  we  start 
with  the  equation 

x*  -  7x  +  7  =  0,  .        .        .        .        (1) 

which  has  one  real  root  whose  first  significant  figures  we  have 
found  to  be  1*3.  Transform  the  equation  into  another  whose 
roots  are  less  by  1*3  than  the  roots  of  (1).  This  is  done  by 
substituting  u  +  IB  for  x.     In  this  way  we  obtain, 

u  +  3-90w2  -  l'93w3  +  -097  =  0.  .  .  (2) 
The  first  significant  figure  of  the  root  of  this  equation  is  0*05.  Lower 
the  roots  of  (2)  by  the  substitution  of  v  +  0*05  for  u  in  (2).  Thus, 
V2  +  4.05^2  _  i-5325t>  +  -010375  =  0. .  .  (3) 
The  next  significant  figure  of  the  root,  deduced  from  (3),  is  "006. 
We  could  have  continued  in  this  way  until  the  root  had  been 
obtained  of  any  desired  degree  of  accuracy. 

Practically,  the  work  is  not  so  tedious  as  just  outlined.  Let  a,  b,  c, 


§  118.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS.  365 

be  the  coefficients  of  the  given  equation,  B  the  absolute  term, 
ax3  +  bx 2  +  ex  +  B  =  0. 

1.  Multiply  a  by  the  first  significant  digits  of  the  root  and  add 
the  product  to  b.     Write  the  result  under  b. 

2.  Multiply  this  sum  by  the  first  figure  of  the  root,  add  the 
product  to  c.     Write  the  result  under  c. 

3.  Multiply  this  sum  by  the  first  figure  of  the  root,  add  the 
product  to  B,  and  call  the  result  the  first  dividend. 

4.  Again  multiply  a  by  the  root,  add  the  product  to  the  last 
number  under  b. 

5.  Multiply  this  sum  by  the  root  and  add  the  product  to  the 
last  number  under  c,  call  the  result  the  first  trial  divisor. 

6.  Multiply  a  by  the  root  once  more,  and  add  the  product  to  the 
last  number  under  b. 

7.  Divide  the  first  dividend  by  the  first  trial  divisor,  and  the 
first  significant  figure  in  the  quotient  will  be  the  second  significant 
of  the  root.  Thus  starting  from  the  old  equation  (1),  whose  root 
we  know  to  be  about  1. 

a  b  c  R  (Boot 

1  +0  -7  +7  (1-3 

1  1  -6 

1-6  1  First  dividend. 

1  2 

2  -  i  First  trial  divisor. 
1 

~T 

8.  Proceed  exactly  as  before  for  the  second  trial  divisor,  using 
the  second  digit  of  the  root,  vie.,  3. 

9.  Proceed  as  before  for  the  second  dividend.  We  finally  ob- 
tain the  result  shown  in  the  next  scheme.  Note  that  the  black 
figures  in  the  preceding  scheme  are  the  coefficients  of  the  second  of 
the  equations  reduced  on  the  supposition  that  x  •*  1*3  is  a  root  of 
the  equation. 

a'  V  &  R  [Boot 

18  -  4  1  (1-35 

08  0-99  -  0-908 


3-8  -  801  0*097  Second  dividend. 

0-8  1-08 


8-6  -  1»93  Second  trial  divisor. 

0-8 

3*9 


366 


HIGHER  MATHEMATICS. 


§118. 


Once  more  repeating  the  whole  operation,  we  get, 


b" 
3-9 
005 

3-95 
005 

4-00 
0-05 

4*05 


c" 

-  1-93 
0-1975 

-  1-7325 
0-2000 


R" 
0*097 
0-086625 


(Root 
(1-356 


0*010375  Third  dividend. 


1*8325  Third  trial  divisor. 


Having  found  about  five  or  seven  decimal  places  of  the  root  in 
this  way,  several  more  may  be  added  by  dividing,  say  the  fifth 
trial  dividend  by  the  fifth  trial  divisor.  Thus,  we  pass  from 
1-356895,  to  1*356895867 ...  a  degree  of  accuracy  more  than 
sufficient  for  any  practical  purpose. 

Knowing  one  root,  we  can  divide  out  the  factor  x  - 1*3569  from 
equation  (1),  and  solve  the  remainder  like  an  ordinary  quadratic. 

If  any  root  is  finite,  the  dividend  becomes  zero,  as  in  one  of 
the  following  examples.  If  the  trial  divisor  gives  a  result  too  large 
to  be  subtracted  from  the  preceding  dividend,  try  a  smaller  digit. 

To  get  the  other  root  whose  significant  digits  are  1*6,  proceed 

as  above,  using  6  instead  of  3  as  the  quotient  from  the  first  dividend 

and  trial  divisor.     Thus  we  get  1*692  . . .  Several  ingenious  short 

cuts  have  been  devised  for  lessening  the  labour  in  the  application 

of  Horner's  method,  but  nothing  much  is  gained,  when  the  method 

has  only  to  be  used  occasionally,  beyond  increasing  the  probability 

of  error.     It  is  usual  to  write  down  the  successive  steps  as  indicated 

in  the  following  example. 

Examples. — (1)  Find  the  root  between  6  and  7  in 
4*3  _  13a;2  _  31a.  _  275. 

4  -  18  -  31  -  275  (6*26 


24 


-  31 
66 


210 


11 

24 

85 
210 

-  65 

51-392 

85 
24 

245 
11-96 

18*608 

13-608 

89 

0*8 

256-96 
12-12 

0 

59-8 
0-8 

269*08 

3*08 

60-6 
0-8 

272-16 

61*4 


§  119.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS. 


367 


The  steps  mark  the  end  of  eaoh  transformation.  The  digits  in  black 
letters  are  the  coefficients  of  the  successive  equations. 

(2)  There  is  a  positive  root  between  4  and  5ina?  +  a?  +  z-  100.  Ansr. 
4-2644 . . . 

(8)  Find  the  positive  and  negative  roots  in  a*  +  8<ca  +  16a;  =  440.  Ansr. 
+  3-976 . . .,  -  4-3504.  To  find  the  negative  roots,  proceed  as  before,  but  first 
transform  the  equation  into  one  with  an  opposite  sign  by  changing  the  sign 
of  the  absolute  term. 

(4)  Show  that  the  root  between  -  3  and  -  4,  in  equation  (1),  is 
-3*0489173396  .  .  .  Work  from  a  =  1,  b  m  -0,  c  =  -7,  B  =  -7. 


§  119.    Yan  der  Waals'  Equation. 

The  relations  between  the  roots  of  equations,  discussed  in  this 
chapter,  are  interesting  in  many  ways  ;  for  the  sake  of  illustration, 
let  us  take  the  van  der  Waals'  relation  between  the  pressure,  p, 
volume,  v,  and  temperature,  T,  of  a  gas. 

(p  +  £)  («  -  b)  =  BT;  or,  #-(*  +  YJv^+^v-  |=0.    (1) 

This  equation  of  the  third  degree  in  v,  must  have  three  roots, 
a»  ft  y,  equal  or  unequal,  real  or  imaginary.     In  any  case, 

(v  -  a)  (v  -  0)  (v  -  y)  =  0.  .         .         (2) 

Imaginary  roots  have  no  physical  meaning  ;  we  may  therefore 
confine  our  attention  to  the  real  roots.  Of  these,  we  have  seen 
that  there  must  be  one,  and  there  may  be  three.  This  means  that 
there  may  be  one  or  three  (different)  volumes,  corresponding  with 
every  value  of  the  pressure,  p,  and  temperature,  T.  There  are 
three  interesting  cases  : 

I.  There  is  only  one  real  root  present.  This  implies  that  there 
is  one  definite  volume,  t>,  corres- 
ponding to  every  assigned  value  of 
pressure,  pt  and  temperature,  T. 
This  is  realized  in  the  _pt>-curve,  of 
all  gases  under  certain  physical 
conditions  ;  for  instance,  the  graph 
of  carbon  dioxide  at  91°  has  only 
one  value  of  p  corresponding  with 
each  value  of  v.  See  curve  GH, 
Fig.  143. 

II.  There  are  three  real  imequal 
roots  present.  The  ^-ourve  of 
carbon  dioxide  at  temperatures  be- 


368  HIGHER  MATHEMATICS.  §  119. 

low  32°,  has  a  wavy  curve  BC  (Fig.  143).  This  means  that  at  this 
temperature  and  a  pressure  of  Op,  carbon  dioxide  ought  to  have 
three  different  volumes  corresponding  respectively  with  the  abscissae 
Oc,  Ob,  Oa.  Only  two  of  these  three  volumes  have  yet  been 
observed,  namely  for  gaseous  C02  at  a  and  for  liquid  C02  at  y, 
the  third,  corresponding  to  the  point  /?,  is  unknown.  The  curve 
AyftaD,  has  been  realized  experimentally.  The  abscissa  of  the 
point  a  represents  the  volume  of  a  given  mass  of  gaseous  carbon 
dioxide,  the  abscissa  of  the  point  y  represents  the  volume  occupied 
by  the  same  mass  of  liquid  carbon  dioxide  at  the  same  pressure. 

Under  special  conditions,  parts  of  the  sinuous  curve  yBfiCa 
have  been  realized  experimentally.  Ay  has  been  carried  a  little 
below  the  line  ya,  and  Da  has  been  extended  a  little  above  the  line 
ya.  This  means  that  a  liquid  may  exist  at  a  pressure  less  than 
that  of  its  own  vapour,  and  a  vapour  may  exist  at  a  pressure  higher 
than  the  "  vapour  pressure  "  of  its  own  liquid. 

III.  There  are  three  real  equal  roots  present.  At  and  above 
the  point  where  a  =  /?  =  y,  there  can  only  be  one  value  of  v  for  any 
assigned  value  of  p.  This  point  K  (Fig.  143)  is  no  other  than  the 
well-known  critical  point  of  a  gas.  Write  pe,  ve,  Tci  for  the  critical 
pressure,  volume,  and  temperature  of  a  gas.     From  (2), 

(v  -  af  =  0;  or,  v  -a;  .  .  .  (3) 
let  ve  denote  the  value  of  v  at  the  critical  point  when  a  =  v  »  vc. 
Therefore,  if  pe  denotes  the  pressure  corresponding  wiih  v  =  ve, 
from  (1),  and  the  expansion  of  (3), 

^  -  \b  +  ~7rr  +  Htv  ~  ^r  *  ^  -  3^2  +  Bv2ov  -  **    (4) 

\  Pe  /  Pc  P« 

This  equation  is  an  identity,  therefore,  from  page  213, 

dvepe  -  bpe  +  BTe ;  3v*ep0  =-  a ;  v*epc  =  ab,  .  (5) 
are  obtained  by  equating  the  coefficients  of  like  powers  of  the 
unknown  v.     From  the  last  two  of  equations  (5), 

ve  =  Bb.         .        .        .        .        (6) 
From  (6)  and  the  second  of  equations  (5), 

Pc  =  27 '  P*  *        *        *        W 

From  (6),  (7),  and  the  first  of  equations  (5), 

e      27   bB'      '        *        •       •'        W 


§  119.        HOW  TO  SOLVE  NUMERICAL  EQUATIONS.  369 

From  these  results,  (6),  (7),  (8),  van  der  Waals  has  calculated  the 
values  of  the  constants  a  and  b  for  different  gases.  Let  p  —  p/pc, 
v  -  v/v„  T  -  T/Tc.     From  (1),  (6),  (7)  and  (8),  we  obtain 

(p+|)(3v-1)-8T,    ...        (9) 

which  appears  to  be  van  der  Waals'  equation  freed  from  arbitrary 
constants.  This  result  has  led  van  der  Waals  to  the  belief  that 
all  substances  can  exist  in  states  or  conditions  where  the  corre- 
sponding pressures,  volumes  and  temperatures  are  equivalent. 
These  he  calls  corresponding  states — uebereinstimmende  Zustande. 
The  deduction  has  only  been  verified  in  the  case  of  ether,  sulphur 
dioxide  and  some  of  the  benzene  halides. 


CHAPTEE  VII. 

HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS. 

•*  Theory  always  tends  to  become  more  abstract  as  it  emerges  success- 
fully from  the  chaos  of  facts  by  processes  of  differentiation  and 
elimination,  whereby  the  essentials  and  their  connections  be- 
come recognized,  while  minor  effects  are  seen  to  be  secondary 
or  unessential,  and  are  ignored  temporarily,  to  be  explained  by 
additional  means." — O.  Heavisidb. 

§  120.    The  Solution  of  a  Differential  Equation  by  the 
Separation  of  the  Variables. 

This  chapter  may  be  looked  upon  as  a  sequel  to  that  on  the 
integral  calculus,  but  of  a  more  advanced  character.  The 
"  methods  of  integration  "  already  described  will  be  found  ample 
for  most  physico-chemical  processes,  but  more  powerful  methods 
are  now  frequently  required. 

I  have  previously  pointed  out  that  in  the  effort  to  find  the 
relations  between  phenomena,  the  attempt  is  made  to  prove  that 
if  a  limited  number  of  hypotheses  are  prevised,  the  observed  facts 
are  a  necessary  consequenoe  of  these  assumptions.  The  modus 
operandi  is  as  follows: — 

1.  To  "anticipate  Nature"  by  means  of  a  "  working  hypoth- 
esis," which  is  possibly  nothing  more  than  a  "convenient  fiction ". 

"From  the  practical  point  of  view,"  said  A.  W.  Biicker  (Presidential 
Address  to  the  B.  A.  meeting  at  Glasgow,  September,  1901),  "it  is  a  matter  of 
secondary  importance  whether  our  theories  and  assumptions  are  correct,  if 
only  they  guide  us  to  results  in  accord  with  facts.  ...  By  their  aid  we  can 
foresee  the  results  of  combinations  of  causes  which  would  otherwise  elude  us." 

2.  Thence  to  deduoe  an  equation  representing  the  momentary 
rate  of  change  of  the  two  variables  under  investigation. 

3.  Then  to  integrate  the  equation  so  obtained  in  order  to 
reproduce  the  "  working  hypothesis "  in  a  mathematical  form 
suitable  for  experimental  verification. 

370 


§  120.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.         371 

So  far  as  we  are  concerned  this  is  the  ultimate  object  of  our 
integration.  By  the  process  of  integration  we  are  said  to  solve 
the  equation.  For  the  sake  of  convenience,  any  equation  contain- 
ing differentials  or  differential  coefficients  will,  after  this,  be  called 
a  differential  equation. 

I. — The  variables  can  be  separated  directly. 

The  different  equations  hitherto  considered  have  required  but 
little  preliminary  arrangement  before  integration.  For  example, 
the  equations  representing  the  velocity  of  chemical  reactions  have 
the  general  type : 

■£-*/(«).    .    .     .     .    a) 

We  have  invariably  collected  all  the  x's  on  one  side,  the  V s,  on 
the  other,  before  proceeding  to  the  integration.  This  separation  of 
the  variables  is  nearly  always  attempted  before  resorting  to  other 
artifices  for  the  solution  of  the  differential  equation,  because  the 
integration  is  then  comparatively  simple. 

Examples. — (1)  Integrate  the  equation,  y .  dx  +  x  .  dy  m  0.  Rearrange 
the  terms  so  that 

by  multiplying  through  with  1/xy.  Ansr.  log  x  +  log  y  —  C.  Two  or  more 
apparently  different  answers  may  mean  the  same  thing.  Thus,  the  solution 
of  the  preceding  equation  may  also  be  written,  log  By  —  loge0;  i.e.,  xy  =■  ea\ 
or  log  xy  =  log  C ;  i.e.,  xy  =  C.  C  and  log  C  are,  of  course,  the  arbitrary 
constants  of  integration. 

(2)  F.  A.  H.  Schreinemaker  (Zeit.  phys.  Chem.,  36,  413,  1901)  in  his 
study  of  the  distillation  of  ternary  mixtures,  employed  the  equation  dy/dx  = 
ay/x.  Hence  show  that  y  =  Cx*.  He  calls  the  graph  of  this  equation  the 
"  distillation  curve  ". 

(8)  The  equation  for  the  rectilinear  motion  of  a  partiole  under  the  in- 
fluence of  an  attractive  force  from  a  fixed  point  if  v  .  dv/dx  +  ax~3  =»  0  ;  .\ 
Jv2  -  a\x  +  C. 

(4)  In  consequence  of  imperfeot  insulation,  the  charge  on  an  electrified 
body  is  dissipated  at  a  rate  proportional  to  the  magnitude  E  of  the  charge. 
Hence  show  that  if  a  is  a  constant  depending  on  the  nature  of  the  body,  and 
E0  represents  the  magnitude  of  the  charge  when  t  (time)  =  0,  E  =  2E0e-«*. 
Hint.  Compound  interest  law.  Integrate  by  the  separation  of  the  variables. 
Interpret  your  result  in  words. 

(5)  Solve  (1  +  x*)dy  =  >Jy  .  dx.    Ansr.  2  Jy  -  tan  -  lx  =  C. 

(6)  Solve  y  -  x  .  dy/dx  =  a(y  +  dy/dx).     Ansr.  y  «=  C(a  +  x)  (1-  a). 

(7)  Abegg's  formula  for  the  relation  between  the  dielectric  oonstant,  JD,  of 

AA* 


372  HIGHER  MATHEMATICS.  §  T20. 

_        Q 

a  fluid  and  temperature  9,  is  -  dD/dO  =  y^D.  Hence  show  that  D  =  Ce  ttrf, 
where  0  is  a  constant  whose  value  is  to  be  determined  from  the  conditions  of 
the  experiment.     Put  the  answer  into  words. 

(8)  What  curves  have  a  slope  -  y\x  to  the  #-axis  ?  Ansr.  The  rectangular 
hyperbolas  xy  =  C.     Hint.  Set  up  the  proper  differential  equation  and  solve. 

(9)  The  relation  between  small  changes  of  pressure  and  volume  of  a  gas 
under  adiabatic  conditions,  is  ypdv  +  vdp—O.    Hence  show  that  pv~l  =  constant. 

(10)  A  lecturer  discussing  the  physical  properties  of  substances  at  very  low 
temperatures,  remarked  "  it  appears  that  the  specific  heat,  o-,  of  a  substance 
decreases  with  decreasing  temperatures,  6,  at  a  rate  proportional  to  the 
speoific  heat  of  the  substance  itself".  Set  up  the  differential  equation  to 
represent  this  "law"  and  put  your  result  in  a  form  suitable  for  experimental 
verification.     Ansr.  (log«r0  -  logo-)/0  =  const. 

(11)  Helmholtz's  equation  for  the  strength  of  an  electric  current,  C,  at  the 
time  t%  is  C  =  E/B  -  (L/B)dC/dt,  where  E  represents  the  electromotive  force 
in  a  circuit  of  resistance  B  and  self-induction  L.  If  E,  B,  L,  are  constants, 
show  that  BC  -  E(l  -  «-■»/*)  provided  0  =  0,  when  t  =  0. 

(12)  The  distance  x  from  the  axis  of  a  thick  cylindrical  tube  of  metal  is  re- 
lated to  the  internal  pressure  _p  as  indicated  in  the  equation  (2p  -  a)dx  +  xdp=0, 
where  a  is  a  constant.     Hence  show  th&tp  =  $a  +  Cx-2. 

(18)  A  substitution  will  often  enable  an  equation  to  be  treated  by  this 
simple  method  of  solution.  Solve  (x  -  y2)dx  +  2xydy  =  0.  Ansr.  xev<ilx  =  O. 
Hint.  Put  2/2  =  v,  divide  by  a;2,  .-.  dxjx  +  d(y/x)  =  0,  etc. 

(14)  Solve  Dulong  and  Petit's  equation:  dd/dt  =  b(c9  -  1),  page  60.  Put 
cP  -  1  =  x  and  differentiate  for  dd  and  dx.  Hence  dx  =  c#  log  c  .  d0, 
dd  =  dx\c$  log  c ;  and  page  213,  Case  1. 

f    dd  f  dx  ..,  ,    ■      x  _, 

]*rzn-=]x(x  +  l)\ogc>  .'.  Wlogc-logj^ri  +  C;  etc. 

(15)  Solve  Stefan's  equation  :  dd/dt  =  a{(273  +  a)4  -  2734},  page  60.  Put 
x  =  273  +  9  and  c  =  273.  Hence  the  given  equation  can  be  written  dxjdt 
=  a(x4  -  c4)  =  a(x  +  c)  (x  -  c)  (<c2  +  c2)  which  can  be  solved  by  Case  3,  page 

216.     Thus,  at  =  -L  (log  lii  -  2  tan-*  -  )  +  c. 
4c3  I    6x  +  c  cJ 

(16)  Solve  du/dr  -  %i\r  —  GA\r2  -  \ar2.  Substitute  v  =  ujr\  .*.  rdv/dr  = 
dujdr  -  uff.     .-.  dvjdr  =  G^r*  -  Jor;  .-.  ujr  =  C2  -  ^Cj/r2  -  i«r2. 

(17)  According  to  the  Glasgow  Herald  the  speed  of  H.M.S.  Sapphire  was 
V  when  the  engines  indicated  the  horse-power  P.     When 

P  =  5012,  7281,  10200,  12650 ; 

V  =  18-47,  20-60,  22-43,  23-63. 

Do  these  numbers  agree  with  the  law  dPjdV—aP,  where  a  is  constant? 
Ansr.  Yes.  Hint.  On  integration,  remembering  that "V  =  0  when  P  =  0,  we 
get  log10P  -  log10O  =  «7,  where  G  is  constant.  Evaluate  the  constants  as 
indicated  on  page  324,  we  get  C  =  181,  a  =  0-07795,  etc. 

II. — The  equation  is  homogeneous  in  x  and  y. 

If  the  equation  be  homogeneous  in  x  and  y,  that  is  to  say,  if 
the  sum  of  the  exponents  of  the  variables  in  each  term  is  of  the 


§  120.      HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.      373 

same  degree,  a  preliminary  substitution  of  x  =  ty,  or  y  =  tx,  ac- 
cording to  convenience,  will  always  enable  variables  to  be  separated. 
The  rule  for  the  substitution  is  to  treat  the  differential  which  in- 
volves the  smallest  number  of  terms. 

Examples. — (1)  Solve  x  +  y  .  dyjdx  -  2y  m  0.  Substitute  y  =  ex ;  or 
dy  =  xdz  +  zdxt  and  rearrange  terms.  We  get  (1  -  2z  +  z%)dx  +  xzdz  =  0 ;  or 
(1  -  zfdx  +  xzdz  =  0 ;  and 

\-0zy  +  \%  =  G;  •'•  f^i +  log(1  ~ z)  +  logx  =  °' ' {X  "  &*!?  =  C' 

(2)  F.  A.  H.  Schreinemaker  (Zeit.  phys.  Ghem.,  36,  413,  1901)  in  studying 
the  vapour  pressure  of  ternary  mixtures  used  the  equation  dyjdx  =  myfx  +  n 
This  becomes  homogeneous  when  x  =  ty  is  substituted.  Hence  show  that 
Cxm  -  nx/(m  -  1)  =  y,  where  G  is  the  integration  constant. 

(3)  Show  that  if  (y  -  x)dy  +  ydx  =  0 ;  y  =  Ce  -*t» . 

(4)  Show  that  if  x*dy  -  yldx  -  xydx  =  0;  x  =  e  -  »*   +  G. 

(5)  Show  that  if  {x*  +  y2)dx  =  2xydy ;  z2  -  jf  =  Cx. 

III. — The  equation  is  non-homogeneous  in  x  and  y. 

Non-homogeneous  equations  in  x  and  y  can  be  converted  into 
the  homogeneous  form  by  a  suitable  substitution.  The  most 
general  type  of  a  non-homogeneous  equation  of  the  first  degree  is, 

(ax  +  by  +  c)dx  +  (a'x  +  b'y  +  c')dy  =  0,        .         (2) 

where  x  and  y  are  of  the  first  degree.  To  convert  this  into  an 
homogeneous  equation,  assume  that  x  =  v  +  h ;  and  y  —  w  +  k, 
and  substitute  in  the^iven  equation  (2).     Thus,  we  obtain 

{av  +  bw  +  (ah  +  bk  +  c)}dv  +  {a'v  +  b'w  +  (a'h  +  b'k  +  c') }dw  =  0.  (3) 
Find  h  and  k  so  that  ah  +  bk  +  c  =  0 ;  a'h  +  b'k  +  c'  =  0. 

'\*-"  Hb^~ab' '  a'b  -  ab"      '        *        (** 

Substitute  these  values  of  fe  and  &  in  (3).     The  resulting  equation 
(av  +  6w)aH;  +  (a'v  +  b'w)dw  =  0,  .         (5) 

is  homogeneous  and,  therefore,  may  be  solved  as  just  indicated. 

Examples.— (1)  Solve  (Sy  -  Ix  -  l)dx  +  (7y  -  3x  -  S)dy  —  0-  Ansr. 
(y  -  x  -  1)%  +  x  +  l)5  =  C.  Hints.  From  (2),  a=-7,  6*3,  c  «=  -  7  : 
a'  =  -  3,  6'  =  7,  c'  =  -  3.  From  (4),  fc  =  -  1,  &  =  0.  Hence,  from  (3),  we 
get  Stock)  -  Ivdv  +  Iwdw  -  Svdiv  =  0.  To  solve  this  homogeneous  equation, 
substitute  w  =  vt,  as  above,  and  separate  the  variables. 

„dv       3  -  It ,.  _  fdv       f  2dt        f  5dt         _ 


374  HIGHER  MATHEMATICS.  §  121. 

.-.  7  log  v  +  21og(*  -  1)  +  51og(*  +  1)  =  C ;  or,  v'(t  -  1)2(*  +  l)5  =  C. 
But  x  *=  v  +  h,  .'.  v  =  x  +  1 ;  y  =  w  +  k,  .\  y  =  w  ;  .'.  t  =  w\v  =  yj(x  +  1),  etc. 
(2)  If  {2y  -  x  -  l)dy  +  (2x  -  y  +  l)dx  =  0;  x2  -  xy  +  y*  +  x  -  y  =  C. 

IV. — Non-homogeneous  equations  in  which  the  constants  have  the 
special  relation  ab'  =  a'b. 

If  a  :  b  =  a'  :  b'  =  1  :  m  (say),  then  h  and  k  are  indeterminate, 
since  (2)  then  becomes 

(ax  +  by  +  c)dx  +  {m(ax  +  by)  +  c'}dy  =  0. 

The  denominators  in  equations  (4)  also  vanish.     In  this  case  put 
z  =  ax  +  by,  and  eliminate  y}  thus,  we  obtain, 

,    z  +  c        dz  /ftv 

a  _  J .  _  -_    -  o,   .         .         .         (b) 

„mz  +c       ax 

an  equation  which  allows  the  variables  to  be  separated. 
Examples.— (1)  Solve  (2x  +  Sy  -  b)dy  +  (2x  +  3y  -  l)dx  =  0. 

Ansr.  x  +  y  -  4  log(2«  +  Sy  +  7)  =  C. 
(2)  Solve  (Sy  +  2x  +  ±)dx  -  (4»  +  6y  +  h)dy  =  0. 

Ansr.  9  log{(21y  +  14a;  +  22)}21(2j/  -  a?) .-  O. 

When  the  variables  cannot  be  separated  in  a  satisfactory  manner, 
special  artifices  must  be  adopted.  We  shall  find  it  the  simplest 
plan  to  adopt  the  routine  method  of  referring  each  artifice  to  the 
particular  class  of  equation  which  it  is  best  calculated  to  solve. 
These  special  devices  are  sometimes  far  neater  and  quicker  pro- 
cesses of  solution  than  the  method  just  described.  We  shall  follow 
the  conventional  x  and  y  rather  more  closely  than  in  the  earlier 
par*  of  this  work.  The  reader  will  know,  by  this  time,  that  his 
x  and  2/'s,  his  p  and  y's  and  his  s  and  t's  are  not  to  be  kept  in 
"water-tight  compartments  ".  It  is  perhaps  necessary  to  make  a 
few  general  remarks  on  the  nomenclature. 

§  121.    What  is  a  Differential  Equation? 

We  have  seen  that  the  straight  line, 

y  =  mx  +  6,  .        .        .        (1) 

fulfils  two  special  conditions :  (i)  It  cuts  one  of  the  coordinate  axes 
at  a  distance  b  from  the  origin ;  (ii)  It  makes  an  angle  tan  a  =»  m, 
with  the  a;-axis.     By  differentiation, 

dy      ™  fO\ 

This  equation  has  nothing  at  all  to  say  about  the  constant  b. 


§  121.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        375 

That  oondition  has  been  eliminated.  Equation  (2),  therefore, 
represents  a  straight  line  fulfilling  one  condition,  namely,  that 
it  makes  an  angle  tan _  lm  with  the  rc-axis.  Now  substitute  (2) 
in  (1),  the  resulting  equation, 

y  =  %x  +  h <8) 

in  virtue  of  the  constant  b,  satisfies  only  one  definite  condition, 
(3),  therefore,  is  the  equation  of  any  straight  line  passing  through 
b.  Nothing  is  said  about  the  magnitude  of  the  angle  tan  ~  1m. 
Differentiate  (2).     The  resulting  equation, 

g-0,         .        .        .        .        (4) 

represents  any  straight  line  whatever.  The  special  conditions 
imposed  by  the  constants  m  and  b  in  (1),  have  been  entirely 
eliminated.  Equation  (4)  is  the  most  general  equation  of  a 
straight  line  possible,  for  it  may  be  applied  to  any  straight  line 
that  can  be  drawn  in  a  plane. 

Let  us  now  find  a  physical  meaning  for  the  differential  equation. 
In  §  7,  we  have  seen  that  the  third  differential  coefficient,  dh/dt* 
represents  "the  rate  of  change  of  acceleration  from  moment  to 
moment".  Suppose  that  the  acceleration  d2s/dt2,  of  a  moving 
body  does  not  change  or  vary  in  any  way.  It  is  apparent  that  the 
rate  of  change  of  a  constant  or  uniform  acceleration  must  be  zero. 
In  mathematical  language,  this  is  written, 

dh/dt*  -  0 (5) 

By  integration  we  obtain, 

d2s/dt2  =  Constant  =  g.  .  .  (6) 

Equation  (6)  tells  us  not  only  that  the  acceleration  is  constant,  but 
it  fixes  that  value  to  the  definite  magnitude  g  ft.  per  second. 
Remembering  that  acceleration  measures  the  rate  of  change  of 
velocity,  and  integrating  (6),  we  get, 

ds/dt  -  gt  +  Gv  .  .  .  .  (7) 
From  §  71,  we  have  learnt  how  to  find  the  meaning  of  Gv  Put 
t  =  0,  then  ds/dt  =  Gv  This  means  that  when  we  begin  to  reckon 
the  velocity,  the  body  may  have  been  moving  with  a  definite  velocity 
Cv  Let  G1  =  v0  ft.  per  second.  Of  course  if  the  body  started  from 
a  position  of  rest,  G1  =  0.  Now  integrate  (7)  and  find  the  value  of 
C2  in  the  result, 

s  =  $gt*  +-V  +  Cv         .         .         .         (8) 


376  HIGHER  MATHEMATICS.  §  121. 

by  putting  t  m  0.  It  is  thus  apparent  that  G2  represents  the  space 
which  the  body  had  traversed  when  we  began  to  study  its  motion. 
Let  C2  *=  s0  ft.     The  resulting  equation 

*  =  $gt2  +  v0t  +  «o,  .  .  .  (9) 
tells  us  three  different  things  about  the  moving  body  at  the  instant 
we  began  to  take  its  motion  into  consideration. 

1.  It  had  traversed  a  distance  of  s0  ft.  To  use  a  sporting  phrase, 
if  the  body  is  starting  from  "  scratch,"  s0  =  0. 

2.  The  body  was  moving  with  a  velocity  of  vQ  ft.  per  second. 

3.  The  velocity  was  increasing  at  the  uniform  rate  of  g  ft.  per 
second. 

Equation  (7)  tells  us  the  two  latter  facts  about  the  moving  body ; 
equation  (6)  only  tells  us  the  third  fact ;  equation  (5)  tells  us  no- 
thing more  than  that  the  acceleration  is  constant.  (5),  therefore,  is 
true  of  the  motion  of  any  body  moving  with  a  uniform  acceleration. 

Examples. — (1)  A  body  falls  from  rest.  Show  that  it  travels  400  ft.  in 
5  sec.    Hint.  Use  g  =  32. 

(2)  A  body  starting  with  a  velocity  of  20  ft.  per  sec.  falls  in  accord  with 
equation  (7) ;  what  is  its  velocity  after  6  seconds  ?    Ansr.  212  ft. 

(3)  A  body  dropped  from  a  balloon  hits  the  ground  with  a  velocity  of 
384  ft.  per  sec.     How  long  was  it  falling  ?     Ansr.  12  seconds. 

(4)  A  particle  is  projected  vertically  upwards  with  a  velocity  of  100  ft. 
per  sec.  Find  the  height  to  which  it  ascends  and  the  time  of  its  ascent. 
Here  d?sjdta  —  -  g ;  multiply  by  2ds/dt,  and  integrate 

ds  &* 
Adt '  dt* 
when  the  particle  has  reached  its  maximum  height  dsjdt  =  0 ;  and,  therefore, 

*  -  Wl9  =  1±irL !  f™m  (7).  since  C,  =  100  «  v0,  t  =  vjg  -  >#• 

(5)  If  a  body  falls  in  the  air,  experiment  shows  that  the  retarding  effect 
of  the  resisting  air  is  proportional  to  the  square  of  the  velocity  of  the  moving 
body.  Instead  of  g,  therefore,  we  must  write  g  -  bv2,  where  b  is  the  variation 
constant  of  page  22.      For  the  sake  of  simplicity,  put  b  =  g\o?  and  show  that 

e9t la  _  e-gt\a  a*         ggtfa  +  e~gtla        tf  gj, 

v  =  V/«  +  g-*'»;  5  ="  7l0« 2 "  ^iogoosh-, 

since  v  =  0,  when  t  —  0,  and  5=0  when  £=0.  Hint.  The  equation  of  motion 
i3  dvjdt  =»  g  -  bv*. 

Similar  reasoning  holds  good  from  whatever  sources  we  may 
draw  our  illustrations.  We  are,  therefore,  able  to  say  that  a 
differential  equation,  freed  from  constants,  is  the  most 
general  way  of  expressing  a  natural  law. 

Any  equation  can  be  freed  from  its  constants  by  combining  it 


d\dt)  ft  da        /ds\* 


§121.     HOW  .TO  SOLVE  DIFFERENTIAL  EQUATIONS.        377 

with  the  various  equations  obtained  by  differentiation  of  the*  given 
equation  as  many  times  as  there  are  constants.  The  operation  is 
called  elimination.  Elimination  enables  us  to  discard  the  ac- 
cidental features  associated  with  any  natural  phenomenon  and  to 
retain  the  essential  or  general  characteristics.  It  is,  therefore, 
possible  to  study  a  theory  by  itself  without  the  attention  being 
distracted  by  experimental  minutiae.  In  a  great  theoretical  work 
like  "Maxwell"  or  "Heaviside,"  the  differential  equation  is 
ubiquitous,  experiment  a  rarity.  And  this  not  because  experi- 
ments are  unimportant,  but  because,  as  Heaviside  puts  it,  they 
are  fundamental,  the  foundations  being  always  hidden  from  view 
in  well-constructed  buildings. 

Examples. — (1)  Eliminate  the  arbitrary  constants  a  and  6,  from  the 
relation  y  =  ax  +  bx2.  Differentiate  twice ;  evaluate  a  and  b ;  and  substitute 
the  results  in  the  original  equation.     The  result, 

is  quite  free  from  the  arbitrary  restrictions  imposed  in  virtue  of  the  presence 
of  the  constants  a  and  b  in  the  original  equation. 

(2)  Eliminate  m  from  y*  =  4maj.    Ansr.  y  m  2x .  dyfdx. 

(3)  Eliminate  a  and  b  from  y  ■■  a  cob*  +  6  sin  x.    Ansr.  d^yjdx2  +  y  *«  0. 

We  always  assume  that  every  differential  equation  has  been 
obtained  by  the  elimination  of  constants  from  a  given  equation 
called  the  primitive.  In  practical  work  we  are  not  so  much 
concerned  with  the  building  up  of  a  differential  equation  by  the 
elimination  of  constants  from  the  primitive,  as  with  the  reverse 
operation  of  finding  the  primitive  from  which  the  differential 
equation  has  been  derived.  In  other  words,  we  have  to  find 
some  relation  between  the  variables  which  will  satisfy  the  differ- 
ential equation.  Given  an  expression  involving  x,  y,  dx/dy, 
d'2x/dy2, . . .,  to  find  an  equation  containing  only  x,  y  and  con- 
stants which  can  be  reconverted  into  the  original  equation  by  the 
elimination  of  the  constants. 

This  relation  between  the  variables  and  constants  which  satisfies 
the  given  differential  equation  is  called  a  general  solution,  or  a 
complete  solution,  or  a  complete  integral  of  the  differential 
equation.  A  solution  obtained  by  giving  particular  values  to  the 
arbitrary  constants  of  the  complete  solution  is  a  particular  solu- 
tion. Thus  y  =  mx  is  a  complete  solution  of  y  =  x .  dy/dx ; 
y  —  x  tan  45°,  is  a  particular  solution. 


378  HIGHER  MATHEMATICS.  §  122. 

A  differential  equation  is  ordinary  or  partial,  according  as 
there  is  one  or  more  than  one  independent  variables  present. 
Ordinary  differential  equations  will  be  treated  first.  Equations 
like  (2)  and  (3)  above  are  said  to  be  of  the  first  order,  because 
the  highest  derivative  present  is  of  the  first  order.  For  a  similar 
reason  (4)  and  (6)  are  of  the  second  order,  (5)  of  the  third  order. 
The  order  of  a  differential  equation,  therefore,  is  fixed  by  that 
of  the  highest  differential  coefficient  it  contains.  The  degree  of  a 
differential  equation  is  the  highest  power  of  the  highest  order  of 
of  differential  coefficient  it  contains.  This  equation  is  of  the  second 
order  and  first  degree : 

It  is  not  difficult  to  show  that  the  complete  integral  of  a  differ- 
ential equation  of  the  nth  order,  contains  n,  and  no  more  than  n, 
arbitrary  constants.  As  the  reader  acquires  experience  in  the 
representation  of  natural  processes  by  means  of  differential  equa- 
tions, he  will  find  that  the  integration  must  provide  a  sufficient 
number  of  undetermined  constants  to  define  the  initial  conditions 
of  the  natural  process  symbolized  by  the  differential  equation.  The 
complete  solution  must  provide  so  many  particular  solutions  (con- 
taining no  undetermined  constants)  as  there  are  definite  conditions 
involved  in  the  problem.  For  instance,  equation  (5),  page  375,  is 
of  the  third  order,  and  the  complete  solution,  equation  (9),  requires 
three  initial  conditions,  g,  s0,  v0  to  be  determined.  Similarly,  the 
solution  of  equation  (4),  page  375,  requires  two  initial  conditions, 
m  and  b,  in  order  to  fix  the  line. 

§  122.    Exact  Differential  Equations  of  the  First  Order. 

The  reason  many  differential  equations  are  so  difficult  to  solve 
is  due  to  the  fact  that  they  have  been  formed  by  the  elimination 
of  constants  as  well  as  by  the  elision  of  some  common  factor  from 
the  primitive.  Such  an  equation,  therefore,  does  not  actually  re- 
present the  complete  or  total  differential  of  the  original  equation 
or  primitive.  The  equation  is  then  said  to  be  inexact.  On  the 
other  hand,  an  exact  differential  equation  is  one  that  has  been 
obtained  by  the  differentiation  of  a  function  of  x  and  y  and  per- 
forming no  other  operation  involving  x  and  y. 

Easy  tests   have   been  described,  on  page   77,  to   determine 


§  122.     HOW  TO  SOLVE  DIFFEKENTIAL  EQUATIONS.        379 

whether  any  given  differential  equation  is  exact  or  inexact.  It 
was  pointed  out  that  the  differential  equation, 

M.dx  +  N.dy~Qt        .        .        .        (1) 

is  the  direct  result  of  the  differentiation  of  any  function  u,  provided, 

This  last  result  was  called  "  the  criterion  of  integrability,"  because, 
if  an  equation  satisfies  the  test,  the  integration  can  be  readily  per- 
formed by  a  direct  process.  This  is  not  meant  to  imply  that  only 
such  equations  can  be  integrated  as  satisfy  the  test,  for  many  equa- 
tions which  do  not  satisfy  the  test  can  be  solved  in  other  ways. 

Examples.— (1)  Apply  the  test  to  the  equations,  ydx  +  xdy  =  0,  and 
ydx  -  xdy  =  0.  In  the  former,  M ~  y,  N -,  x\  .-.  dM/dy  =  1,  dtydx  -  U 
.*.  ?)Mfdy  =*  dN/dx.  The  test  is,  therefore,  satisfied  and  the  equation  is  exact. 
In  the  other  equation,  M  =  y,  N  =  -  x,  .:  ?>Mfdy  =  1,  "dNfdx  =  -  1.  This 
does  not  satisfy  the  test.  In  oonsequence,  the  equation  cannot  be  solved  by 
the  method  for  exact  differential  equations. 

(2)  Show  that  {a?y  +  x*)dx  +  (b3  +  a?x)dy  =  0,  is  exact. 

(8)  Is  the  equation,  (x  +  2y)xdx  +  (x*  -  y*)dy  =  0,  exaot  ?  M  —  x(x  +  2y), 
N=x*  -y2;  .-.  dMfdy  *  2x,  dNfdx  =  2x.  .  The  condition  is  satisfied,  the 
equation  is  exact. 

(4)  Show  that  (siny  +  y  coBx)dx  +  (since  +  x cos y)dy  =  0,  is  exaot. 

I.  Equations  which  satisfy  the  criterion  of  integrability. 
We  must  remember  that  M  is  the  differential  coefficient  of  u 
with  respect  to  x,  y  being  constant,  and  N  is  the  differential  co- 
efficient of  u  with  respect  to  y,  x  being  constant.  Hence  we  may 
integrate  Mdx  on  the  supposition  that  y  is  constant  and  then  treat 
Ndy  as  if  x  were  a  constant.  The  complete  solution  of  the  whole 
equation  is  obtained  by  equating  the  sum  of  these  two  integrals  to 
an  undetermined  constant.     The  complete  integral  is 

u  =  O (3) 

Example. — Integrate  x(x  +  2y)dx  +  (x2  -  y2)dy  =  0,  from  the  preceding 
set  of  examples.  Since  the  equation  is  exact,  M  =  x(x  +  2y) ;  JV«o  x2  -  y2; 
.'.  jMdx  =  jx(x  +  2y)dx  =  %x3  +  x^y  =  Y,  where  Fis  the  integration  constant 
which  may,  or  may  not,  contain  y,  because  y  has  here  been  regarded  as  a  con- 
stant. Now  the  result  of  differentiating  %xz  +  x*y  =  F,  should  be  the  original 
equation.  On  trial,  x2dx  +  2xydx  +  x2dy  —  dY.  On  comparison  with  the 
original  equation,  it  is  apparent  that  dY  =  y2dy ;  .*.  F  =  ^y3  +  C.  Sub- 
stitute this  in  the  preceding  result.  The  complete  solution  is,  therefore, 
±2?  +  x^y  -  ly3  =  C.  The  method  detailed  in  this  example  can  be  put  into  a 
more  practical  shape. 


380  HIGHER  MATHEMATICS.  §  122. 

To  integrate  an  exact  differential  equation  of  the  type 

M .  dx  +  N .  dy  =  0, 

first  find  jM .  dx  on  the  assumption  that  y  is  constant  and  substi- 
tute the  result  in 

E.g.,  in  x(x  +  2y)dx  +  (x2  -  y2)dy  =  0,  it  is  obvious  that  jMdx  is 
Ja?8  +  x2y,  and  we  may  write  down  at  once 

^3  +   X2y   +  jj^  _  y,  _   ^3  +  x^dy  =   G. 

.\  $x*  +  a%  +  j(x2  -  y2  -  x2)dy  =  C;  or,  \xz  +  x2y  -  \y*  -  C. 
If  we  had  wished  we  could  have  used 


^Ndy  +^(m  -^Ndy^dx  =  C, 


in  place  of  (4),  and  integrated  jN .  dy  on  the  assumption  that  x  is 
constant.  •» 

In  practice  it  is  often  convenient  to  modify  this  procedure.  If 
the  equation  satisfies  the  criterion  of  integrability,  we  can  easily 
pick  out  terms  which  make  Mdx  +  Ndy  *=  0,  and  get 

Mdx  +  Y;  and  Ndy  +  X, 
where  T  cannot  contain  x  and  X  cannot  contain  y.  Hence  if  we 
can  find  Mdy  and  Ndx,  the  functions  X  and  T  will  be  determined. 
In  the  above  equation,  the  only  terms  containing  x  and  y  are 
%xydx  +  x2dy,  which  obviously  have  been  derived  from  x2y.  Hence 
integration  of  these  and  the  omitted  terms  gives  the  above  result. 

Examples.— (1)  Solve  (x2  -  ±xy  -  2y2)dx  +  (y2  -  4<cy  -  2x2)dy  =  0.  Pick 
out  terms  in  x  and  y,  we  get  -  (±xy  +  2y2)dx  -  (±xy  +  2x*)dy  =0.  Integrate. 
.•.  -  2x2y  -  2xy2  =  constant.  Pick  out  the  omitted  terms  and  integrate  for 
the  complete  solution.    We  get, 

jx2dx  +  jy2dy  -  2aty  -  2xy2  =  O ;  .'.x'A  -  5x*y  -  6xy2  +  y3  =»  constant. 

(2)  Show  that  the  solution  of  (a?y  +  x2)dx  +  (63  +  a2x)dy  =  0,  furnishes 
the  relation  a?xy  +  b3y  +  $xs  =  C.    Use  (4). 

(3)  Solve  {x2  -  y2)dx  -  2xydy  =  0.    Ansr.  %x2  -  y*  =  Cfx.    Use  (4). 

II.  Equations  which  do  not  satisfy  the  criterion  of  integrability. 
As  just  pointed  out,  the  reason  any  differential  equation  does 
not  satisfy  the  criterion  of  exactness,  is  because  the  "  integrating 
factor  "  has  been  cancelled  out  during  the  genesis  of  the  equation 
from  its  primitive.  If,  therefore,  the  equation 
Mdx  +  Ndy  »=  0, 


§  123.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        381 

does  not  satisfy  the  criterion  of  integrability,  it  will  do  so  when 
the  factor,  previously  divided  out,  is  restored.  Thus,  the  pre- 
ceding equation  is  made  exact  by  multiplying  through  with  the 
integrating  factor  /x.     Hence, 

^(Mdx  +  Ndy)  =  0, 
satisfies  the  criterion  of  exactness,  and  the  solution  can  be  obtained 
as  described  above. 

§  123.   How  to  find  Integrating  Factors. 

Sometimes  integrating  factors  are  so  simple  that  they  can  be 
detected  by  simple  inspection. 

Examples. — (1)  ydx  -  xdy  «  0  is  inexact.  It  becomes  exact  by  multipli- 
cation with  either  x -2,  x-1  ,y~\  or  y-2. 

(2)  In  (y  -  x)dy  +  ydx  —  0,  the  term  containing  ydx  -  xdy  is  not  exact, 
but  becomes  so  when  multiplied  as  in  the  preceding  example. 

dy     xdy  -  ydx     .  .  x      ~ 

.'.  J  -  -~ f- 0;  or,  logy  -  -  -  0. 

We  have  already  established,  in  §  $6,  that  an  integrating  factor 
always  exists  which  will  make  the  equation 

Mdx  +  Ndy  m  0, 
an  exact  differential.  Moreover,  there  is  also  an  infinite  number 
of  such  factors,  for  if  the  equation  is  made  exact  when  multiplied 
by  /a,  it  will  remain  exact  when  multiplied  by  any  function  of  //,. 
The  different  integrating  factors  correspond  to  the  various  forms 
in  which  the  solution  of  the  equation  may  present  itself.  For 
instance,  the  integrating  factor  a;"1^"1,  of  ydx  +  xdy  =  0,  corre- 
sponds with  the  solution  log  a?  +  logy  =  G.  The  factor  y~2  corre- 
sponds with  the  solution  xy  «  G.  Unfortunately,  it  is  of  no 
assistance  to  know  that  every  differential  equation  has  an  infinite 
number  of  integrating  factors.  No  general  practical  method  is 
known  for  finding  them ;  and  the  reader  must  consult  some  special 
treatise  for  the  general  theorems  concerning  the  properties  of  in- 
tegrating factors.  Here  are  four  elementary  rules  applioable  to 
special  cases. 

Role  I.  Since 

d(xmyn)  =  xm~lyn-\mydx  +  nxdy), 
an  expression  of  the  type  mydx  +  nxdy  =  0,  has  an  integrating 
factor  xm~1yn~1 ;  or,  the  expression 

x«yP(mydx  +  nxdy)  =  0,     .         •        .         (1) 


382  HIGHER  MATHEMATICS.  §  123. 

has  an  integrating  factor 

or  more  generally  still, 

ajj—i-y*-!-^  .        .        .        (2) 

where  k  may  have  any  value  whatever. 

Example. — Find  an  integrating  factor  of  ydx  -  xdy  =  0.  Here,  a  =  0, 
fi  =  0,  m  =  1,  n  «=  -  1  /.  y-2  is  an  integrating  factor  of  the  given  equation. 

If  the  expression  can  be  written 

x^imydx  +  nxdy)  +  xa'yP'(m'ydx  +  n'xdy)  =  0,  .         (3) 
the  integrating  factor  can  be  readily  obtained,  for 

x*n  -  1  -  aykn  -  1  -  fi  .    ftnd  tf'm'  -  1  -  ay^  -  1  -  ^ 

are  integrating  factors  of  the  first  and  second  members  respectively. 
In  order  that  these  factors  may  be  identical, 

km  -  1  -  a  =  k'm'  -  1  -  a' ;  kn  -  1  -  /J  =  k'n'  -  1  -  p. 
Values  of  k  and  k'  can  be  obtained  to  satisfy  these  two  conditions 
by  solving  these  two  equations.     Thus, 

h       n'(a  -  a')  -  rn'tf  -  /?') .   v  _  n(a  -  a')  -  m(/3  -  ff) 

wn'  -  m'n  *  mn'  -  m'n  '     ^  ' 

Examples.— (J)  Solve  y*{ydx  -  2xdy)  +  x*(2ydx  +  xdy)  =  0.  Hints.  Show 
that  a  =  0,  0  =  3,  m  =  1,  n  =  -  2  ;  o'  =  4,  0'  =  0,  m'  =  2,  n'  =  1 ;  .-. 
xk-iy-vt-4  ^  an  integrating  factor  of  the  first,  xw ~ 5y*' "" 1  of  the  second 
member.  Hence,  from  (4),  k  m  -  2,  k'  =  1,  .*.  x~5  is  an  integrating  factor  of 
the  whole  expression.     Multiply  through  and  integrate  for  2x*y  -  y4  =  Ccc2. 

(2)  Solve  (y3  -  2ya;2)da;  +  (2a*/2  -  ic3)^  =  0.  Ansr.  afyV  -  a;2)  =  0.  In- 
tegrating factor  deduced  after  rearranging  the  equation  is  xy. 

Rule  II.   If  the  equation  is  homogeneous  and  of  the  form : 
Mdx  +  Ndy  =  0,  then  (Mx  +  Ny)  ~1  is  an  integrating  factor. 
Let  the  expression 

Mdx  +  Ndy  =  0 
be  of  the  mth  degree  and  /x,  an  integrating  factor  of  the  nth  degree, 
.-.  fiMdx  +  fjiNdy  =  du,     .         .         .         (5) 
is  of  the  (m  +  n)th  degree,  and  the  integral  u  is  of  the  (m  +  n  +  l)th 
degree.     By  Euler's  theorem,  §  23, 

.-.  fiMx  +  fiNy  =  (m  +  n  +  l)u.      .        .        (6) 
Divide  (5)  by  (6), 

Mdx  +  Ndy  _  1  du 

Mx  +  Ny    ~  m  +  n  +  1'  ~u~' 
The  right  side  of  this  equation  is  a  complete  differential,  conse- 


§  123.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.         383 

quently,  the  left  side  is  also  a  complete  differential.  Therefore, 
the  factor  (Mx  +  Ny)  " 1  has  made  Mdx  +  Ndy  =  0  an  exact 
differential  equation. 

Examples. — (1)  Show  that  (x*y  -  xy3)-1  ia  an  integrating  factor  of 
i??y  +  y3)dx  -  2xy2dy  =  0. 

(2)  Show  that  l/(a;2  -  nyx +ya)  is  an  integrating  factor  of  ydy  +  (x-  ny)dx =0. 

The  method,  of  oourse,  cannot  be  used  if  Mx  +  Ny  is  equal  to  zero.  In 
this  case,  we  may  write  y  **  Cx,  a  solution. 

Rule  III.  If  the  equation  is  homogeneous  and  of  the  form : 

A(x>  y)vdx  +  /»(*!  v)xdv  =  o> 

then  (Mx  -  Ny)  "  *  is  an  integrating  factor. 

Example.— Solve  (1  +  xy)ydx  +  (1  -  xy)xdy  =  0.  Hint.  Show  that  the 
integrating  factor  is  l/2a;V-    Divide  out  £.    .'.  \Mdx  =  -  ljxy  +  log  x.    Ansr. 

x  =  Cye . 

y 
If  Mx  -  Ny  =  0,  the  method  fails  and  xy  m  G  is  then  a  solu- 
tion of  the  equation.     E.g.,  (1  +  xy)ydx  +  (1  +  xy)xdy  =.  0. 

Rule  IY.  If  jd-^ y-jwa  function  of  x  only,  ejAx)dx  is  an 

integrating  factor.  Or,  if  -g^  -  -^  J  -  f(y),  then  e^)dv  is  an 
integrating  factor.     These  are  important  results. 

Examples.— (1)  Solve  (as2  +  y%)dx  -  2xydy  =  0.  Ansr.  x2  -  y2  =  Gx. 
Hint.  Show  f(x)=  -Ste-l.  The  integrating  factor  is  *-'**-i*=a=e-i»<** 
=  a;-2.  Prove  that  this  is  an  integrating  factor,  and  solve  as  in  the  preced- 
ing section. 

(2)  Solve  {y4  +  2y)dx  +  (xy3  +  2y*  -  ix)dy  =  0.    Ansr.  xy2+  y4  +  2x=Cy2. 

We  may  now  illustrate  this  rule  for  a  special  case,  as  we  shall 
want  the  result  later  on.  The  steps  will  serve  to  recall  some  of  the 
principles  established  in  some  earlier  chapters.     Let 

|  +  Pj,  =  0,    .       .       .       .       (7) 

where  P  and  Q  are  either  constants  or  functions  of  x.  Let  fi  be  an 
integrating  factor  which  makes 

ay  +  (Py  -  Q)dx  «  0,       .         .         .         (8) 
an  exact  differential. 

.-.  fidy  +  ix(Py  -  Q)dx  =  Ndy  +  Mdx. 

•••  is  -  £ •  %  -  w  -  % + p"-  •••  s§- <**  -  «>sf  +  *v 

•••  £dx  -  (Py  -  <?)|^r  +  P^fe  =  -  |fidy  +  P^te. 


384  HIGHER  MATHEMATICS.  §  124. 

•*•  £dx  +  tydy  =  */*  -  pfjdx-  •*• p  -  -  -£ >  •*•  $pdx  =  log/* ; 

and  since  log^e  =  1,  (jPdx)  log  e  =  log  ju,  ;  consequently 

.ygL-fP* (9) 

is  the  integrating  factor  of  the  given  equation  (7). 

§  124.    Physical  Meaning  of  Exact  Differentials. 

Let  AP  (Fig.  144)  be  the  path  of  a  particle  under  the  influence 
of  a  force  F  making  an  angle  0  with  the  tangent 
PT  of  the  curve  at  the  point  P(x,  y).  Let  W 
denote  the  work  done  by  the  particle  in  passing 
from  the  fixed  point  A(a,  b)  to  its  present  position 
P(x,  y).  Let  the  length  AP  be  s.  The  work, 
dW,  done  by  the  particle  in  travelling  a  distance 
ds  will  now  be 

dW  =  F.  gob e.ds.  (1) 

Let  PT  and  PF  respectively  make  angles  a  and  /?  with  the  #-axis. 
Hence,  as  on  page  126,  dx/ds  =  cos  a  ;  dy/ds  =»  sin  a  ;  .*.  0  =  a  -  ft 

.-.  .Foos  0  =  .Fcos  (a  -  j3)  =  J7 cos  a .  cos/3  +  Fqw.  a .  sin  ft 
by  a  well-known  trigonometrical  transformation  (24)  page  612. 

,.F».-F~.&+F*L&_JI»+y%.         (2) 

where  X  is  put  for  F  cos  ft  and  Y  for  F  sin  ft;  Xand  Fare  ob- 
viously the  two  components  of  the  force  parallel  with  the  coordinate 
axes.     From  (1), 

<*-(*£+*D*.  •  •  •  w 

I.  Let  Xdx  +  Ydy  be  a  complete  differential. 

Let  us  assume  that  Xdx  +  Ydy  is  a  complete  differential  of  the 
function  u  =  f(x,  y).     Hence 

1Trr       fin   dx      ~bu   dy\ , 

by  partial  differentiation.  In  order  to  fix  our  ideas,  let 
u  =  tan  "  l(yl%)  Fig.  144.     Hence,  Ex.  (5),  page  49, 

**  -  ^%  -  ^», 
where  r2  is  put  in  place  of  x2  +  y2.     From  (4), 


^TF  _  fx    dy       y  dx\  _  du 
"aT  ~  \^'  ds  ~  r^ds)  =  ds' 


§  124.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        385 

The  rate  dW/ds  at  which  the  work  is  performed  by  the  particle 
changes  as  it  moves  along  the  curve  and  is  equal  to  the  rate,  du/ds, 
at  which  the  function  f(x,  y)  changes.  Any  change  in  W  is  accom- 
panied by  a  corresponding  change  in  the  value  of  u.  Hence,  as 
the  particle  passes  from  A  to  P,  the  work  performed  will  be 

rl  x  ' 

and  by  integration, 

W  =  U  +  constant. 

This  means  that  the  work  done  by  the  particle  in  passing  from  a 
fixed  point  A  to  another  point  P(x,  y)  depends  only  upon  the  value 
of  u,  and  u  is  a  function  of  the  coordinates,  x  and  y,  of  the  point  P. 
It  will  be  obvious  that  if  the  particle  moves  along  a  closed 
curve  the  work  done  will  be  zero.  If  the  origin  0  lies  within  the 
closed  curve,  u  will  increase  by  2-n-  when  P  has  travelled  round  the 
curve.  In  that  case  the  work  done  is  not  zero.  The  function  u  is 
then  a  multi- valued  function. 

Example. — If  X  =  y,  and  Y  =  x,  dW  =  (xdx  +  ydy)  =  d(xy) ;  or,  by  in- 
tegration W  =  xy  +  C.  We  do  not  need  to  know  the  equation  of  the  path. 
The  work  done  is  simply  a  function  of  the  coordinates  of  the  end  state.  The 
constant  C  serves  to  define  the  initial  position  of  the  point  A  (a,  b). 

The  first  law  of  thermodynamics  states  that  when  a  quantity  of 
heat,  dQ,  is  added  to  a  substance,  one  part  of  the  heat  is  spent  in 
changing  the  internal  energy,  dU,  of  the  substance  and  another 
part,  dW,  is  spent  in  doing  work  against  external  forces.  In 
symbols, 

dQ  =  dU  +  dW. 

In  the  special  case,  when  that  work  is  expansion  against  atmospheric 
pressure,  dW  =  p.dv.    Now  let  the  substance  pass  from  any  state 
A  to  another  state  B  (Fig.  145).     The  internal  energy  of  the  sub- 
stance in  the  state  B  is  completely  deter- 
mined by  the  coordinates  of  that   point, 
because    U  is    quite    independent   of   the 
nature    of    the    transformation   from   the 
state  A   to  the  state   B.     It   makes   no 
difference  to  the  magnitude  of  U  whether 
that  path  has   been   vid  ABB    or   AQB. 


jp 


B' 

In  this  case  U  is  completely  defined  bv  ™      «,,- 

,.  ,    !  •  -.  FlG-  145- 

the  coordinates  of  the  point  corresponding 

to  any  given  state.     In  other  words  d  U  is  a  complete  differential. 

BB 


386  HIGHER  MATHEMATICS.  §  124. 

On  the  other  hand,  the  external  work  done  during  the  trans- 
formation from  the  one  state  to  another,  depends  not  only  on  the 
initial  and  final  states  of  the  substance,  but  also  on  the  nature  of 
the  path  described  in  passing  from  the  state  A  to  the  state  B. 
For  example,  the  substance  may  perform  the  work  represented  by 
the  area  AQBB'A'  or  by  the  area  APBB'A',  in  its  passage  from  the 
state  A  to  the  state  B.  In  fact  the  total  work  done  in  the  passage 
from  A  to  B  and  back  again,  is  represented  by  the  area  APBQ. 
In  order  to  know  the  work  done  during  the  passage  from  the  state 
A  to  the  state  B,  it  is  not  only  necessary  to  know  the  initial  and 
final  states  of  the  substance  as  defined  by  the  coordinates  of  the 
points  A  and  B,  but  we  must  know  the  nature  of  the  path  from 
the  one  state  to  the  other. 

Similarly,  the  quantity  of  heat  supplied  to  the  body  in  passing 
from  one  state  to  the  other,  not  only  depends  on  the  initial  and  final 
states  of  the  substance,  but  also  on  the  nature  of  the  transforma- 
tion. All  this  is  implied  when  it  is  said  that  "  dW  and  dQ  are  not 
perfect  differentials  ".  dW  and  dQ  can  be  made  into  complete  differ- 
entials by  multiplying  through  with  the  integrating  factor  /x.  The 
integrating  factor  is  proved  in  thermodynamics  to  be  equivalent  to 
the  so-called  Carnot's  function.  To  indicate  that  d W  and  dQ  are 
not  perfect  differentials,  some  writers  superscribe  a  comma  to  the 
top  right-hand  corner  of  the  differential  sign.  The  above  equation 
would  then  be  written, 

d'Q  =  dU  +  d'W. 

II.  Let  Xdx  +  Ydy  be  an  incomplete  differential. 

Now  suppose  that  Xdx  +  Ydy  is  not  a  complete  differential. 
In  that  case,  we  cannot  write  X  =  ~bu[bx  and  Y  =  'dufby  as  in  (3) 
and  (4).  But  from  equation  (3),  by  a  rearrangement  of  the  terms, 
we  get 

dW=(x+Y^)dx.        ...        (5) 

And  now,  to  find  the  work  done  by  the  particle  in  passing  from  A 
to  P,  we  must  be  able  to  express  y  in  terms  of  x  by  using  the 
equation  of  the  path.  Let  X  =  -  y,  and  Y  =  x  ;  let  the  equation 
of  the  path  be  y  =  ax2,  .'.  dy/dx  =  2ax.     From  (5) 

dW  =  ( -  y  +  2ax2)dx  =  ( -  ax2  +  2ax2)dx  =  %ax*  +  G. 

It  is  now  quite  clear  that  the  value  of  X+  Ydy/dx  will  be  different 


§  125.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        387 

for  different  paths.     For  example,  if  y  =  ax3, 

dW  =  (ax3  +  Sax3)dx  =  ax\ 
So  that  the  work  done  depends  upon  the  coordinates  of  the  point 
P  as  well  as  upon  the  equation  of  the  path. 

Example. — If  dU=dQ  -  pdv,  and  dU  is  a  oomplete  differential,  show 
that  dQ  is  not  a  complete  differential.     Hint.  We  know,  page  80,  that 

*»-(&)**+ (IS,*'  •••<w=(l§W  (g -*),*•  <6> 

If  U  is  a  complete  differential, 

From' (6),  if  dQ  is  a  complete  differential, 

32Q   _  d2Q 
dvdT     dTdv' 
Hence  (6)  and  (7)  cannot  both  be  true. 

The  question  is  discussed  from  another  point  of  view  in  Technics, 
1,  615,  1904, 

§  125.    Linear  Differential  Equations  of  the  First  Order. 

A  linear  differential  equation  of  the  first  order  involves  only 
the  first  power  of  the  dependent  variable  y  and  of  its  first  differ- 
ential coefficients.  The  general  type,  sometimes  called  Leibnitz' 
equation,  is 

%  +  Py-Q,    •      •      •      •      (i) 

where  P  and  Q  may  be  functions  of  x  and  explicitly  independent 
of  y,  or  constants.  We  have  just  proved  that  e/pdx  is  an  integrat- 
ing factor  of  (1),  therefore 

efpdx(dy  +  Pydx)  -  e/pdxQdx, 

is  an  exact  differential  equation.  Consequently,  the  general  solu- 
tion is, 

ye'***  =  \efpdxQdx  +  C;  or,  y  -  e-/pdx\e/pdxQdx  +  Ce-/Pd*.  (2) 

The  linear  equation  is  one  of  the  most  important  in  applied  mathe- 
matics. In  particular  cases  the  integrating  factor  may  assume  a 
very  simple  form. 

Examples. — (1)  Solve  (1  +  x2)dy  =  (m  +  xy)dx.  Eeduce  to  the  form  (1) 
and  we  obtain 

dy  x        _      m 

dx  ~  T+~x^y  ~  T+1F 
BB  * 


388  HIGHER  MATHEMATICS.  §  125. 

//*  xdx 

Remembering  logl  =0,  loge  =  1,  the  integrating  factor  is  evidently, 
log  e"***  =  log  1  -  log /s/ITaJ5 ;  or  ef*<  =  Jh  +  g& 

Multiply  the  original  equation  with  this  integrating  factor,  and  solve  the 
resulting  exact  equation  as  §  122,  (4),  or,  better  still,  by  (2)  above.  The 
solution :  y  =  mx  +  G  ^/(l  +  x2)  follows  at  once. 

(2)  Ohm's  law  for  a  variable  current  flowing  in  a  circuit  with  a  coefficient 
of  self-induction  L  (henries),  a  resistance  B  (ohms),  and  a  current  of  G 
(amperes)  and  an  electromotive  force  E  (volts),  is  given  by  the  equation, 
E  =  BG  +  LdCjdt.  This  equation  has  the  standard  linear  form  (1).  If  E 
is  constant,  show  that  the  solution  is,  C  =  EjB  +  Be~Rt,L,  where  B  is  the 
arbitrary  constant  of  integration  (page  193).  Show  that  G  approximates  to 
E\B  after  the  current  has  been  flowing  some  time,  t.  Hint  for  solution. 
Integrating  factor  is  em,L. 

(3)  The  equation  of  motion  of  a  particle  subject  to  a  resistance  varying 
directly  as  the  velocity  and  as  some  force  which  is  a  given  function  of  the 
time,  is  dv/dt  +  kv  =  f(t).  Show  that  v  =  C«-**  +  «-»f«»/($)<«.  If  the  force 
is  gravitational,  say  g,  v  =  Ge~u  +  gjk. 

(4)  Solve  xdy+ydx=xsdx.    Integrating  factor = x.     Ansr.  y^a^  +  G/x. 

(5)  We  shall  want  the  integral  of  dyjdt  +  k2y  =  k2a(l  -  e-*if)  very  shortly. 
The  solution  follows  thus : 

y  =  Ce-'W  -  e-'**uf*r*gp{  -  k2a(l  -  «-*i«)}<&; 
=  Ce  -  *tf    +  e  -  **{k2aje**  -  je^-h^dt ; 
=  Ce-**    +a-JWe_htt 

(6)  We  shall  also  want  to  solve 

dy        Ky  Kx    . 


Here 


dx      a  —  x      a  —  x 


(Kdx 
e  a~x  =  e-*log(«-x)  =  e-log[a-x)K  = 


(a  -  x)R 

.         V        _r,    K[       ^dx  x        _  v[     dx 

' '  {a  -  x)*  ~  "  "*"  ^J  (a  -  x)^  +  1  ~  u  +  (a  -  x)*         J  (a  -  x)* 

on  integrating  by  parts.     Finally,  if  x  *=  0,  when  y  =  0, 

K(a  -  x)     „  1 

y  =  C(a-x)x  +  a--±r-I>-,  C  =  {K_1)aK_1. 

Many  equations  may  be  transformed  into  the  linear  type  of 
equation,  by  a  change  in  the  variable.  Thus,  in  the  so-called 
Bernoulli's  equation, 

%  +  Py  =  Qyn.        ...       (3) 

Divide  by  yn,  multiply  by  (1  -  n)  and  substitute  yl~n  =  v,  in  the 
result.     Thus, 


§  125.      HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        389 

...^  +  (1  -  n)Pv  =  0(1  -  n), 

which  is  linear  in  v.     Hence,  the  solution  follows  at  sight, 
veii -nvpdx  =  (1  _  ri)jQe{1-n)/pd*dx  +  C. 
.-.  yi-»ell-nVPdx  =  (1  -  ri)\Qe(l-n)fpdxdx  +  G. 

Examples. — (1)  Solve  dyjdx  +  yjx  =  y2.  Substitute  v  =  Ijy.  Integration 
factor  is  e -/<**/*  =  g-iog*  =  x~x.     Ansr.  Cxy  -  xylogx  =  1. 

(2)  Solve  dyjdx  +  x  sin  2y  =  a^cos2?/.  Divide  by  cos2!/.  Put  tan  y  =  v. 
The  integration  factor  is  e^***,  i.e.,  e3?.  Ansr.  e*2(tany  -  %x2  +  %)  =  G. 
Hint.  The  steps  are  sec2?/  dyjdx  +  2a;  tan y  =  a;3;  dvjdx  +  2xu  =  a;3;  to  solve 
ve*2  =  jx^dx  +  C.  Put  x2  =  z,  .'.  2xdx  m  dz,  and  this  integral  becomes 
\\zezdz\  or,  \&{z  -  1),  (4)  page  206,  etc. 

(3)  Here  is  an  instructive  differential  equation,  which  Harcourt  and  Esson 
encountered  during  their  work  on  chemical  dynamics  in  1866. 

y2   dx      y       x 

I  shall  give  a  method  of  solution  in  full,  so  as  to  revise  some  preceding  work. 

The  equation  has  the  same  form  as  Bernoulli's.     Therefore,  substitute 

1     .       dv  1     dy  dv  K 

v  =  - :  i.e.,  -5-  =  — z  •  <r*     .*•  -j —  Kv  +  —  =  0, 

y '        '  dx         y2   dx  dx  x 

an  equation  linear  in  v.     The  integrating  factor  is  efpdx\  or,  e~Kx\  Q,  in 
(2),  =  -  E/x ;  therefore,  from  (2), 

ve-Kx  =  _  l—e-xxdx  +  C. 
J  v 
From  the  method  of  §  111,  page  341, 

But  v  =  Ijy.    Multiply  through  with  yeKx,  and  integrate. 

l  =  ^{Cl-logx  +  ^-^f  +  J^-...}y. 

We  shall  require  this  result  on  page  437.     Other  substitutions  may  convert 
an  equation  into  the  linear  form,  for  instance : 

(4)  I  came  across  the  equation  dxjdt  =  k(a  -  x)  (x  -  y),  where  y =a(l  -  e~mt), 
in  studying  some  chemical  reactions.     Put  z  =  l/k(a  -  x);  .».  dx  =  -  dz/kz2, 

.♦.  dzjdt  -kze-"*  =  1. 
This  equation  is  linear.     For  the  integrating  factor  note  that  kedt  =  -  kejm 
=  —ut  say.     Consequently, 

1       [ferdu      „\         e~uL  1    u2      1    v?  „\ 


390  HIGHER  MATHEMATICS.  §  126. 

z  =  0  when  t  =  0 ;  .*.  u  =  k/m  when  £  =  0.  Let  S  denote  the  sum  of  the 
series  when  k/m  is  substituted  in  place  of  u,  and  s  the  sum  as  it  stands  above. 
When  z  =  0,  t  =  0,  C  =  -  S. 

_e~u{8  -  s)      8  -  s  _  m&* 

•'•  *  ~         m         ~  ~^~  ;  •''  U  "  a  ~  k(8  -  s)' 
if  u  is  less  than  unity  the  series  is  convergent. 

(5)  J.  W.  Mellor  and  L.  Bradshaw  {Zeit.  phys.  Ohem.,  $8,  353,  1904)  have 
for  the  hydrolysis  of  cane  sugar 

dufdt  +  bu  =  Ab(l  -  e-fe) ;  .-.  U&*  =  Ae*  -  belb-*)ftb  -  k)  +  C. 

(6)  The  law  of  cooling  of  the  sun  has  been  represented  by  the  equation 
dT/dt  =  aT3  -  bT.  To  solve,  divide  by  Tz ;  put  T  - 2  =  z,  and  hence 
T  ~  zdT  =  -  \dz ;  hence  dzjdt  -  2bz  =  -  2a.  This  is  an  ordinary  linear 
equation  with  the  solution  z  =  ajb  +  Ce2it.  Restore  the  value  of  z.  The 
constant  0  can  be  evaluated  in  the  usual  manner. 

§  126.  Differential  Equations  of  the  First  Order  and  of  the 
First  or  Higher  Degree. — Solution  by  Differentiation. 

Case  i.  The  equation  can  be  split  up  into  factors.  If  the 
differential  equation  can  be  resolved  into  n  factors  of  the  first 
degree,  equate  each  factor  to  zero  and  solve  each  of  the  n  equa- 
tions separately.  The  n  solutions  may  be  left  either  distinct,  or 
combined  into  one. 

Examples. — (1)  Solve  x{dy\dxf=y.  Resolve  into  factors  of  the  first  degree, 
dxjdy=  ±  sjyjx.  Separate  the  variables  and  integrate,  jx~idx±jy~idy=  ±  ,JG, 
where  JO  is  the  integration  constant.  Hence  *Jx  ±  >Jy  =  ±  >JC,  which,  on 
rationalization,  becomes  (x  -  y)2  -  2C{x  +  y)  +  C2  =  0.  Geometrically  this 
equation  represents  a  system  of  parabolic  curves  each  of  which  touches  the 
axis  at  a  distance  C  from  the  origin.  The  separate  equations  of  the  above 
solution  merely  represent  different  branches  of  the  same  parabola. 

(2)  Solve  xy(dyldx)2-(x2-y2)dyldx-xy=0.  Ansr.  xy=C,  or  x2~y2=C. 
Hint.  Factors  (xp  +  y)  (yp  -  z),  where  p  =  dyjdx.  Either  xp  +  y  =  0,  or 
vp  -  x  =  0,  etc. 

(3)  Solve  (dyjdx)2  -  ldy\dx  +  12  =  0.     Ansr.  y  =  ix  +  C,  or  Sx  +  O. 

Case  ii.  The  equation  cannot  be  resolved  into  factors,  but  it  can 
be  solved  for  x,  y,  dyjdx,  or  y/x.  An  equation  which  cannot  be 
resolved  into  factors,  can  often  be  expressed  in  terms  of  x,  y,  dyjdx, 
or  y/x,  according  to  circumstances.  The  differential  coefficient  of 
the  one  variable  with  respect  to  the  other  may  be  then  obtained 
by  solving  for  dyjdx  and  using  the  result  to  eliminate  dyjdx  from 
the  given  equation. 

Examples.— (1)  Solve  dyjdx  +  2xy  =  x2  +  y2.     Since  (x  -  y)2 =x2-  2xy + y* 


§  127.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        391 

y  =  x  +  Jdy/dx.    Put  p  in  place  of  dy/dx.     Differentiate,  and  we  get 

dy  _  1      dp 

dx~     +  2  ijp~ '  dx' 
Separate  the  variables  x  and  p,  solve  for  dy/dx,  and  integrate  by  the  method  of 
partial  fractions. 

.  "am  dP  1,      slv-l  fdlj      C  +  e2x 

•  • dx  -  I^T^-i)  •  •••  •  -  a  '<*  ^+  log  C ;  VJ  =  7TT* 

On  eliminating  2?  by  means  of  the  relation  y  =  cc  +  \/p,  we  get  the  answer 
y  =  x  +  (C  +  e**)j(G  -  e2*). 

(2)  Solve  x(dyfdx)2  -  2y(dy/dx)  +  ax  =  0.  Ansr.  y  =  $(Cx2  +  aje).  Hint. 
Substitute  for  p.  Solve  for  y  and  differentiate.  Substitute  pdx  for  dy,  and 
clear  of  fractions.  The  variables  p  and  a;  can  be  separated.  Integrate 
p  =  xC.     Substitute  in  the  given  equation  for  the  answer. 

(3)  Solve  y  (dy/dx)2  +  2x(dy/dx)  -  y  =  0.  Ansr.  y2  =  C(2x  +  C).  Hint. 
Solve  for  x.  Differentiate  and  substitute  dy\p  for  dx,  and  proceed  as  in 
example  (2).    yp  =  C,  etc. 

Case  iii.  The  equation  cannot  be  resolved  into  factors,  x  or  y  is 
absent.  If  x  is  absent  solve  for  dy/dx  or  y  according  to  conveni- 
ence ;  if  y  is  absent,  solve  for  dx/dy  or  x.  Differentiate  the  result 
with  respect  to  the  absent  letter  if  necessary  and  solve  in  the 
regular  way. 

Examples.— (1)  Solve  (dy/dx)2  +  x(dy/dx)  +  1=0.  For  the  sake  of  greater 
ease,  substitute  p  for  dy/dx.    The  given 'equation  thus  reduces  to 

-  x  =  p  +  1/p (1) 

Differentiate  with  regard  to  the  absent  letter  y,  thus, 

Combining  (1)  and  (2),  we  get  the  required  solution. 

(2)  Solve  dy/dx  =  y  +  1/y.    Ansr.  y2  =  Ce2*  -  1. 

(3)  Solve  dy/dx  =  x  +  1/x.    Ansr.  y  =  %x2  +  log  x  +  G. 

§  127.    Clairaut's  Equation. 

The  general  type  of  this  equation  is 

*  =  *!+/(!);     •    •    •     « 

or,  writing  dy/dx  =  p,  for  the  sake  of  convenience, 

y  =  px  +  f(p).  ...         (2) 

Many  equations  of  the  first  degree  in  x  and  y  can  be  reduced 
to  this  form  by  a  more  or  less  obvious  transformation  of  the  vari- 
ables, and  solved  in  the  following  way:  Differentiate  (2)  with 
respect  to  x,  and  equate  the  result  to  zero 

i>-j.V^+/'(P)|;«,{a+/Mi-o. 


392  HIGHEE  MATHEMATICS.  §  128. 

Hence,  either  dp/dx  =  0 ;  or,  a?  +f'(p)  =  0.     If  the  former, 

where  C  is  an  arbitrary  constant.      Hence,  dy  =  Cdx,  and  the 
solution  of  the  given  equation  is 

y=Cx+f(G). 
Again,  p  in  x  +f'(p)  may  be  a  solution  of  the  given  equation. 
To  find  p,  eliminate  p  between 

y  =px  +  f(p),  and  x  +  f'(p)  =  0. 
The  resulting  equation  between  x  and  y  also  satisfies  the  given 
equation.     There  are  thus  two  classes  of  solutions  to  Clairaut's 
equation. 

Examples.— (1)  Find  both  solutions  in  y  =  px  +  p2.     Ansr.  Cx  +  C2  =  y  ; 
and  x2  +  Ay  =  0. 

(2)  If  (y  -  px)  (p  -l)=p;  show  (y  -  Cx)  (C  -  1)  =  C ;  Jy  +  Jx  =  1. 

(3)  In  the  velocity  equation,  Ex.  (6),  page  388,  if  K=  2,  put  dyjdx  =  p, 
solve  for  y,  and  differentiate  the  resulting  equation, 

_  a  -  x      dy  p      a  -  x   dp       dx  dp 

V-x-      2    P>  dx~p  =  1  +  2  2~'dlc;  'a~^x~  =  ~p~^Z 

Integrate,  and  -  log(a  -  x)  =  -  log(p  -  2) ;  .-.  a  -  x  =  p  -  2,  and  we  obtain 

y  =  2x  -  a  -  (a  -  x)2,  which  is  the  equation  of  a  parabola  yl  =  x^>  if  we 

substitute  x  =  a  +  1  +  x1;  y  =  a  +  1  -  y^ 

After  working  out  the  above  examples,  read  over  §  67,  page  182. 

§  128.   Singular  Solutions. 

Clairaut's  equation  introduces  a  new  idea.  Hitherto  we  have 
assumed  that  whenever  a  function  of  x  and  y  satisfies  an  equation, 
that  function,  plus  an  arbitrary  constant,  represents  the  complete 
or  general  solution.  We  now  find  that  a  function  of  x  and  y  can 
sometimes  be  found  to  satisfy  the  given  equation,  which,  unlike 
the  particular  solution,  is  not  included  in  the  general  solution. 
This  function  must  be  considered  a  solution,  because  it  satisfies 
the  given  equation.  But  the  existence  of  such  a  solution  is  quite 
an  accidental  property  confined  to  special  equations,  hence  their 
cognomen,  singular  solutions.  Take  the  equation 
dy  a  a 

y  =  d^  +  ^;oi>y=px  +  p-    •    •    « 

dx 
Eemembering  that^)  has  been  written  in  place  of  dy/dx,  differentiate 
with  respect  to  x,  we  get,  on  rearranging  terms, 


§  128.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        393 


(•-?) 


dp 
dx 


where  either  x  -  a/p2  =  0 ;  or,  dp/dx  =  0.     If  the  latter, 

p=  G;  or,  y  =  Gx  +  a/C;  .        .        (2) 

and  if  the  former,  p  =  Jajx,  which,  when  substituted  in  (1),  gives 
the  solution, 

y2  =  ±ax (3) 

This  is  not  included  in  the  general  solution,  but  yet  it  satisfies 

the  given  equation.      Hence,  (3)  is  the  singular  solution  of  (1). 

Equation  (2),  the  complete  solution  of   (1),  has   been  shown  to 

represent  a  system  of  straight  lines  which  differ  only  in  the  value 

of  the  arbitrary  constant  G ;  equation  (3),  containing  no  arbitrary 

constant,  is  an  equation  to  the  common  parabola.     A  point  moving 

on  this  parabola  has,  at  any  instant,  the  same  value  of  dy/dx  as  if 

it  were  moving  on  the  tangent  of  the  parabola,  or  on  one  of  the 

straight  lines  of  equation  (2).     The  singular  solution  of  a  differential 

equation  is  geometrically  equivalent  to  the  envelope  of  the  family  of 

curves  represented  by  the  general  solution.     The  singular  solution 

is  distinguished  from  the  particular  solution,  in  that  the  latter  is 

contained  in  the  general  solution,  the  former  is  not. 

'  Again  referring  to  Fig.  96,  it  will  be  noticed  that  for  any  point 

on  the  envelope,  there  are  two  equal  values  of  p  or  dy/dx,  one  for 

the  parabola,  one  for  the  straight  line. 

In  order  that  the  quadratic 

ax2  +  bx  +  c  =  0, 

may  have  equal  roots,  it  is  necessary  (page  354)  that 

b2  =  ±ac ;  or,  b2  -  4ac  =  0.  .         .         .         (4) 

This  relation  is  called  the  discriminant.     From  (1),  since 

a 
y  =  px  +  -;  .\  xp2  -  yp  +  a  =  0.     .        .        (5) 

Ji 

In  order  that  equation  (5)  may  have  equal  roots, 

y2  =  iax, 
as  in  (4).     This  relation  is  the  locus  of  all  points  for  which  two 
values  of  p  become  equal,  hence  it  is  called  the  p-discriminant 
of  (1). 

In  the  same  way  if  G  be  regarded  as  variable  in  the  general 
solution  (2), 

y  =  Gx  +  -p ;  or,  xC2  -  yC  +  a  =  0. 


394 


HIGHER  MATHEMATICS. 


§128. 


The  condition  for  equal  roots,  is  that 

y1  =  4:ax, 

which  is  the  locus  of  all  points  for  which  the  value  of  C  is  the 
same.     It  is  called  the  C-discriminant. 

Before  applying  these  ideas  to  special  cases,  we  may  note  that 
the  envelope  locus  may  be  a  single  curve  (Fig.  96)  or  several 
(Fig.  97).  For  an  exhaustive  discussion  of  the  properties  of  these 
discriminant  relations,  I  must  refer  the  reader  to  the  text-books 
on  the  subject,  or  to  M.  J.  M.  Hill,  "  On  the  Locus  of  Singular 
Points  and  Lines,"  Phil.  Trans.,  1892.     To  summarize: 

1.  The  envelope  locus  satisfies  the  original  equation  but  is 
not  included  in  the  general  solution  (see  xx\  Fig.  146). 


S  /hodalfocusj 


Fig.  146.— Nodal  and  Tac  Loci. 

2.  The  tac  locus  is  the  locus  passing  through  the  several 
points  where  two  non-consecutive  members  of  a  family  of  curves 
touch.  Such  a  locus  is  represented  by  the  lines  AB  (Fig.  97),  PQ 
(Fig.  146).  The  tac  locus  does  not  satisfy  the  original  equation, 
it  appears  in  the  ^-discriminant,  but  not  in  the  O-discriminant. 

3.  The  node  locus  is  the  locus  passing  through  the  different 
points  where  each  curve  of  a  given  family  crosses  itself  (the  point 
of  intersection — node — may  be  double,  triple,  etc.).  The  node 
locus  does  not  satisfy  the   original  equation,  it   appears  in  the 

C-discriminant  but  not  in 
the  ^-discriminant.  BS 
(Fig.  147)  is  a  nodal  locus 
passing  through  the  nodes 
A,...,B,...,  C,...,4f.«. 
4.  The  cusp  locus 
passes  through  all  the 
cusps  (page  169)   formed 


y 


Fig.  147.— Cusp  Locus. 


§  128.      HOW  TO  SOLVE  DIFFEKENTIAL  EQUATIONS.     395 

by  the  members  of  a  family  of  curves.  The  cusp  locus  does  not 
satisfy  the  original  equation,  it  appears  in  the  p-  and  in  the  G- 
discriminants.  It  is  the  line  Ox  in  Fig.  147.  Sometimes  the 
nodal  or  cusp  loci  coincide  with  the  envelope  locus. 

Examples. — (1)  Find  the  singular  solutions  and  the  nature  of  the  other 
loci  in  the  following  equations :  (1)  xp2  -  2yp  +  ax  =  0.  For  equal  roots 
y*  _.  ax%m  TtuS  satisfies  the  original  equation  and  is  not  included  in  the  general 
solution :  x2  -  2Cy  +  C2  =  0.     y2  =  ax2  is  thus  the  singular  solution. 

(2)  ixp2  =  (3»  -  a)2.  General  solution  :  (y  +  Of  =  x(x  -  a)2.  For  equal 
roots  in  p,  4#(3x  -  a)2  =  0,  or  8(38  -  a)2  =  0  (p-discriminant).  For  equal 
roots  in  C,  differentiate  the  general  solution  with  respect  to  0.  Therefore 
(8  +  C)dxjdG  =  0,  or  C  =  -  x.  .-.  x(x  -  a)2  =  0  (C-discriminant)  is  the  con- 
dition to  be  fulfilled  when  the  O-discriminant  has  e  qual  roots,  x  =  0  is 
common  to  the  two  discriminants  and  satisfies  the  original  equation  (singular 
solution) ;  x  =  a  satisfies  the  G-discriminant  but  not  the  ^-discriminant  and, 
since  it  is  not  a  solution  of  the  original  equation,  x  =  a  represents  the  node 
locus ;  8  =  \a  satisfies  the  p-  but  not  the  C-discriminant  nor  the  original 
equation  (tac  locus). 

(3)  p2  +  2xp  -  y  =  0.  General  solution :  (283  +  3xy  +  C)2  =  4(82  +  yf  ; 
p-discriminant :  82  +  y  —  0;  C-discriminant :  (xl  4-  y)3  =  0.  The  original 
equation  is  not  satisfied  by  either  of  these  equations  and,  therefore,  there  is 
no  singular  solution.  Since  (82  +  y)  appears  in  both  discriminants,  it  repre- 
sents a  cusp  locus. 

(4)  Show  that  the  complete  solution  of  the  equation  y2(p2  +  1)  =  a2,  is 
y2  +  (x  -  C)2  =  a1 ;  that  there  are  two  singular  solutions,  y  =  +  a ;  that 
there  is  a  tao  locus  on  the  8-axis  for  y  =  0  (Fig.  97,  page  183). 

A  trajectory  is  a  curve  which  cuts  another  system  of  curves 
at  a  constant  angle.  If  this  angle  is  90°  the  curve  is  an  orthog- 
onal trajectory. 

Examples. — (1)  Let  xy  =  C  be  a  system  of  rectangular  hyperbolas,  to 
find  the  orthogonal  trajectory,  first  eliminate  C  by  differentiation  with  respect 
to  x,  thus  we  obtain,  xdy/dx  +  y  =  0.  If  two  curves  are  at  right  angles 
(£tt  =  90°),  then  from  (17),  §  32,  %n  =  (o'  -  o),  where  a,  a'  are  the  angles 
made  by  tangents  to  the  curves  at  the  point  of  intersection  with  the  8-axis. 
But  by  the  same  formula,  tan(+  Jir)  =  (tana'  -  tana)/(l  +  tana,  tan  a'). 
Now  tan  +  Jir  =  oo  and  l/oo  =0,  .*.  tan  a  =  -  cot  a;  or,  dyjdx  =  -  dx/dy. 
The  differential  equation  of  the  one  family  is  obtained  from  that  of  the  other 
by  substituting  dy/dx  for  -  dx/dy.  Hence  the  equation  to  the  orthogonal 
trajectory  of  the  system  of  rectangular  hyperbolas  is,  xdx  +  ydy  =  0,  or 
x2  -  y2  =■  G,  a  system  of  rectangular  hyperbolas  whose  axes  coincide  with 
the  asymptotes  of  the  given  system.  For  polar  coordinates  it  would  have 
been  necessary  to  substitute  -  (drjr)dd  for  rdejdr. 

(2)  Show  that  the  orthogonal  trajectories  of  the  equipotential  curves 
1/r  -  1//  =  O,  are  the  magnetic  curves  cos  d  +  cos  0'  =  C. 


396  HIGHER  MATHEMATICS.  §  130. 

§  129.    Symbols  of  Operation. 

It  has  been  found  convenient,  page  68,  to  represent  the 
symbol  of  the  operation  u  d/dx  "  by  the  letter  "D  ".  If  we  assume 
that  the  infinitesimal  increments  of  the  independent  variable  dx 
have  the  same  magnitude,  whatever  be  the  value  of  x,  we  can 
suppose  D  to  have  a  constant  value.     Thus, 

respectively.  The  operations  denoted  by  the  symbols  D,  D2,  . .  ., 
satisfy  the  elementary  rules  of  algebra  except  that  they  are  not 
commutative2  with  regard  to  the  variables.  For  example,  we 
cannot  write  D{xy)  =  D(yx).     But  the  index  law 

DmDnu  =  Dm  +  nu 
is  true  when  m  and  n  are  positive  integers.     It  also  follows  that  if 

Du  =  v\  u  =  D~h)',  or,  u  =  j=jv;  .*.  v  =  D .D~lv\  or,  D .D~1  =  l ; 

that  is  to  say,  by  operating  with  D  upon  D  *  lv  we  annul  the  effect 
of  the  D  ~ 1  operator.     In  this  notation,  the  equation 

g_(a  +  /})g  +  a/32/  =  o, 

can  be  written, 

{£>2  -  (a  +   fS)D  +  a/3}y  =  0;   or,  (D  -  a)  (D  -  /%  =  0. 

Now  replace  D  with  the  original  symbol,  and  operate  on  one 
factor  with  y}  and  we  get 

(ii  ~a)(i-  v)y  -°i&  - a)  (I  -  f*v)  =  °- 

By  operating  on  the  second  factor  with  the  first,  we  get  the  original 
equation  back  again. 

§  130.    Equations  of  Oscillatory  Motion. 

By  Newton's  second  law,  if  a  certain  mass,  m,  of  matter  is 
subject  to  the  action  of  an  "  elastic  force,"  FQ,  for  a  certain  time, 
we  have,  in  rational  units, 

Fq  =  Mass  X   Acceleration  of  the  particle. 
If  the  motion  of  the  particle  is  subject  to  friction,  we  may  regard 
the  friction  as  a  force  tending  to  oppose  the  motion  generated  by 

1  See  footnote,  page  177. 


§  130.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        397 

the  elastic  force.  Assume  that  this  force  is  proportional  to  the 
velocity,  V,  of  the  motion  of  the  particle,  and  equal  to  the  product 
of  the  velocity  and  a  constant  called  the  coefficient  of  friction, 
written,  say,  p.  Let  Fx  denote  the  total  force  acting  on  the  par- 
ticle in  the  direction  of  its  motion, 

Fx  =  F0  -  fiV  -  md2s/dt2.  ...         (1) 

If  there  is  no  friction,  we  have,  for  unit  mass, 

FQ  =  d2s/dt2 (2) 

The  motion  of  a  pendulum  in  a  medium  which  offers  no  resist- 
ance to  its  motion,  is  that  of  a  material  particle  under  the  influence 
of  a  central  force,  F,  attracting  with  an  intensity  which  is  pro- 
portional to  the  distance  of  the  particle  away  from  the  centre  of 
attraction.  We  shall  call  F  the  effective  force  since  this  is  the 
force  which  is  effective  in  producing  motion.     Consequently, 

F  =  -q2s,  .  .  •  .  (3) 
where  q2  is  to  be  regarded  as  a  positive  constant  which  tends  to 
restore  the  particle  to  a  position  of  equilibrium — the  so-called  co- 
efficient of  restitution.  It  is  written  in  the  form  of  a  power  to 
avoid  a  root  sign  later  on.  The  negative  sign  shows  that  the 
attracting  force,  F}  tends  to  diminish  the  distance,  s,  of  the  particle 
away  from  the  centre  of  attraction.  If  s  =»  1,  q2  represents  the 
magnitude  of  the  attracting  force  unit  distance  away.     Prom  (2), 

therefore, 

d2s 

w  =  -^s w 

The  integration  of  this  equation  will  teach  us  how  the  particle 
moves  under  the  influence  of  the  force  F.  We  cannot  solve  the 
equation  in  a  direct  manner,  but  if  we  multiply  by  2ds/dt  we  can 
integrate  term  by  term  with:  regard  to  s  ;  thus, 

„ds  d2s       ~  -  ds       ~  fds\2  _  „       _ 

Let  us  replace  the  constant  G  by  the  constant  q2r2 ;  separate  the 
variables,  and  integrate  again ;  we  get  from  Table  II.,  page  193, 

Jds  f  s 

i  2  _  g2  =  ±  q\dt  J  or,  sin  -  !-  =  +  qt  +  c;  or,  s  =  +  r  sin  (qt  +  c), 

where  €  is  a  new  integration  constant.  Here  we  have  s  as  an 
explicit  function  of  t.  We  have  discussed  this  equation  in  an 
earlier  chapter,  pages  66  and  138.  It  is,  in  fact,  the  typical  equa- 
tion of  an  oscillatory  motion.     The  particle  moves  to  and  fro  on  a 


398  HIGHER  MATHEMATICS.  §  130. 

straight  line.  The  value  of  the  sine  function  changes  with  time 
between  the  limits  +  1  and  -  1,  and  consequently  x  changes 
between  the  limits  +  r  and  -  r.  Hence,  r  is  the  amplitude  of  the 
swing  ;  c  is  the  phase  constant  or  epoch  of  page  138.  The  sine  of 
an  angle  always  repeats  itself  when  the  angle  is  increased  by  2ir, 
or  some  multiple  of  2?r.  Let  the  time  t  be  so  chosen  that  after  the 
elapse  of  an  interval  of  time  T0  the  particle  is  passing  through  the 
same  position  with  the  same  velocity  in  the  same  direction, 
hence, 

qT0  =  27r;  or,  T0  =  j.     .         .        .         (5) 

The  two  undetermined  constants  r  and  e  serve  to  adapt  the  relation 

s  =  rsin(qt  +  c) 
to  the  initial  conditions.     This  is  easily  seen  if  we  expand  the 
latter  as  indicated  in  (23)  and  (24),  page  612 : 

s  =  r  sin  € .  cos  qt  ±  r  cos  € .  sin  qt. 
Let   G1  and  C2  denote  the  undetermined  constants  r  sin  e,  and 
r  cos  c  respectively,  such  that 

as  indicated  on  page  138.     Now  differentiate 

ds 
s  =  G^osqt  +  C2sinqt;  .-.  -tt  =  -  qC^inqt  +  qC2cosqt. 

Let  s0  denote  the  position  of  the  particle  at  the  time  t  =  0  when 
moving  with  a  velocity  V0.  The  sine  function  vanishes,  and  the 
cosine  function  becomes  unity.  Hence,  Gx  =  s0 ;  G2  =  V0/q,  and 
the  constants  r  and  e  may  be  represented  in  terms  of  the  initial 
conditions : 

In  the  sine  galvanometer,  the  restitutional  force  tending  to 
restore  the  needle  to  a  position  of  equilibrium,  is  proportional  to 
the  sine  of  the  angle  of  deflection  of  the  needle.  If  /  denotes  the 
moment  of  inertia  of  the  magnetic  needle  and  G  the  directive 
force  exerted  by  the  current  on  the  magnet,  the  equation  of  motion 
of  the  magnet,  when  there  is  no  other  retarding  force,  is 

^--ffBin*.  ...        (6) 

For  small  angles  of  displacement,  <j>  and  sin  <f>  are  approximately 


§  131.  HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.  399 
equal.     Hence, 

^--^  m 

dp  ~      J* Vi 

From  (4),  q  =  s/G/T,  and,  therefore,  from  (5), 

T0  =  %rJ7Ja,    ....  (8) 

a  well-known  relation  showing  that  the  period  of  oscillation  of  a 
magnet  in  the  magnetic  field,  when  there  is  no  damping  action 
exerted  on  the  magnet,  is  proportional  to  the  square  root  of  the 
moment  of  inertia  of  the  magnetic  needle,  and  inversely  proportional 
to  the  square  root  of  the  directive  force  exerted  by  the  current  on  the 
magnet, 

§  131.    The  Linear  Equation  of  the  Second  Order. 

As  a  general  rule  it  is  more  difficult  to  solve  differential  equa- 
tions of  higher  orders  than  the  first.  Of  these,  the  linear  equation 
is  the  most  important.  A  linear  equation  of  the  nth  order  is 
one  in  which  the  dependent  variable  and  its  n  derivatives  are  all 
of  the  first  degree  and  are  not  multiplied  together.  If  a  higher 
power  appears  the  equation  is  not  linear,  and  its  solution  is,  in 
general,  more  difficult  to  find.  The  typical  form  is 
dny       -r>dn-ly  ~         _ 

Or,  in  our  new  symbolic  notation, 

D»y  +  X1D»-iy  +  ...  +  X^y  -  X, 
where  P,  Q,. . .,  B  are  either  constant  magnitudes,  or  functions  of 
the  independent  variable  x.  If  the  coefficient  of  the  highest 
derivative  be  other  than  unity,  the  other  terms  of  the  equation 
can  be  divided  by  this  coefficient.  The  equation  will  thus  assume 
the  typical  form  (1). 

I.  Linear  equations  with  constant  coefficients. 

Let  us  first  study  the  typical  linear  equation  of  the  second 
order  with  constant  coefficients  P  and  Q, 

g+pg+fl*-0.       .         .         .         (1) 

The  linear  equation  has  some  special  properties  which  consider- 
ably shorten  the  search  for  the  general  solution.  For  example, 
let  us  substitute  e™  for  y  in  (1).  By  differentiation  of  e™*,  we 
obtain  dx/dt  »  mx  ;  and  d2x/dt2  =  m2x,  therefore, 


400  HIGHER  MATHEMATICS.  §  131. 

d2pmx  d,pmx 

~aW  +  P~dx~  +  QemX  =  ^  +  Pm  +  ®emx  "  °» 
provided 

m2  +  Pm  +  Q  =  0.         .        .        .         (2) 

This  is  called  the  auxiliary  equation.     If  m1  be  one  value  of  m 

which  satisfies  (2),  then   y  —  emix  is   one   integral   of   (1),    and 

y  =  em2x  is  another.     But  we  must  go  further. 

If  we  know  two  or  more  solutions  of  a  linear  equation,  each 

can  be  multiplied  by  a  constant,  and  their  sum  is  an  integral  of 

the  given  equation.     For  example,  if  u  and  v  are  solutions  of  the 

equation 

d2x 

jp=-q*x,.         .         .         .         (3) 

each  is  called  a  particular  integral,  and  we  can  substitute  either 
u  or  v  in  place  of  x  and  so  obtain 

d2u  d2v 

W2=-q*u;oT,-^=-q*v.  .         .         (4) 

Multiply  each  equation  by  arbitrary  constants,  say,  C1  and  C2 ;  add 
the  two  results  together,  and  C^u,  +  G2v  satisfies  equation  (1), 

.•■^y^-^q,^^).  ■      .      (5) 

This  is  a  very  valuable  property  of  the  linear  equation.  It  means 
that  if  u  and  v  are  two  solutions  of  (3),  then  the  sum  Gxu  +  C2v 
is  also  a  solution  of  the  given  equation.  Since  the  given  equation 
is  of  the  second  order,  and  the  solution  contains  two  arbitrary  con- 
stants, the  equation  is  completely  solved.  The  principle  of  the 
superposition  of  particular  integrals  here  outlined  is  a  mathe- 
matical expression  of  the  well-known  physical  phenomena  discussed 
on  page  70,  namely,  the  principle  of  the  coexistence  of  different 
reactions ;  the  composition  of  velocities  and  forces  ;  the  super- 
position of  small  impulses,  etc.  We  shall  employ  this  principle 
later  on,  meanwhile  let  us  return  to  the  auxiliary  equation. 

1.  When  the  auxiliary  equation  has  two  unequal  roots,  say  m1 
and  m2,  the  general  solution  of  (1)  may  be  written  down  without 
any  further  trouble. 

y  =  C^i*  +  G2em2x.        .         .         .         (6) 

This  result  enables  us  to  write  down  the  solution  of  a  linear  equa- 
tion at  sight  when  the  auxiliary  has  unequal  roots. 


§  131.       HOW  TO  SOLVE  DIFFERENTIAL  EQUATONS.        401 

Examples.— (1)  Solve  (D2  +  UD  -  S2)y  =  0.  Assume  y  =  Cemx  is  a 
solution.  The  auxiliary  becomes,  m2  +  14m  -  32  =  0.  The  roots  are  m  =  2 
or  -  16.     The  required  solution  is,  therefore,  y  =  Cxe2*  +  C2e~Wx. 

(2)  Solve  d2yjdx2  +  Adyjdx  +  3y  =  0.     Ansr.  y  =  Op-3"  +  Ctf.~x. 

(3)  Fourier's  equation  for  the  propagation  of  heat  in  a  cylindrical  bar,  is 
d'-Vldx*  -  0>V  =  0.     Hence  show  that  V  =  C^x  +  Ce~  ?x. 

2.  When  the  two  roots  of  the  auxiliary  are  equal.  If  mx  =  ra2, 
in  (6),  it  is  no  good  putting  (G1  +  G2)emi*  as  the  solution,  because 
Cx  +  G2  is  really  one  constant.  The  solution  would  then  contain 
one  arbitrary  constant  less  than  is  required  for  the  general  solu- 
tion. We  can  find  the  other  particular  integral  by  substituting 
m2  =  m1  +  h,  in  (6),  where  h  is  some  finite  quantity  which  is  to  be 
ultimately  made  equal  to  zero.  Substitute  m2  =  m1  +  h  in  (6) ; 
expand  by  Maclaurin's  theorem,  and,  at  the  limit,  when  h  =  0, 
we  have 

y  =  emi*(A  +  Bx).  ...         (7) 

This  enables  us  to  write  down  the  required  solution  at  a  glance. 
For  equations  of  a  higher  order  than  the  second,  the  preceding 
result  must  be  written, 

y  =  e-i^d  +  G2x  +  Cp*  +  . . .  +  Cr  _  &  ~ l),     .         (8) 
where  r  denotes  the  number  of  equal  roots. 

Examples. — (1)   Solve  d^y/dx3  -  dPyjdx2  -  dy/dx  +  y  =  0.      Assume 
y  =  Ge™*.     The  auxiliary  equation  is  m3  -  m2  -  m  +  1  =  0.     The  roots  are 
1,  1,  -  1.     Hence  the  general  solution  can  be  written  down  at  sight : 
y  =  Cxe-*  +  (C2  +  C3x)e*. 

(2)  Solve  (D3  +  3D2  -  4)y  =  0.  Ansr.  e~2*(C1  +  C.2x)  +  C#*.  Hint. 
The  roots  are  obtained  from  (x  -  2)  (x  -  2)  (x  -  1)  =  x3  +  Sx2  -  4. 

3.  When  the  auxiliary  equation  has  imaginary  roots,  all  un- 
equal. Eemembering  that  imaginary  roots  are  always  found  in 
pairs  in  equations  with  real  coefficients,  let  the  two  imaginary 
roots  be 

ml  =  a  +  i/?  J  and  m2  =  a  -  i/?. 
Instead  of  substituting  y  =  e™*  in  (6),  we  substitute  these  values 
of  m  in  (6)  and  get 

y  =  <7ie(a  +  t0)a;  +   Q^a-iftx  _  qox^Q^x  +   C2e~^x)', 

where  Gx  and  G2  are  the  integration  constants.  From  (13)  and 
(15),  p.  286, 

y  =  e^O^cos  fix  +  i  sin  fix)  +  eaa;C2(cos  fix  -  t  sin  (Sx).       (9) 
Separate  the  real  and  imaginary  parts,  as  in  Ex.  (3),  p.  351, 

.-.  y  =eax{(G1  +  C2)cos/to  +  i(0l  -  C2)sin/ta}, 
CO 


402  HIGHER  MATHEMATICS.  §  131. 

If  we  put  Gl  +  C2  =  A,  and  l(C1  -  G2)  =  B,  we  can  write  down 
the  real  form  of  the  solution  of  a  linear  equation  at  sight  when  its 
auxiliary  has  unequal  imaginary  roots. 

y  =  e°*(A  cos  fix  +  B  sin  fix).  .  .  (10) 
In  order  that  the  constants  A  and  B  in  (10)  may  be  real,  the 
constants  C1  and  G2  must  include  the  imaginary  parts. 

The  undetermined  constants  A  and  B  combined  with  the  par- 
ticular integrals  u  and  v  may  be  imaginary.  Thus,  u  and  v  may 
be  united  with  Oj  and  iClf  and  Au  +  iBv  is  then  an  integral  of  the 
same  equation.  It  is  often  easier  to  find  a  complex  solution  of 
this  character  than  a  real  expression.  If  we  can  find  an  integral 
u  +  iv,  of  the  given  equation,  u  and  v  can  each  be  separately 
regarded  as  particular  integrals  of  the  given  equation. 

Examples. — (1)  Show,  from  (9),  and  (2)  and  (3)  of  page  347,  that  we  can 
write  y  =  (cosh  ax  +  sinh  ax)  (A-^  cos  fix  +  Bx  sin  fix). 

(2)  Integrate  (Py/dx2  +  dyfdx  +  y=0.  The  roots  are  a  =  -  J  and  fi = %  V3 ; 
.•.  y  =  e-xP(Acos$\f3.  x  +  B  sin  £  */&  .  x). 

(3)  The  equation  of  a  point  vibrating  under  the  influence  of  a  periodic 
force,  is,  d2xjdt2  +  q2x  =  0,  the  roots  are  given  by  (D  +  ta)  (D  -  to)  =  0.  From 
(10),  y  =  A  cos  ax  +  B  sin  ax. 

(4)  In  the  theory  of  electrodynamics  (Encyc.  Brit.,  28,  61,  1902)  and  in 
the  theory  of  sound,  as  well  as  other  branches  of  physics,  we  have  to  solve 
the  equation 

dr2       r   dr  r 

Multiply  by  r  and  notice  that 

d(r<(>)  _    dr        d<p  _  d<p  d2((pr)  _  d<p      dr    d<p        d2<j> 

~W  ~<pd?  +  rdr'-<t}  +  rdr';  ~W  ~  dr~  +  dr"  dr~  +  rdr2' 

.  JE±  .   od<t>  _  <*2(4>r) 
*  *   dr2  +  *dr      ~W*~' 
Hence  we  may  write 

^^  =  k2(r<f>)  =  (D2  +  k2)<pr  =(D  +  ik)  (D  -  tk)<pr  =  0. 

.*.  <pr  =  Aeikr  +  Be~iJer ;  .'.  <p=  -(A  cosfrr  +  Esin  fcr),asin(9)  and  (10)  above. 

4.  When  some  of  the  imaginary  roots  of  the  auxiliary  equation 
are  equal.  If  a  pair  of  the  imaginary  roots  are  repeated,  we  may 
proceed  as  in  Case  2,  since,  when  mx  =  m2,  GYemlx  +  C2em2x,  is 
replaced  by  (A  +  Bx)emlx ;  similarly,  when  ra3  =  ra4,  G3emZx  +  G^ 
may  be  replaced  by  (G  +  Dx^™**.     If,  therefore, 

mx  =  m2  =  a  +  l/3  ;  and  m3  =  m4  =  a  -  i/?, 
the  solution 

y  =  (ox  +  c2xy«+w*  +  (o8  +  Ctxy—w, 


§  131.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        403 

becomes 

y  =  e«*{(A  +  Bx)  cos  fix  +  (0  +  Dx)  sin  fix}.     .       (11) 

Examples.— (1)  Solve  (Di  -  12D3  +  62D2  -  156D  +  16%  =  0.  Given  the 
roots  of  the  auxiliary :  3  +  2i,  3  +  2i,  3  -  2t,  3  -  2i.  Hence,  the  solution  is 
y  =  eix{{Cx  +  G&)  sin  2x  +  (03  +  C^x)  cos  2x}. 

(2)  If  (D2  +  if(D -i)2y=o,y=(A  +  Bx)  sin  x+{C+ Dx)  cos x  +  ( E  +  Fx)e*. 

II.  Linear  equations  with  variable  coefficients. 

Linear  equations  with  variable  coefficients  can  be  converted 
into  linear  equations  with  constant  coefficients  by  means  of  the 
substitution 

x  =  e' ;  or,  z  =  log  x,        .     -    .        .      (12) 

as  illustrated  in  the  following  example  : — 
Example. — Solve  the  equation 

by  means  of  the  substitution  (12).     By  differentiation  of  (12),  we  obtain 

—  =  e*  -    -  e&  =  ^  •    •  ^  =  -  •  ^ 
dz         '  "    dx      dz'  ' '  dx      x'  dz 

Again, 

d?y  =  ld?y  _dx   dy     _   d?y  _  l_(d?y  _  dy\ 

dx      x  dz       x*'  dz'  '"'  dx%  ~  x\dz*      dz)' 

since,  from  (12),  dx  =  xdz.  Introducing  these  values  of  dyjdx  and  d'hjfdx2  in 
the  given  equation,  we  get  the  ordinary  linear  form 

S  +  4  +  *-* 

with  constant  coefficients.     Hence,  y  =  Gx&*  +  Ctf*  =  Cxx2  +  C^c. 

If  the  equation  has  the  form  of  the  so-called  Legendre's 
equation,  say, 

(a  +  a;)2S'-5(<l  +  a!)l  +  62'  =  0;  '        <13> 

the  substitution  z  =  a  +  x  will  convert  it  into  form  (12),  and  the 
substitution  e*  for  a  +  x  will  convert  it  into  the  linear  equation 
with  constant  coefficients.  Hence,  dx  =  (a  +  x)dt ;  dx2  =  (a  +  x)2dt2, 
and 

%  -  5%  +  6y  -  ° ;  •*•  y = cie* +  °^  -  °i(a  +  *)2 + ^ + *o3. 


Example. — In  the  theory  of  potential  we  meet  with  the  equation 
dW  ,  2    dV      _  2d27      „  <Z7     A 

The  roots  of  the  auxiliary  are  m  =  0,  and  w  =  -  1.    Hence,  7=  Ox  +  C/ 


404  HIGHER  MATHEMATICS.  §  132. 

§  132.    Damped  Oscillations. 

d2s 
The  equation  -r^  =  -  q2s  takes  no  account  of  the  resistance  to 

which  a  particle  is  subjected  as  it  moves  through  such  resisting 
media  as  air,  water,  etc.  We  know  from  experience  that  the 
magnitude  of  the  oscillations  of  all  periodic  motions  gradually 
diminishes  asymptotically  to  a  position  of  rest.  This  change  is 
called  the  damping  of  the  oscillations. 

When  an  electric  current  passes  through  a  galvanometer,  the  needle  is 
deflected  and  begins  to  oscillate  about  a  new  position  of  equilibrium.  In 
order  to  make  the  needle  come  to  rest  quickly,  so  that  the  observations  may 
be  made  quickly,  some  resistance  is  opposed  to  the  free  oscillations  of  the 
needle  either  by  attaching  mica  or  aluminium  vanes  to  the  needle  so  as  to 
increase  the  resistance  of  the  air,  or  by  bringing  a  mass  of  copper  close  to 
the  oscillating  needle.  The  currents  induced  in  the  copper  by  the  motion  of 
the  magnetic  needle,  react  on  the  moving  needle,  according  to  Lenz'  law,  so 
as  to  retard  its  motion.  Such  a  galvanometer  is  said  to  be  damped.  When 
the  damping  is  sufficiently  great  to  prevent  the  needle  oscillating  at  all,  the 
galvanometer  is  said  to  be  "dead  beat"  and  the  motion  of  the  needle  is 
aperiodic.    In  ballistic  galvanometers,  there  is  very  much  damping. 

It  is  a  matter  of  observation  that  the  force  which  exerts  the 
damping  action  is  directed  against  that  of  the  motion  ;  and  it  also 
increases  as  the  velocity  of  the  motion  increases.  The  most 
plausible  assumption  we  can  make  is  that  the  damping  force,  at 
any  instant,  is  directly  proportional  to  the  prevailing  velocity,  and 
has  a  negative  value.  To  allow  for  this,  equation  (4)  must  have 
an  additional  negative  term.  We  thus  get  the  typical  equation  of 
the  second  order, 

d2s  fo        2 

<ft*  "  ~  l*dt~~  q  s' 
where  /a  is  the  coefficient  of  friction.     For  greater  convenience,  we 
may  write  this  2/,  and  we  get 

d2s  ds 

aF  +  a/aj  +  ^-o.     .      .      .      (i) 

Before  proceeding  further,  it  will  perhaps  make  things  plainer 
to  put  the  meaning  of  this  differential  equation  into  words.  The 
manipulation  of  the  equations  so  far  introduced,  involves  little  more 
than  an  application  of  common  algebraic  principles.  Dexterity  in 
solving  comes  by  practice.  Of  even  greater  importance  than  quick 
manipulation  is  the  ability  to  form  a  clear  concept  of  the  physical 
process  symbolized  by  the  differential  equation.     Some  of  the  most 


§  132.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        405 

important  laws  of  Nature  appear  in  the  guise  of  an  "  unassuming 
differential  equation  ".  The  reader  should  spare  no  pains  to  acquire 
familiarity  with  the  art  of  interpretation  ;  otherwise  a  mere  system 
of  differential  equations  maybe  mistaken  for  "laws  of  Nature". 
The  late  Professor  Tait  has  said  that  "  a  mathematical  formula, 
.however  brief  and  elegant,  is  merely  a  step  towards  knowledge,  and 
an  all  but  useless  one  until  we  can  thoroughly  read  its  meaning  ". 

Buler  once  confessed  that  he  often  could  not  get  rid  of  the 
uncomfortable  feeling  that  his  science  in  the  person  of  his  pencil 
surpassed  him  in  intelligence.  I  dare  say  the  beginner  will  have 
some  such  feeling  as  he  works  out  the  meaning  of  the  above 
innocent-looking  expression.  The  term  d2s/dt2,  page  17,  denotes  the 
relative  change  of  the  velocity  of  the  motion  of  the  particle  in  unit 
time  ;  while  2/ .  ds/dt  shows  that  this  motion  is  opposed  by  a  force 
which  tends  to  restore  the  body  to  a  position  of  rest,  the  greater 
the  velocity  of  the  motion,  the  greater  the  retardation  ;  and  q2s  re- 
presents another  force  tending  to  bring  the  moving  body  to  rest, 
this  force  also  increases  directly  as  the  distance  of  the  body  from  the 
position  of  rest.  The  whole  equation  represents  the  mutual  action 
of  these  different  effects.  To  investigate  this  motion  further,  we 
must  find  a  relation  between  s  and  t.  In  other  words,  we  must 
solve  the  equations. 

We  can  write  equation  (1)  in  the  symbolic  form 
(Z>2  +  2/D  +  q*)s  -  0, 
the  roots  of  the  auxiliary  are,  pages  353  and  354, 

*--f+JF^;  P--f~'Jfr^'      •       (2) 

The  solution  of  (1)  thus  depends  on  the  relative  magnitudes  of 
/  and  q.  There  are  two  important  cases  :  the  roots,  a  and  ft,  may 
be  real  or  imaginary.  Both  have  a  physical  meaning  and  represent 
two  essentially  different  types  of  motion.  Suppose  that  we  know 
enough  about  the  moving  system  to  be  able  to  determine  the  in- 
tegration constant.     When  t  —  0,  let  V  =  V0  and  s  =  0. 

I.  The  roots  of  the  auxiliary  equation  are  imaginary,  equal 
and  of  opposite  sign.  For  equal  roots  of  opposite  sign,  say  +  q, 
we  must  have  /  =  0,  as  indicated  upon  page  401.  In  this  case,  as 
in  the  typical  equation  for  Case  3  of  the  preceding  section, 

s  =  GY  sin  qt  +  G2  cos  qt.  .  .  .  (3) 
To  find  what  this  means,  let  us  suppose  that  t  =  0,  s  =  0,  V0  =  1, 
q  =  2,  /  =  0.     Differentiate  (3), 

ds/dt  m  qCx  cos  qt  -  qC2  sin  qt. 


406 


HIGHER  MATHEMATICS. 


§132. 


2  x  C2  x  0,  or  G1  =  J ;  .-.  G2  =  0.     Hence 


Hence  1  =  2Gl  x  1 
the  equation, 

s  =  J  sin  2* (4) 

Curve  1  (Pig.  148)  was  obtained,  by  plotting,  from  equation  (4)  by  assign- 
ing arbitrary  values  to  t  in  radians ;  converting  the  radians  into  degrees  ;  and 
finding  the  sine  of  the  corresponding  angle  from  a  Table  of  Natural  Sines. 
Suppose  we  put  %  =  45°,  then  sine  45°  m  0*79  ;  t  =  22-5°  =  0*39  radians  from 
Table  XIII. ;  if  2t  =  630°,  sin  630°  =  sin  45°  in  the  fourth  quadrant,  it  is  there- 
fore negative  ;  t  =  320° ;  .-.  t  =  (3-1416  +  £  of  3-1416  +  0-39)  radians.  The 
numbers  set  out  in  the  first  three  columns  of  the  following  table  were  calcu- 
lated from  equation  (4)  for  the  first  complete  vibration  : — 


s  =  £  sin  2t. 

,  =  ^-o-i«sin  1-997^ 

t 
radians. 

sin  2t. 

a. 

t. 

sin  V7t. 

e-°-lt. 

«. 

0 

0 

0 

0 

0 

1-00 

0 

0-39 

+  0-79 

+  0-39 

0-46 

+  0-79 

0-96 

+  3-84 

0-78 

+  1-00 

+  0-50 

0-92 

+  i-oo 

0-91 

+  4-55 

1-18 

+  0-79 

+  0-39 

1-38 

+  0-79 

0-87 

+  3-84 

1-57 

0 

0 

1-80 

0 

0-84 

0 

1-96 

-  0-79 

-0-39 

2-30 

-  0-79 

0-79 

-  3-84 

2-44 

-  1-00 

-  0-50 

2-77 

-  100 

0-76 

-  4-55 

2-69 

-  0-79 

-0-39 

3-20 

-  0-79 

0-73 

-  3-84 

3-14 

0 

0 

3-70 

0 

0-69 

0 

II.  The  roots  of  the  auxiliary  equation  are  imaginary.  For 
imaginary  roots,  -  f  ±  Jif2  -  q2),  or,  say  -  a  ±  bi,  it  is  neces- 
sary that  f  <  q  (page  354).     In  this  case, 

s  =  e  ~  "(Ci  sin  bt  +  G2  co3  bt).  .  .  (5) 
Let  the  coefficient  of  friction,/  =  0*1,  q  •—  2,  t  =  0,  s  =  0,  F0  =  1. 
The  roots  of  the  auxiliary  are  m=  -01  ±  VO'01  -  4  =  01±  V-3'99 
=  -  04  +  il-97,  where  i  =  J  -  1.  Hence  a  =  0'1,  b  =  1-997. 
Differentiate  (5), 

ds/dt  =  -  ae-at(G1Qinbt  +  C2  cos  bt)  +  be~  at  (C  l  cos  bt-  C2  sin  bt). 
From  (5),  C2  =  0,  and,  therefore,  Cj  =  1/6  =  0*5.     Therefore, 

s  =  0-56-0'1'  sin  1-997*,  ...  (6) 
a  result  which  differs  from  that  which  holds  for  undamped  oscilla- 
tions by  the  introduction  of  the  factor  e~°'u.  The  last  four 
columns  of  the  above  table  has  the  numbers  computed  for  the 
first  complete  vibration  from  equation  (6).  The  graph  of  the 
equation  is  curve  2  of  Fig.  148. 

The  simple  harmonic  curve,  1,  Fig.  148,  represents  the  un- 


§  132.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        407 

damped  oscillations  of  a  particle.  The  effects  of  damping  are 
brought  out,  by  the  diagram,  curve  2,  in  an  interesting  manner. 
The  net  result  is  a  damped  Yibration,  which  dies  away  at  a  rate 
depending  on  the  resistance  of  the  medium  (2fv)  and  on  the 
magnitude  of  the  oscillations  (q2s).      Such   is   the   motion   of   a 


/X                               ^X*^                         ^V 

J       1$*                  J       /S  V            4             S-=^ 

f  tsX     t/  XX   L   -,X^4 

o   «,N,2r|   -xTph  %    KJX*Z   %">    3 

5\\  /Z         V  2-^      V  l. 

^5-               Kz                ^Z 

Fig.  148. 

magnetic  or  galvanometer  needle  affected  by  the  viscosity  of  the 
air  and  the  electromagnetic  action  of  currents  induced  in  neigh- 
bouring masses  of  metal  by  virtue  of  its  motion ;  it  also  represents 
the  natural  oscillations  of  a  pendulum  swinging  in  a  medium  whose 
resistance  varies  as  the  velocity.  The  effects  of  damping  are  two- 
fold : 

1.  The  period  of  oscillation  is  augmented  by  damping,  from 
T0  to  T.     From  equation  (5),  we  can  show,  page  138,  that 

s  =  e  *  °*A  sin  bt (7) 

The  amplitude  of  this  vibration  corresponds  to  the  value  of  t  for 
which  s  has  a  maximum  or  a  minimum  value.  These  values  are 
obtained  in  the  usual  way,  by  equating  the  first  differential  co- 
efficient to  zero,  hence 

e~at(bcosbt  -  asmbt)  =  0.  .         .         .         (8) 

If  we  now  define  the  angle  <£  such  that  bt  =  </>,  or 

tan<£  =  b/a,  ....  (9) 
<£>  tymg  between  0  and  \ir  {i.e.,  90°),  becomes  smaller  as  a  in- 
creases in  value.  We  have  just  seen  that  the  imaginary  roots  of 
-  /  ±  JP  -  q2  are  -  a  ±  ub,  for  values  of  /  less  than  q.  Conse- 
quently, 
(-/+  sff^i-f-  JfZ¥)  =  (icL  +  ^)(a-Lb);  or,  a*  +  b2  =  q2.  (10) 

The  period  of  oscillation  of  an  undamped  oscillation  is,  by 
(5),  page  398,  T0  =  Q-rr/q,  and  similarly,  for  a  damped  oscillation 
T  =  2nlb 


T2       q2      a2  +  b2               a2 

T 

'T\~  b2~       b2       "       '   b2'  ' 

'To 

n/o2  +   W 


(11) 


408  HIGHER  MATHEMATICS.  §  132. 

which  expresses  the  relation  between  the  periods  of  oscillation 
of  a  damped  and  of  an  undamped  oscillation.  Consequently, 
OT  -  OT0  =  1-019  (Fig.  148). 

2  The  ratio  of  the  amplitude  of  any  vibration  to  the  next,  is 
constant.    The  amplitudes  of  the  undamped  vibrations  M-J?^  M2P2, 
...  become,  on  damping  NxQlt  N2Q2,  ...  It  is  easy  to  show,  by 
plotting,  that  tan  <f>,  of  (9),  is  a  periodic  function  such  that 
tan  <£  =  tan  (<£  +  ?r)  =  tan  (<£  +  %tt)  =  . . . 

Hence  <f> ;  <f>  +  -n- ;   <f>  +  2tt  ;  ...  satisfy  the  above  equation.      It 
also  follows  that  btx ;  bt2  +  Tr;  bt2+  27r;  . . .  also  satisfy  the  equation, 
where  tv  t2,  t3,  ...  are  the  successive  values  of  the  time.     Hence 
bt^b^  +  Tr;  tag  =  6^  +  271-;  ...;  .\  t^^  +  ^T;  t^^  +  T;  ... 
Substitute  these  values  in  (7)  and  put  sv  s2,  s3,  ...  for  the  cor- 
responding displacements, 

.'.s1  =  Ae'^sinb^;  -  s2  =  Ae~  "^siabt^;  ... 
where  the  negative  sign  indicates  that  the  displacement  is  on  the 
negative  side.     Hence  the  amplitude  of  the  oscillations  diminishes 
according  to  the  compound  interest  law, 

h  =  e—«i-<3)  =  eh«r.   £2  =  H  =  ^  >  m  =  e¥x%        %  (12^ 

s2  s3      s4 

This  ratio  must  always  be  a  proper  fraction.  If  a  is  small,  the 
ratio  of  two  consecutive  amplitudes  is  nearly  unity.  The  oscilla- 
tions diminish  as  the  terms  of  a  geometrical  series  with  a  common 

ratio  eaT'2.  By  tak- 
ing logarithms  of  the 
terms  of  a  geometrical 
series  the  resulting 
arithmetical  series  has 
every  succeeding  term 
smaller  than  the  term 
which  precedes  it  by 
a  constant  difference. 
149.— Strongly  Damped  Oscillations.  This  difference  can  be 

found  by  taking  logarithms  of  equations  (12). 

Plotting  these  successive  values  of  s  and  t,  in  (12),  we  get  the 
curve  shown  in  Fig.  149.  The  ratio  of  the  amplitude  of  one  swing 
to  the  next  is  called  the  damping  ratio,  by  Kohlrausch  ("Damp- 
fungsverhaltnis  ").  It  is  written  k.  The  natural  logarithm  of  the 
damping  ratio,  is  Gauss'  logarithmic  decrement,  written  X  (the 


§  132.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        409 


ordinary  logarithm  of  ky  is  written  L).     Hence 

A  =  log  k  =  JaTlog  e  =  \aT  =  air/h, 
and  from  (11), 


T^  X_2 


r.^(i^.S^...). 


(13) 


(14) 


Hence,  if  the  damping  is  small,  the  period  of  oscillation  is  aug- 
mented by  a  small  quantity  of  the  second  order.  The  logarithmic 
decrement  allows  the  "  damping  constant  "  or  "  frictional  co- 
efficient"  (/a  of  pages  397  and  404)  to  be  determined  when  the 
constant  a  and  the  period  of  oscillation  are  known.  It  is  there- 
fore not  necessary  to  wait  until  the  needle  has  settled  down  to 
rest  before  making  an  observation.  The  following  table  contains 
six  observations  of  the  amplitudes  of  a  sequence  of  damped  oscilla- 
tions : 


Observed 
Deflection. 

*. 

A. 

L. 

69 
48 
33-5 
23-5 
16-5 
11-5 
8 

1-438 
1-434 
1-426 
1-425 
1-435 
1-438 

0-3633 
0-3604 
0-3548 
0-3542 
0-3612 
0-3633 

0-1578 
0-1565 
0-1541 
0-1538 
0-1569 
0-1578 

Observations  of  oscillating  pendulums,  vibrating  needles,  etc., 
play  an  important  part  in  the  measurement  of  the  force  charac- 
terized by  the  constant  q,  whether  that  be  the  action  of,  say, 
gravity  on  a  pendulum,  of  a  magnetic  field  on  the  motion  of  a 
magnet.  The  small  oscillations  of  a  pendulum  in  a  viscous 
medium  furnish  numerical  values  of  the  magnitude  of  fluid 
friction  or  viscosity. 

III.  The  roots  of  the  auxiliary  equation  are  real  and  unequal. 
The  condition  for  real  roots  -  a  and  -  /?,  in  (2),  is  that  /  be  greater 
than  q  (page  354).     In  this  case, 

8=C1e~,u  +  C&-f»,  .  .  .  (15) 
solves  equation  (1).  To  find  what  this  means,  let  us  suppose  that 
f  -  3,  q  «  2,  t  -  0,  *  -  0,  Vv  -  1,     From  (2),  therefore, 

m  =  -  3  ±  sj9^~4  =  -  3  ±  2-24  =  -  -76  and  -  5-24. 
Substitute  these  values  in  (15)  and  differentiate  for  the  velocity  v 


410  HIGHER  MATHEMATICS.  §  133. 

or  ds/dt.     Thus 

s  =  C,e-™  +  C2e-™;  ds/dt  =  -  5'MCtf-™  -  0'76C2e-™. 

.-.  -  5-240!  -  -76C2  =  1. 

From  (15),  when  t  =  0,s  =  0;  and  d+C^O;  or  -  Cx=  +  C2  =  0'225, 

.-.  s  =  0-225(e-'r*  -  e"5"24().  .         .         (16) 

Assign  particular  values  to  t,  and  plot  the  corresponding  values 

of  s  by  means  of  Table  IV.,  page  616.     Curve  3  (Fig.  150)  was 


^ 


Fig.  150. 

obtained  by  plotting  corresponding  values  of  s  and  t  obtained  in 
this  way.     The  curves  have  lost  the  sinuous  character,  Fig.  148. 

IV.  The  roots  of  the  auxiliary  equation  are  real  and  equal. 
The  condition  for  real  and  equal  roots  is  that  /  =  q. 

.v*-(G1  +  ^)«^;  .  .  .  (17) 
As  before,  let  /  =  2,  q  =  2,  t  =  0,  s  =  0,  V0  =  1.  The  roots  of  the 
auxiliary  are  -  2  and  -  2.  Hence,  to  evaluate  the  constants, 
s  =  (G1  +  C2t)e -  *  js  ds/dt  =  G2e ~*  -  2(<71  +  G2t)e '  2< ;  C2  -  2G1  =  1 ; 
Gx»  0  and  C2  =  1 ; 

.-.  s  =  £e-2' (18) 

Plot  (18)  in  the  usual  manner.     Curve  4  (Fig.  150)  was  so  ob- 
tained. 

Compare  curves  3  and  4  (Fig.  148)  with  curves  1  and  2 
(Fig.  150).  Curves  3  and  4  (Fig.  150)  represent  the  motion  when 
the  retarding  forces  are  so  great  that  the  vibration  cannot  take 
place.  The  needle,  when  removed  from  its  position  of  equilibrium, 
returns  to  its  position  of  rest  asymptotically,  i.e.,  after  the  lapse  of 
an  infinite  time.  What  does  this  statement  mean  ?  E.  du  Bois 
Raymond  calls  a  movement  of  this  character  an  aperiodic  motion. 

§  133.   Some  Degenerates. 

There  are  some  equations  derived  from  the  general  equation 
by  the  omission  of  one  or  more  terms.  The  dependent  or  the 
independent  variable  may  be  absent.  I  have  already  shown, 
pages    249    and    401,    how    to    solve   equations   of   this    form : 


§  133.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        411 

d2y  <Py  dhi 

&-9:  ■&  +  &-*'.-&- ft-*.      .      (i) 

where  g  and  q  are  constants. 

Example. — The  general  equation  for  the  deflection  of  a  horizontal  beam 
uniformly  loaded  and  subjected  to  the  pressure  of  its  supports  is 
dhj  d4y 

where  a  and  w  are  constants.  If  the  beam  has  a  length  Z  and  is  supported  at 
both  ends,  the  integration  constants  are  evaluated  on  the  supposition  that 
y  =  0,  and  dhjjdx2  =  0  both  when  x  =  0  and  x  =  I.  Hence  show  that  the 
integration  constants,  taken  in  the  same  order  as  they  appear  in  the  inte- 
gration, are  Cx  =  -  %bl ;  C2  =  0 ;  C9  =  -fabl3  ;  C4  =  0.  Hence  the  solution 
V  =  ■<ribx(x3  -  2lx  +  Z3).  If  the  beam  is  clamped  at  both  ends,  y  =  0,  and 
dy/dx  =  0  for  x  =  0  and  x  =  I.  Show  that  the  constants  now  become 
Gx  =  -  Ibl ;  C  =  ,ij&Z2 ;  C3  =  0  ;  C4  =  0.  .-.  y  =  ^bx^x2  -  2lx  +  Z2).  If  the 
beam  is  clamped  at  one  end  and  free  at  the  other,  y  =  0,  dy\dx  =  0  for 
x  =  0 ;  and  dhjjdx2  =  0  and  d^y/dx3  =  0  when  <c  =  Z.  Show  that  C^  =  -  bl ; 
C  =  %bl ;  03  =  0  ;  C4  =  0.     .-.  y  =  to2(x2  -  4Zz  +  6Z2). 

If  hx  denotes  the  "pull "  of  a  spring  balance  when  stretched  a 
distance  x,  at  the  time  t,  the  equation  of  motion  is 
d2x       k 
W  +  m*-9>    ....        (2) 

where  m  denotes  the  stretching  weight ;  g  is  the  familiar  gravita- 
tion constant.  For  the  sake  of  simplicity  put  k/m  =  a2,  and  we 
can  convert  (2)  into  one  of  the  above  forms  by  substituting 

9         d2u        9 
X  =  U  +  i2'"''W+ahl  =  0' 
Solving  this  latter,  as  on  page  401,  we  get 

u  =  Oj  cos  at  +  C2  sin  at ;  .-.  x  =  G2  cos  at  +  C2  sin  at  +  g/a2. 
Or,  you  can  solve  (2)  by  substituting 

„  =  % .    .  ^V  _  dp  _  dy   <fy  __    dp 
v      dx'  "  dx2      dx~  dx'  dy  ~  pdy'        '        W 
so  as  to  convert  the  given  equation  into  a  linear  equation  of  the 
first  order.     For  the  sake  of  ease,  take  the  equation 
dW     1   dV_      P 

dr2  +  r'  dr  ~  l^  '  '  *  ^ 
which  represents  the  motion  of  a  fluid  in  a  cylindrical  tube  of 
radius  r  and  length  I.  The  motion  is  supposed  to  be  parallel  to 
the  axis  of  the  tube  and  the  length  of  the  tube  very  great  in 
comparison  with  its  radius  r.  P  denotes  the  difference  of  the 
pressure  at  the  two  ends  of  the  tube.  If  the  liquid  wets  the 
walls  of  the  tube,  the  velocity  is  a  maximum  at  the  axis  of  the 


412  HIGHER  MATHEMATICS.  §133. 

tube  and  gradually  diminishes  to  zero  at  the  walls.  This  means 
that  the  velocity  is  a  function  of  the  distance,  rv  of  the  fluid  from 
the  axis  of  the  tube,  fx  is  a  constant  depending  on  the  nature  of 
the  fluid.     First  substitute  p  =  dV/dr,  as  in  (3) 

dp     P         P         ,         ,  P  ,  P 

•''d7  +  7=-  Vp>  •'•  "*P+i*fr-  "  Jfdr;  r.pr=  -  j^  +  Cfc® 

<ZF  P         ft"  -"„  P   9      ~  . 

•••  sf  -  -  W  +  * ;  F=  "  ^r  +  °llogr  +  °2' 

To  evaluate  C^  in  (5),  note  that  at  the  axis  of  the  tube  r  =  0.  This 
means  that  if  Gx  is  a  finite  or  an  infinite  magnitude  the  velocity 
will  be  infinite.  This  is  physically  impossible,  therefore,  GY  must 
be  zero.  To  evaluate  (72,  note  that  when  r  =  rlf  V  vanishes  and, 
therefore,  we  get  the  final  solution  of  the  given  equation  in  the 
form, 

P(ri2_r2) 

ilfJL 

which  represents  the  velocity  of  the  fluid  at  a  distance  rx  from  the 
axis. 

Examples. — (1)  Show  that  if  dPy/dx2  =  32,  and  a  particle  falls  from  rest, 
the  velocity  at  the  end  of  six  seconds  is  6  x  32  ft.  per  second  ;  and  the 
distance  traversed  is  £  x  32  x  36  ft. 

(2)  The  equation  of  motion  of  a  particle  in  a  medium  which  resists  di- 
rectly as  the  square  of  the  velocity  is  d2s/dt2  =  -  a(ds/dt)2.  Solve.  Hint. 
Substitute  as  in  (3);  .-.dplp2  +  adt  —  0;  ,'.p~l  =  at  +  Gl;  .*.  as  =  \og(at  +  Cj)  +  G2 ; 
etc. 

(3)  Solve  y  .  d^yjdx2  +  {dyjdxf  =  1.  Ansr.  if  =  x*  +  G^x  +  G2.  Hint. 
Use  (3),  pdp\dy  +  p2\y  =  \\y ;  .'.  py  =  x  +  Gx,  etc.   . 

Exact  equations  may  be  solved  by  successive  reduction.  The 
equation  of  motion  of  a  particle  under  the  influence  of  a  repulsive 
force  which  varies  inversely  as  the  distance  is 

d2s      a         ds  t  ,,*..„ 

dt*=!>   '''dt  =  al0^G'1;  .•.2/  =  ^Gog£--l)+C2, 

on  integration  by  parts.     The  equation  of  motion  of  a  particle 

under  the  influence  of  an  attractive  force  which  varies  inversely 

as  the  nth.  power  of  the  distance  is 

<Ps_       a  fdsy         2a    f    1  1    \ 

dt2~~      ?; -'•'■  \dt)    ~w-lVsn_1      aw-V*     *        ^ 

Again  integrating,  we  get 

„      „    In  -  1  f '       si{n-"ds 


Ja« 


i  _ 


§  134.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.         413 

According  to  the  tests  of  integrability,  this  may  be  integrated  when 
»  =  •  •  ■  h  f,  i  "  1,  f,  S,  •  .  •  ;    or  when  n  =  .  .  .  f ,  f ,  \,  0,  2,  |,  .  .  .    (7) 
as  indicated  on  page  210. 

Examples. — (1)  If  n  =  £,  we  get  (ds/dt)2  =  4a(a*  -  s*) ;  consequently, 
2sJa  .dt  =  -  (a^  -  s^)~^ds.  The  negative  sign  is  taken  because  s  and  t  are 
inverse  functions  of  one  another.  Add  and  subtract  2\/a/(3\/Wa^  -  sty- 
We  get  on  rearranging  terms, 

,_  /-  3\/s  +  2sfa  2\/o"         \ 

'*' t  =  3jl{s*(a* " s^*  +  2x/*(a*  -  si)*i  =  rr-(s*  + 2ai)  (a*  - s*)*- 

(2)  If  rf^/d*2  -  z.dz\dt  =  0;  show  that  Cx  +  z  =  {Cx  -  *)*>&'+ <h\ 
Hint.  We  get  on  substituting  p  =  ds/cft ;  dHjdt2  =  dp/d<  =  dzjdt   x  dp/da 

(3)  The  equation  of  motion  of  a  thin  revolving  disc  is 

<Pu      du      u  u  C,      ar* 

Hint.  Add  and  subtract  rdu/dr. 

(   d2u  du\      (  du        \  ■  d  /    dzA        d ,     ,       d  /ar4\      „ 

On  integration  we  get  an  ordinary  equation  of  the  first  order  which  can  be 
solved  (Ex.  (16),  p.  372)  by  substituting  vr  =  u. 

§  134.  Forced  Oscillations. 

We  have  just  investigated  the  motion  of  a  particle  subject  to 
an  effective  force,  d2s/dt2,  and  to  the  impressed  forces  of  restitu- 
tion, q2s,  and  resistance,  2/V.  The  particle  may  also  be  subjected 
to  the  action  of  a  periodic  force  which  prevents  the  oscillations 
dying  away.  This  is  called  an  external  force.  It  is  usually 
represented  by  the  addition  of  a  term  f(t)  to  the  right-hand  side 
of  the  regular  equation  of  motion,  so  that 

d2s      n.ds 

^+2/^  +  ^=/».      .         .         .         (1) 

The  effective  force  and  the  three  kinds  of  impressed  force  all 
produce  their  own  effects,  and  each  force  is  represented  in  the 
equation  of  motion  by  its  own  special  term.  The  term  comple- 
mentary function,  proposed  by  Liouville  (1832),  is  applied  to  the 
complete  solution  of  the  left  member  of  (1),  namely, 
d2s  ds 

p.+*3i  +  a*-J*    •     •     •     (2) 

The  complementary  function  gives  the  oscillations  of  the  system 


414  HIGHEK  MATHEMATICS.  §  134. 

when  not  influenced  by  disturbing  forces.  This  integral,  there- 
fore, is  said  to  represent  the  free  or  natural  oscillations  of 
the  system.  The  particular  integral  represents  the  effects  of  the 
external  impressed  force  which  produce  the  forced  oscillations. 
The  word  "  free  "  is  only  used  in  contrast  with  "  forced  ".  A  free 
oscillation  may  mean  either  the  principal  oscillation  or  any  motion 
represented  by  any  number  of  terms  from  the  complementary 
function. 

Let  equation  (1)  represent  the  motion  of  a  pendulum  when 
acted  upon  by  a  force  which  is  a  simple  harmonic  function  of  the 
time,  such  that 

d2s         ds 

dt2  +  ^dt  +  qh  =  koosnt  •  .  .  (3) 
We  have  already  studied  the  complementary  function  of  this 
equation  in  connection  with  damped  oscillations.  Any  particular 
integral  represents  the  forced  vibration,  but  there  is  one  particu- 
lar integral  which  is  more  convenient  than  any  other.     Let 

s  =  A  cos  nt  +  B  sin  nt,  .         .         (4) 

be  this  particular  integral.  The  complementary  function  contains 
the  two  arbitrary  constants  which  are  necessary  to  define  the 
initial  conditions ;  consequently,  the  particular  integral  needs  no 
integration  constant.  We  must  now  determine  the  forced  oscil- 
lation due  to  the  given  external  force,  and  evaluate  the  constants 
A  and  B  in  (4). 

First  substitute  (4)  in  (3),  and  two  identical  equations  result. 
Pick  out  the  terms  in  cos  nt,  and  in  sin  nt.  In  this  manner  we 
find  that 

-  An2  +  2Bfn  +  q2A  =  & ;  and,  -  Bn2  -  2Afn  +  a2B  =  0. 
Solve  these  two  equations  for  A  and  J5,  and  we  get 

_  k(q2  -  n2)        .        _  2fe/n.  ' 

(q2  -  n2)2  +Af2n2  '  (q2  -  n2)2  +  4/V  ^ 

It  is  here  convenient  to  collect  these  terms  under  the  symbols 
B,  cos  c,  and  sin  e,  so  that 

q2  -  n2  =  cos  e ;  2fn  =  sin  e ;  e  =  tan*1- 


q*  -  nl 
B  = 


(6) 


^2    _   W2)2  +   4y2w2  • 

.-.  A  =  Bcose;  B  =  i?sine.  .         .         (7) 

From  (4)  we  may  now  write  the  particular  integral 

s  =  B(co3  e  .  cos  nt  +  sin  €  .  sin  nt)}     .         .         (8) 


§  134.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        415 

or,  making  a  well-known  trigonometrical  substitution, 

s  =  B  cos  (nt  -  c).  .         .         .         (9) 

This  expression  represents  the  forced  oscillations  of  the  system 
which  are  due  to  the  external  periodic  force.  The  forced  oscil- 
lation is  not  in  the  same  phase  as  the  principal  oscillation  induced 
by  the  effective  force,  but  lags  behind  a  definite  amount  e. 

R  in  (6)  always  has  the  same  sign  whatever  be  the  signs  of  n  and  q ;  2/  is 
positive,  hence  sin  €  is  positive  and  the  angle  €  lying  in  the  first  two  quadrants 
ranges  from  0  to  ir.  On  the  other  hand,  the  sign  of  cos  e  does  depend  upon 
the  relative  magnitudes  of  n  and  q.  If  q  be  greater  than  n,  e  is  in  the  first 
quadrant ;  if  q  is  less  than  n,  e  is  in  the  second  quadrant  (see  Table  XV.,  page 
610) ;  if  q  =  n,  e  =  £ir.  The  amplitude,  jB,  of  the  forced  vibration  is  propor- 
tional to  the  intensity,  k,  of  the  external  force.  If  /  be  small  enough,  we  can 
neglect  the  term  containing  /  under  the  root  sign,  and  then 

R=       * 

22  -  n2' 

In  that  case  the  more  nearly  the  numerical  value  of  q  approaches  n,  the 
greater  will  be  the  amplitude,  R,  of  the  forced  vibration.  Finally,  when 
q  =  n,  we  should  have  an  infinitely  great  amplitude.  Consequently,  when 
q  =  n,  we  cannot  neglect  the  magnitude  of  /2  and  we  must  have 

Umax  -  2nf 

so  that  the  magnitude  of  R  is  conditioned  by  the  damping  constant.  If /=0 
as  is  generally  assumed  in  the  equation  of  motion  of  an  unresisted  pendulum, 

d2s 

^p  +  g2s  =  ft  cos  ni, (10) 

the  particular  integral  of  which 

is  indeterminate  when  n  =  q.  The  physical  meaning  of  this  is  that  when  a 
particle  is  acted  upon  by  a  periodic  force  "in  step"  with  the  oscillations  of 
the  particle,  the  amplitude  of  the  forced  vibrations  increase  indefinitely,  and 
equation  (10)  no  longer  represents  the  motion  of  the  pendulum.  See  page 
404.  As  a  matter  of  fact,  equation  (10)  is  only  a  first  approximation  obtained 
by  neglecting  the  second  powers  of  small  quantities  (see  E.  J.  Routh's  Ad- 
vanced Rigid  Dynamics,  London,  222,  1892).  I  assume  that  the  reader  knows 
the  meaning  of  q  and  n,  if  not,  see  pages  137  and  397. 

If  the  motion  of  the  particle  is  strongly  damped,  the  maximum  excita- 
tion does  not  occur  when  n  =  q,  but  when  the  expression  under  the  root  sign 
is  a  minimum.  If  n  be  variable,  the  expression  under  the  root  sign  is  a 
minimum  when  n2  =  q2  -  2/2,  as  indicated  in  Ex.  (5),  page  166 ;  and  R  is  there- 
fore a  maximum  under  the  same  conditions.  If  n  be  gradually  changed  so 
that  it  gradually  approaches  q,  and  at  the  same  time  /be  very  small,  R  will 
remain  small  until  the  root  sign  approaches  its  vanishing  point,  and  the 
forced  oscillations  attain  a  maximum  value  rather  suddenly.     For  example, 


416 


HIGHER  MATHEMATICS. 


§134. 


if  a  tuning  fork  be  sounded  about  a  metre  away  from  another,  the  minute 
movements  of  air  impinging  upon  the  second  fork  will  set  it  in  motion. 

If  /  be  large,  the  expression  under  the  root  sign  does  not  vanish  and 
there  is  no  sudden  maximum.  The  amplitudes  of  the  free  vibration  changes 
gradually  with  variations  of  n.  The  tympanum  of  the  ear,  and  the  receiver 
of  a  telephone  or  microphone  are  illustrations  of  this.  Every  ship  has  its 
own  natural  vibration  together  with  the  forced  one  due  to  the  oscillation  of 
the  waves.  If  the  two  vibrations  are  synchronous,  the  rolling  of  the  ship 
may  be  very  great,  even  though  the  water  appears  relatively  still.  "  The  ship 
Achilles"  says  White  in  his  Manual  of  Naval  Architecture,  "was  remarkable 
for  great  steadiness  in  heavy  weather,  and  yet  it  rolled  very  heavily  off  Port- 
land in  an  almost  dead  calm."  The  natural  period  of  the  ship  was  no  doubt 
in  agreement  with  the  period  of  the  long  swells.  Iron  bridges,  too,  have 
broken  down  when  a  number  of  soldiers  have  been  marching  over  in  step 
with  the  natural  period  of  vibration  of  the  bridge  itself.  And  this  when  the 
bridge  could  have  sustained  a  much  greater  load. 

The  complete  solution  of  the  linear  equation  is  the  sum  of  the 
particular  integral  and  the  complementary  function.  If  the  latter 
be  given  by  II,  page  406,  the  solution  of  (3)  must  be  written 

s  =  Bcos(nt  -  e)  +  e~at(Clco$qt  +  C2sing£). 
We  can  easily  evaluate  the  two  integration  constants  Gx  and  C2, 
when  we  know  the  initial  conditions,  as  illustrated  in  the  preced- 
ing section.     If  the  particle  be  at  rest  when  the  external  force 
begins  to  act, 

fn  .  If        \ 

C1  =  -Bco3€;  02  =  -  Bl  -sine  +  —  cosel. 

At  the  beginning,  therefore,  the  amplitude  of  the  free  vibrations  is 


-y    a  b 

Fig.  151. 

of  the  same  order  of  magnitude  as  the  forced  oscillation.  If  the 
damping,  2/,  is  small,  and  n  is  nearly  equal  to  q,  the  damping 
factor,  e'"*,  will  be  very  great  nearly  unity;  2f/q  is  nearly  zero; 
n/q  is  nearly  unity ;  and  e  is  nearly  ^tt.  In  that  case,  Gx  =  0,  and 
G2  =  -  B.     The  motion  is  then  approximately 

s  =  B(sin  nt  -  sin  qt). 
The  two  oscillations  sin  nt  and  -  sin  qt  are  superposed  upon  one 


§  134.     HOW  TO  SOLVE  DIFFEKENTIAL  EQUATIONS.        417 

another.  If  these  two  harmonic  motions  functions  are  plotted 
separately  and  conjointly,  as  in  Fig.  151,  we  see  at  once  that  they 
almost  annul  one  another  at  the  beginning  because  the  one  is 
opposed  by  the  other.  This  is  shown  at  A.  In  a  little  while,  the 
difference  between  q  and  n  becomes  more  marked  and  the  ampli- 
tude gradually  increases  up  to  a  maximum,  as  shown  at  B.  These 
phenomena  recur  at  definite  intervals,  giving  rise  to  the  well- 
known  phenomena  of  interference  of  light  and  sound  waves,  beats, 
etc. 

Examples. — (1)  Ohm's  law  for  a  constant  current  is  E  =  BC ;  for  a 
variable  current  of  G  amperes  flowing  in  a  circuit  with  a  coefficient  of  self- 
induction  of  L  henries,  with  a  resistance  of  B  ohms  and  an  electromotive 
force  of  E  volts,  Ohm's  law  is  represented  by  the  equation, 

E  =^BC  +  L  .  djCdt,  ....  (11) 
where  dC/dt  evidently  denotes  the  rate  of  increase  of  current  per  second,  L  is 
the  equivalent  of  an  eleotromotive  force  tending  to  retard  the  current. 

(i)  When  E  is  constant,  the  solution  of  (11)  has  been  obtained  in  a  preced- 
ing set  of  examples,  G=E/B  +  Be-mlLt  where  B  is  the  constant  of  integration. 
To  find  Bt  note  that  when  t  =  0,  0  =  0.  Hence,  G  =  E{1  -  »-  ni^/B.  The 
second  term  is  the  so-called  "  extra  current  at  make,"  an  evanescent  factor 
due  to  the  starting  conditions.  The  current,  therefore,  tends  to  assume  the 
steady  condition :  G  =  E/B,  when  t  is  very  great. 

(ii)  When  C  is  an  harmonic  function  of  the  time,  say,  O  =  C0  sin  qt ; 
.•.  dGjdt  =  C0qcosqt.  Substitute  these  values  in  the  original  equation  (11), 
and  E  =  BC0  sin  qt  +  LCQq  cos  qt,  or,  E  =  G0  *JB?  +  L2q* .  sin  (qt  +  e),  on 
compounding  these  harmonic  motions,  page  138,  where  e  =  t&n-l(LqlB),  the 
so-called  lag l  of  the  current  behind  the  electromotive  force,  the  expression 
s/(B2  +  Z/2g2)  is  the  so-called  vmpedance. 

(iii)  When  E  is  a  function  of  the  time,  say,  f(t), 

C  =  B6~%+%L\eLf(t)dt, 


K^ffi 


where  B  is  the  constant  of  integration  to  be  evaluated  as  described  above. 

(iv)  When  E  is  a  simple  harmonic  function  of  the  time,  say,  E= EQsm  qt, 
then, 

«       EjBBwqt-LqcoBqt) 
U  =  Be    l+  £2  +  Ly 

The  evanescent  term  e-MlL  may  be  omitted  when  the  current  has  settled 
down  into  the  steady  state.     (Why  ?) 

1An  alternating  (periodic)  current  is  not  always  in  phase  (or,  "in  step")  with 
the  impressed  (electromotive)  force  driving  the  current  along  the  circuit.  If  there 
is  self-induction  in  the  circuit,  the  current  lags  behind  the  electromotive  force ;  if 
there  is  a  condenser  in  the  circuit,  the  current  in  the  condenser  is  greatest  when  the 
electromotive  force  is  changing  most  rapidly  from  a  positive  to  a  negative  value,  that 
is  to  say,  the  maximum  current  is  in  advance  of  the  electromotive  force,  there  is  then 
said  to  be  a  lead  in  the  phase  of  the  current. 

DD 


418  HIGHER  MATHEMATICS.  §  135. 

(v)  When  E  is  zero,  0  =  Be- miL.  Evaluate  the  integration  constant 
B  by  putting  C  =  C0,  when  t  =  0. 

(2)  The  relation  between  the  charge,  q,  and  the  electromotive  force,  E,  of 
two  plates  of  a  condenser  of  capacity  C  connected  by  a  wire  of  resistance  B,  is 
E  =  B  .  dqjdt  +  q/C,  provided  the  self-induction  is  zero.  Solve  for  q.  Show 
that  when  if  E  be  0 ;  q  =  Q^e-'!™;  (Q0  is  the  charge  when  t  =  0).  If  E  be 
constant ;  q  =  CE  +  Be-"*0.    If  E  =  f{t) ; 


1         ±  f     t_  _£_ 

q  =  j=e    no  I  excf(t)dt  +  Be    rc. 


(3)  Show  if  JJ—Mn*;  4  =  ife-F.+  ^'^^r8^. 

§  135.    How  to  find  Particular  Integrals. 

The  particular  integral  of  the  linear  equation, 

(D*  +  PD  +  Q)y  =  f(x),'  .  .  .  (1) 
it  will  be  remembered,  is  any  solution  of  this  equation — the  simpler 
the  better.  The  particular  integral  contains  no  integration  con- 
stant. The  complete  solution  of  the  linear  equation  is  the 
sum  of  the  complementary  function  and  the  particular  integral. 
Complete  solution  =  complementary  function  +  particular  integral. 

We  must  now  review  the  processes  for  finding  particular  in- 
tegrals. Let  B  be  written  in  place  of  f(x),  so  that  (1)  may  be 
written  f(D)y  =  B.     Consequently,  we  may  write, 

2/=/(D)-^;  <>T,y  =  ~.  .         .         (2) 

The  right-hand  side  of  either  of  equations  (2),  will  furnish  a 
particular  integral  of  (1).  The  operation  indicated  in  (2)  depends 
on  the  form  of  f(D).     Let  us  study  some  particular  cases. 

J.  When  the  operator  f(D)  ~  l  can  be  resolved  into  factors. 

Suppose  that  the  linear  equation 

3 -£+*-* 

can  be  factorized.  The  complementary  function  can  be  written 
down  at  sight  by  the  method  given  on  page  401, 

(Z>2  -  5D  +  6)2/  =  0 ;  or,  (D  -  3)  (D  -  2)y  =  0. 
According  to  (2),  the  particular  integral,  yv  is 

Vl  =  (D  -  3)  (D  -  2)^  =  \D~^3  ~  W^1JB  ; 
On  page  396  we  have  defined  f(D)  ~  lB  to  be  that  function  of  x 
which   gives  B  when    operated  upon   by  f(D).      Consequently, 


§  135.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        419 

D  ~  lx2  m  jx2dx.  Hence,  D  -  3  acting  upon  (D  -  3)  ~  XB  must,  by 
definition,  give  B.  But  (D  -  3)  "  XB  is  the%  particular  integral  of 
the  equation 

so  that  from  (2),  page  387,  y  -  e-'-vje'-vjRdx  =  e^je-^Bdx. 

.\y1  =  e**\e  ~  **Bdx  -  e**fe  "  **Bdx. 
from  (2).     The  general  solution  is 

...  y  =  GYe**  +  Ce2*  +  eZx\e  ~  ^Bdx  -  e^je  '  2xBdx. 

Examples. — (1)  In  the  preceding  illustration,  put  B  =  e4x  and  show  that 
the  general  solution  is,  Gx&x  +  C2e2a:  +  %eix. 

(2)  If  (D2  -  4D  +  S)y  =  2e*x\  y  =  Gxex  +  G^x  +  are3*. 

(3)  Solve  cPyjdx2  -  3dy/dx  +  2y  =  a3*.  In  symbolic  notation  this  will 
appear  in  the  form,  (D  -  1)  (D  -  2)y  =  e3x.  The  complementary  function  is 
y  =  Gxex  +  G^P*.     The  particular  integral  is  obtained  by  putting 

Vl  =  (D  -  2)  (D  -  l)6**  "  (dTT~  F^i)^' 
according  to  the  method  of  resolution  into  partial  fractions.     Operate  with 
the  first  symbolic  factor,  as  above,  yx  =  e^je  ~  2xe?xdx  -  ^je  ~  xeSxdx  =  %e3x. 
The  complete  solution  is,  therefore,  y  =  Gxex  +  C^2*  +  $«3*« 

II.  When  Bis  a  rational  function  of  x,  say  x*. 

This  case  is  comparatively  rare.  The  procedure  is  to  expand 
f(D)  ~ 1  in  ascending  powers  of  D  as  far  as  the  highest  power  of  x 
in  B.  The  expansion  may  be  done  by  division  or  other  convenient 
process. 

Examples. — (1)  Solve  dPyfdx2  -  Idyjdx  +  4y  ■  x*.  The  complementary 
function  is  y  =  e^A  +  Bx) ;  the  particular  integral  is : 

(2TDJi*  =  K1  +  D  +  iP'Y  ~lP*  +  *X  +  8)* 
You  will,  of  course,  remember  that  the  operation  Dx2  is  2x  ;  and  JD2a?2  is  2. 

(2)  If  d^y/dx2  -y  =  2  +  5x;  y  =  Gxex  +  G#  -«  -  5x  -  2. 

(3)  The  particular  integral  of  (D3  +  3D2  +  2D)yt  =  x2  is  ^a;(2a;2-9a;  +  21) ; 
the  complementary  function  is  Ox  +  G&-2*  +  Gg~x.    The  steps  are 

2D  +  3D*  +  D*x9ls2D\l  +  \D  +  \Dyx*=W\}  "  2D  +  lD*)&- 

Now  proceed  as  in  Ex.  (1)  for  the  operation  Dx2  and  D2x2.    Then  note  that 

-=x  =  jxdx;  -^a;2=»  jx2dx't  etc 

111.  When  B  contains  an  exponential  factor,  so  that  B  =  e^X. 

Two  cases  arise  according  as  X  is  or  is  not  a  function  of  xt  a  is 
constant, 

DD* 


420  HIGHER  MATHEMATICS.  §  135. 

(i)  When  X  is  a  function  of  x.     Since  IPe"*  =  aneax,  where  n  is 
any  positive  integer,  we  have 

D{eTX)  =  eaxDX  +  ae^X  =  eax(D  +  a)X, 
and  generally,  as  in  Leibnitz'  theorem,  page  67, 
DneaxX  =  eax(D  +  a)nX; 

J)neax  2  I 

•'•  Ur+lifZ  =  fX ;  a„d  (p^Xe-  =  e-'WnX.         (3) 

Consequently,  the  operation  /(D)  -  V*X  is  performed  by  trans- 
planting eax  from  the  right-  to  the  left-hand  side  of  the  operator 
/(D) " x  and  replacing  D  by  D  +  a.  This  will,  perhaps,  be  better 
understood  from  Exs.  (1)  and  (2)  below. 

(ii)  When  X  is  constant,  operation  (3)  reduces  to 

The  operation  /(D)  "  V*  is  simply  performed  by  replacing  D  by  a. 

Examples.— -(1)  Solve  d^yjdx^  -  2dy/dx +y= a?VK    The  complete  solution, 
by  page  418,  is  (02  +  xG^e?  +  (D  +  2D  + 1)  -  1o¥Ja:.     The  particular  integral  is 

jy  -  lb  +  ix2e3x  =  (fr-iMB-i)8*"- 

By  rule  :  tPx  may  be  transferred  from  the  right  to  the  left  side  of  the  operator 
provided  we  replace  D  by  D  +  3. 

We  get  from  J  above,  edx(±x*-%x  +  %),  as  the  value  of  the  particular  integral. 

(2)  Evaluate  (D  -  1)  _  Vlog  x.     Ansr.  e*(x  log  a;  -  a;) ;   or  aa*log  (xje). 
Integrate  jlogxdx  by  parts. 

(3)  Find  the  particular  integral  in  (D2  -  3D  +  2)y  =  e*x, 

F4D  +  2*3"  =  32  -  3 .  3  +  2***  =  &**• 

(4)  Show  that  \ex,  is  a  particular  integral  in  d^y/dx2  +  2dy[dx  +  y  =  eXt 

(5)  Repeat  Ex.  1, I",  above,  by  this  method. 

An  anomalous  case  arises  when  a  is  a  root  of  /(D)  =  0.     By  this 
method,  we  should  g^t  for  the  particular  integral  of  dy/dx -y  =  ex. 
1      *         e* 

The  difficulty  is  evaded  by  using  the  methodi(3)  instead  of  (4).  Thus, 

l  i 

ex  =  e%r.  1  =  xe*. 


D  -  1  D 

The  complete  solution  is,  therefore,  y  =  Gex  +  xe*. 

Another  mode  of  treatment  is  the  following  :  If  a  is  a  root  of 
/(D)  =  0,  then  as  on  page  354,  D  -  a  must  be  a  factor  of  /(D). 


§  135.     HOW  TO  SOLVE  DIFFEKENTIAL  EQUATIONS.        421 

Consequently, 

/£!>)-  (D-ay(D); 

and  the  particular  integral  is 

•'•  ]{Wf'  -  W^-fWf"  =  W1^  •  TW6** =  e<"7w\dx-  (5) 

If  the  factor  D  -  a  occurs  twice,  then  following  the  same  rule 

pf" = ur^ymr = e^)  -mr = e<aml\dxdx- ;  (6) 

and  so  on  for  any  number  of  factors. 

Examples. — (1)  Find  the  particular  integrals  in,  (D  +  l)3y  =  e~x.  Ansr. 
\<x?e~x.  Hint.  Use  (6)  extended;  or,  since  the  root  a  is  -  1,  we  have  to 
evaluate  e  ~  * .  D  ~ 3 ;  that  is,  e  ~  xjjjdxdxdx. 

(2)  (D3  -  l)y  =  xex.  Ansr.  e*fox2  -  {x).  Hint.  By  the  method  of  (3), 
and  Ex.  (8),  II.,  above, 

i  i  i  i/i  i  \      v®.^ 

IV.  When  B  contains  sine  or  cosine  factors. 

By  the  successive  differentiation  of  sin  nx}  page  67,  we  find  that 

_  .  d(smnx)  ^0  .  d2(8mnx) 

Dsmnx  =  -^-j =  noosnx ;  D2smnx  =       ,  2 — -  =»  -  w2sin nx ; . . . 

.*.  (D2)nain(nx  +  a)  «  (-  n2)nsin(nx  +  a), 
where  w  and  a  are  constants.     And  evidently 

/(D2)8in(wo;  +  a)  =  /(  -  w2)sin(rac  +  a). 
By  definition  of  the  symbol  of  operation,  f{D2)  ~  1,  page  396,  this 
gives  us 

•*•  y^2)8in(^  +  <0  -  j^-^Bm(nx  +  a).         .         (7) 
It  can  be  shown  in  the  same  way  that, 

y^fOB(nx  + a)  -  ^  _  rffoafttx  +  a).  .  *      (8) 

Examples.— (1)  cPyjdx*  +  cPyjdx*  +  dy/da  +  y  =  sin  2a.  Find  the  par- 
ticular integral. 

R  *  1  1 

/(5)  =  D3  +  Z>2  +  D  +  1 8in  2x  m  (i*  +  l)  +  D(D»+l)  8m  2a?' 
Substitute  -  22  for  D2  as  in  (7).     We  thus  get  -  £(£  +  1)  ~  asin  2x.     Multiply 
and  divide  by  D  -  1  and  again  substitute  D2  =  ( -  22)  in  the  result.    Thus  we 
get  jV(D  -  1)  sin  2x  ;  or  ^(2  cos  2x  -  sin  2x). 

(2)  Solve  d?yldaP-k*y=s cos  ma?.     Ansr.  C^tf **  +  C^  ~  **  -  (cosmo?)/(m2  +  A;a). 

(3)  If  a  and  £  are  the  roots  of  the  auxiliary  equation  derived  from  Helm- 
holtz's  equation,  diyjdf-  +  mdyjdt  +  n2y  =  a  sin  nt,  for  the  vibrations  of  a  tuning- 
fork,  show  that  y=Cxe^  +  C^  -  (a  cos  nt)lmn  is  the  complete  solution. 


422  HIGHER  MATHEMATICS.  •    §  136. 

An  anomalous  case  arises  when  D2  in  D2  +  n2  is  equal  to  -  n2. 
For  instance,  the  particular  integral  of  d2y/dx2  +  n*y  =  S£rac,  is 
(D2  +  n2)  ~  l  JS  nx.  If  the  attempt  is  made  to  evaluate  this,  by 
substituting  D2  »  -  n2,  we  get  $*  nx{  -  n2  +  n2)  ~  1  =  oo  £*  rac. 
We  were  confronted  with  a  similar  difficulty  on  page  420.  The 
treatment  is  practically  the  same.  We  take  the  limit  of 
(D2  +  n2)  - l  SS  nx,  when  n  of  2£  nx  and  -  D2  become  n  +  h  and  /i 
converges  towards  zero.  In  this  manner  we  find  that  the  par- 
ticular integral  assumes  the  form  . 

x  sin  nx                                      x  cos  nx 
+  — o —  if  B  =  cos  nx ;  and „ — t  if  B  =  sin  nx.      (9) 

Examples. — (1)  Evaluate  (D2  +  4)  _  ^os  2sc.    Ansr.  %x  sin  2x. 

(2)  Show  that  -  %x  cos  x,  is  the  particular  integral  of  (Z)2  +  l)y  =  sin  x. 

(8)  Evaluate  (D2  +  4)  -  Jsin  2a;.    Ansr.  -  \x  cos  2a;. 

V.  When  B  contains  any  function  of  x,  say  X,  such  that  B  =  xX. 

The  successive  differentiation  of  a  product  of  two  variables, 
xX,  gives,  pages  40  and  67, 

DnxX  =  xDnX  +  nTT-^X. 
.-.  f(D)xX  =  xf(D)X  +  /XD)X       .         .       (10) 
Substitute   F  =  f(D)X,  where  T"  is  any  function  of  x.     Operate 
with /(D)  "  K     We  get  the  particular  integral 

1  f         1  ffljy  m, 

/(Df  A  ~  V7(D)      /(D)2 J A*  •         '      (11) 

where /'(D)//(D)  is  the  differential  coefficient  of /(D)  -  * 

Examples. — (1)  Find  the  particular  integral  in  cPyjdx3  -  y  =  xe2*.  Ke- 
member  that  f'(D)  is  the  differential  coefficient  of  D3  -  1.  From  (11)  the 
particular  integral  is 

{.  -  1A-v  3D* }B£rif.  =  {*  -  i.8,4}ie-  =  (=  -  g)* 

(2)  Show  in  this  way,  that  the  particular  integral  of  (D  -  l)y  =  x  sin  x 
is 

1  1  D  +  l    .  (D  +  l)2    . 

a;5Tlsina!  -  (D  -  1)2  sin  a;  =  ^^—\  amx  ~  (j)2  _  1)aBing; 

=s  £a;(.D  +  1)  cos  x  -  %(D2  +  2D  +  1)  cos  a;  =  £a:(cos  x  +  sin  x)  -  £  cos  a?. 

(3)  If  d*y/dx2  -  y  =  «2cosa;;  y  =  Gyex  +  C<#-x+  xsiiix  +  %  cosa?(l  -  a;8). 
Hint.  By  substituting  xX  in  place  of  X  in  (10),  the  particular  integral  may 
be  transformed  into 

fm*x  =  {*fk*^fw+f7m)x'     ■    ■   (12) 

where  /'(Z>)//(2))2,  and  f"(D)jf{D3)  respectively  denote  the  first  and  second 
differential  coefficient  of  f(D)  ~  K  Successive  reduction  of  x^X  furnishes  a 
similar  formula.    The  numerical  coefficients  follow  the  binomial  law.     Re- 


§  136.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.       423 

turning  to  the  original  problem,  the  first  and  second  differential  coefficients 
of  (D2  -  1)  -  »  are  -  2D(D2  -  1)  -  2,  and  (2D2  -  2)  (D2  -  1)  - 3.    Hence, 

jjpf*x»  *  -  {  ^"STTTT  +  2a;  (jy  _  1)2  +   (2)3  _  i)  jcos  0 ; 
.•.  y1  =  -  ^oj^os  sc  -  $  sin  sc  +  £  cos  x. 
(4)  Solve  (Pyjda?  -  y  =  x  sin  z.     The  particular  integral  consists  of  two 
parts,  %{(x  -  3)  cos  x  -  x  sin  x).    Tho  complementary  function  is 

(Vs  +  e"4*{a,sin(£\/3a;)  +  C3cos(£  -s/3aj)}. 

§  136.    The  Gamma  Function. 

The   equation  of  motion  of  a  particle  of  unit  mass  moving 

under  the  influence  of  an  attractive  force  whose  intensity  varies 

inversely  as  the  distance  of  the  particle  away  from  the  seat  of 

attraction  is  obviously 

dh  _      a 

dt*  ~      s' 

where   a  is  a   constant,  and   the   minus   sign  denotes  that  the 

influence  of  the  force  upon  the  particle  diminishes  as  time  goes 

on.     To  find  the  time  occupied  by  a  particle  in  passing  from  a 

distance  s  =  5  to  s  =  s0,  we  must  integrate  this  equation.     Here, 

on  multiplying  through  by  2ds/dt,  we  get 


(ds   d2s         a   (1    ds         1/dsV  , 

)  when  s  =  sQ, 

J.„Vlog^ 


(1) 


For  the  sake  of  convenience,  let  us  write  y  in  place  of  logs0/s. 

From   the   well-known    properties    of    logarithms   discussed    on 

page  24,  it  follows  that  if  s  =  s0,  y  —  0 ;  and  if  s  =  0,  y  =  oo. 

Hence,  passing  into  exponentials, 

s  s 

log-0  =  y  ;  j  =  ev  ;  s  =  sae  -»;  ds  =  -  sQe~vdy  ; 

_••■— Mt?-  *!>-"-*  •  « 

It  is  sometimes  found  convenient,  as  here,  to  express  the  solu- 
tion of  a  physical  problem  in  terms  of  a  definite  integral  whose 
numerical  value  is  known,  more  or  less  accurately,  for  certain 
valines  of  the  variable.     For  example,  there  is  Soldner's  table  of 


424  HIGHER  MATHEM  A.TICS.  §  136. 

jJ(loga;)  ~  ldx  ;  Gilbert's  tables  of  Fresnel's  integral  JJcos  \irv  .  dv,  or 
JJsin  i-n-v .  dv  ;  Legendre's  tables  of  the  elliptic  integrals  ;  Kramp's 
table  of  the  integral  j™e  *  *2 .  dt ;  and  Legendre's  table  of  the  in- 
tegral Qe~xxn~l.  dx,  or  the  so-called  "gamma  function".  We 
shall  speak  about  the  last  three  definite  integrals  in  this  work. 

Following  Legendre,  the  gamma  function,  or  the  "  second 
Eulerian  integral,"  is  usually  symbolised  by  T(n).  By  definition, 
therefore, 

r(n)  =  J   e-xxn~l.dx.  (3> 

Integrate  by  parts,  and  we  get 

e  ~  xxn .  dx  =  n  \    e  ~  *xn  " 1 .  dx  -  e  -  xxn.       .         (4) 
o  Jo 

The  last  term  vanishes  between  the  limits  x  =  0  and  x  =  oo. 
Hence 

e~*xn.dx  =  n\    e"xxn^1.dx.      .        .         (5) 

o  Jo 

In  the  above  notation,  this  means  that 

T(n  +  1)  =  »r(n).  ...         (6) 

If  n  is  a  whole  number,  it  follows  from  (6),  that 

V(n  +  1)  =  1 .  2  .  3  . . .  n  =  n !  .  .  .  (7) 
This  important  relation  is  true  for  any  function  of  n,  though  n ! 
has  a  real  meaning  only  when  n  is  integral. 

The  numerical  value  of  the  gamma  function  has  been  tabu- 
lated for  all  values  of  n  between  1  and  2  to  twelve  decimal  places. 
By  the  aid  of  such  a  table,  the  approximate  value  of  all  definite 
integrals  reducible  to  gamma  functions  can  be  calculated  as  easily 
as  ordinary  trigonometrical  functions,  or  logarithms.  There  are 
four  cases : 

1.  n  lies  between  0  and  1.     Use  (16). 

2.  n  lies  between  1  and  2.     Use  Table  V.,  below. 

3.  n  is  greater  than  2.  Use  (6)  so  as  to  make  the  value  of  the 
given  expression  depend  on  one  in  which  n  lies  between  1  and  2. 

4.  r(l)  =  1;  T(2)  =  1;  T(0)  -  oo;  r(j)  -  JZ.   .       ,        (8) 

I.  The  conversion  of  definite  integrals  into  the  gamma  function. 

The  following  are  a  few  illustrations  of  the  conversion  of  de- 
finite integrals  into  gamma  functions.  For  a  more  extended 
discussion  special  text-books  must  be  consulted.      If  a  is  inde- 


§  136.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        425 
pendent  of  xf 


1 


e-a*xm^  i    dx  =  —v(m) ;  (9) 


o  <* 


Jo         K  J  Jo(l+^r+M       r(m  +  w)  V     ; 

The  first  member  of  (10)  is  sometimes  called  the  first  Eulerian 
integral,  or  beta  function.  It  is  written  B(m,  n).  The  beta  func- 
tion is  here  expressed  in  terms  of  the  gamma  function.  Substitute 
x  =  ay  lb  in  the  second  member  of  (10),  and  we  get 

J"     ym~ldy  T(m)T(ri) 

0(ay  +  b)m+n  ~  ambnT(m  +  n)' 


(11 


Other  relations  are : 

Jo  Jo  r^n  +  1) 

f  sin** .  oob*  .  ^  =  T[ittJ)]'T[tiqn1)l         ^ 

Jo  2rB0>  +  ff)  +  i] 

f1  „,    AVj        r(n+i)  f1  ml      n,      (-i)wr(w  +  i)  „., 
)*•  logy  ib  -  ^n^  ]o-rlogon&  =    (J+  lr+1   .  (H) 

[\ne-«*dx  =»  a-("+1>r(w  +  1) ;    fV«2*2dz  =  ffiil  =  i^.    (15) 
Jo  Jo  a  a 

You  can  now  evaluate  (2).     We  get 

Compare  the  result  with  that  obtained  by  the  process  of  integra- 
tion described  on  pages  342  and  344. 

Examples.— (1)  Evaluate  P"  sin^a? .  dx.     Hint.  From  (12),       r(6)^' 

n£   llilllidL-^  _  ^    JL   !    5   3    X 
•''    2    '      5.4.3.2.1      ~2'l0'8'6*4'2 

,OV    -T,       ,  /■"  -K     J  TT        ,„V  r(6)         5.4.3.2.1 

(2)  Evaluate  /    g-^.dz.    Use  (9).     Ansr.  -~V  = « 

/*  a^-^a:  7r  tt 

-,   ,  _     =     ■    „,  ,  show  that  r(ra) .  r(l  -  m)  =    .    ^    ;  and 
0     1  +  x  sin  mw  \    /      \  /      gm  W7r 

r(l  +  m) .  r(l  -  m)  =    .         ,  by  putting  m  +  n  =  1  in  the  beta  function,  etc. 

OX  XI    if  LTV 

These  two  results  can  be  employed  for  evaluating  the  gamma  function  when 
n  lies  between  0  and  1.     By  division 

.    ,      r(l  +  m)  n„ 

r(«)  =      m     •  •       •       •       •       (16) 

If    m  =  i,    r(i)  -  3-6254;    log  r(£)  =  0-5594;    if    w  -  J,    r(£)  =  1-7725; 


426 


HIGHER  MATHEMATICS. 


§  137. 


log10  r(J)  -  0-2486;  and'  if  m  =  |,  r(|)  =  1'2253;  log10  r(f)  =  0-0883;  where 
the  bar  shows  that  the  figure  has  been  strengthened. 

II.   Numerical  computations. 

Table  V.  gives  the  value  of  log10r(w)  to  four  decimal  places  for 
all  values  of  n  between  1  and  2.  It  has  been  adapted  from  Le- 
gendre's  tables  to  twelve  decimal  places  in  his  Exercises  de  Galcul 
Integral,  Paris,  2,  18,  1817.  For  all  values  of  n  between  1  and  2, 
log  Y(n)  will  be  negative.  Hence,  as  in  the  ordinary  logarithmic 
tables  of  the  trigonometrical  functions,  the  tabular  logarithm  is 
often  increased  by  the  addition  of  10  to  the  logarithm  of  T(n). 
This  must  be  allowed  for  when  arranging  the  final  result. 


Table  V.- 

-Common  Logarithms  of  r(n)  from  n  = 

1-00  to  n  =  1-99. 

n. 

0-00. 

o-oi. 

0*02. 

0-03. 

004. 

0-05. 

0-06. 

0-07. 

0-08. 

0-09. 

1-0 
1-1 
1-2 
1-3 
1-4 

1-5 
1-6 
1-7 
1-8 
1-9 

o-oooo 

1-9783 
1-9629 
1-9530 
1-9481 

1-9475 
1-9511 
T-9584 
1-9691 
1-9831 

1-9975 
1-9765 
1-9617 
1-9523 
1-9478 

1-9477 
1-9517 
1-9593 
1-9704 
1-9846 

1-9951 
1-9748 
1-9685 
1-9516 
1-9476 

1-9479 
1-9523 
1-9603 
1-9717 
1-9862 

1-9928 
1-9731 
1-9594 
1-9510 
1-9475. 

1-9482 
1-9529 
1-9613 
1-9730 
1-9878 

1-9905 
1-9715 
1-9583 
1-9505 
1-9473 

1-9485 
1-9536 
1-9623 
1-9743 
1-9895 

1-9883 
1-9699 
1-9573 
1-9500 
1-9473 

1-9488 
I  9543 
1-9633 
P9757 
19912 

1-9862 
1-9684 
1-9564 
1-9495 
1-9472 

1-9492 
1-9550 
1-9644 
1-9771 
1-9929 

1-9841 
1-9669 
T-9554 
1-9491 
1-9473 

1-9496 
1-9558 
1-9656 
1-9786 
1-9946 

1-9821 
1-9655 
1-9546 
1-9487 
1-9473 

1-9501 
19566 
T-9667 
1-9800 
1-9964 

1-9802 
1-9642 
1-9538 
1-9483 
1-9474 

1-9506 
1-9575 
1-9679 
1-9815 
1-9982 

log10vV  =  0-24857493635  =  log10r(£). 

f\n      

Example. — Evaluate    I     vsin  x .  dx.    Ansr.  1*198.    Hint.  Use  (13),  q  =  0, 


p  =  £.     Hence, 

*»   , 

vsin  x  .  dx 


/: 


Idllii).,    r®r^ 


ar(f) 


log 


2r(f) 


=  log  r(f )  +  log  r(£)  -  log  2  -  log  r(f) 


=  0-0883  +  0-2485  -  0'3010  -  1*9573  =  0-0823  =  log  1-198. 

§  137.  Elliptic  Integrals. 

The  equation  of  motion  of  a  pendulum  swinging  through  a 
finite  angle  is 

g --*-•.      ■       •      CD 

where   6  represents  the   angle,  BOA  (Fig.    152), 

described  by  the  pendulum   on   one   side  of  the 

vertical  at  the  time  t,  reckoned  from  the  instant 

Fig.  152  the  pendulum  was  vertical ;  g  is  the  constant  of 


§  137.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        42* 
gravitation  j  I  the  length  of  the  string  AO.     Hence,  show  that 


\   _|COS0=C;   ...  G=-fcOSa 


'''di  =  ±Vt(cos^  "  C0Sa)  =  *  2VKsin2i  "  sin22)'  (2) 
since  cos  a  =  1  -  2  sin2Ja ;  cos  0  =  1  -  2  sin2J0,  and  a  is  the  value 
of  0  when  d#/d£  =  0,  that  is,  a  is  the  angle,  less  than  180°,  through 
which  the  pendulum  oscillates  on  each  side  of  the  vertical.  Since 
0  is  always  less  than  a,  we  retain  the  negative  sign. 

The  period  of  an  oscillation,  or  double  swing,  T,  can,  therefore 
be  obtained  from  (2).     We  have J 


ivm: 


"y  .       (3) 

Vl  -  sin2£a  .  sin>2<£ ' 
since  to  pass  from  0  to  \T,  0  increases  from  0  to  a,  and  <f>  from  0 
to  \ir.     Hence,  we  may  write 

d<f> 


rvr-Jo 


Jl  -  FsinV  '         '         (4) 

The  expression  on  the  right  is  called  an  elliptic  integral  of  the 

first  elass,  and  usually  written  F(k,  <f>).     The  constant  sin  Ja  is 

called  the  modulus,  and  it  is  usually  represented  by  the  symbol  k. 

The  modulus  is  always  a  proper  fraction,  i.e.,  less  than  unity.     <£ 

is  called  the  amplitude  of  T  JgJT,  and  it  is  written  <£  =  am  s]g\l  T. 

We  can  always  transform  (2)  by  substituting  sin  |0  =  x  sin  Ja, 

where  x  is  a  proper  fraction.     By  differentiation, 

h  cos  J0  .  d$  =  sin  |a .  dx  ;  .  \  dO  =  2(1  -  8in24a .  x2)  ~ 1  /2sin  Ja .  dx. 

This  leads  to  the  normal  form  of  the  elliptic  integrals  of  the  first 

class,  namely, 

dx 

■        ■        (5) 


u-i 


-oVa  -a2)(i  -wy 

commonly  written  F(k,  x).     We  can  evaluate  these  integrals  in 


1  The  expression  on  the  right  of  (2)  can  be  put  in  a  simpler  form  by  writing 
sin  %0  =  sin  £a  .sin  <p  ;  .-.  $  cos  %6  .  dd  =  sin  Ja .  cos  <p  .  d<p. 
2  sin  £a .  cos  <p  .  d<f>  __  2  sin  fra  .  cos  <pd(f>  _      2  sin  $a  cos  <pd<p 

cos£0  \/l  -  sin2£0  Jl  -  sin^a .  cos2</>  ' 

2  sin  $a  cos  <pd<p    2  cos  <ft^</> 

de  *Jl~-  sin2$asin<fr        Jl  -  sin2£a  .  sin2<f> 

>/sin*£a T-  sin2£0  =  sin  £a\/l  -  sin2^>  =  "  cos  0 

Hence  (3)  above.     These  results,  follow  directly  from  the  statements  on  pages  611  and 
612, 


T=% 


428  HIGHEE  MATHEMATICS.  §  137. 

series  as  shown  on  Ex.  (4),  page  342.  In  this  way  we  get,  from 
(4),  for  the  period  of  oscillation, 

When  the  swing  of  the  pendulum  is  small,  the  period  of  oscillation' 
T=  2/r  sfljg  seconds.  If  the  angle  of  vibration  is  increased,  in 
the  first  approximation,  we  see  that  the  period  must  be  increased 
by  the  fraction  J(sin  Ja)2  of  itself. 

The  integral  (3)  is  obviously  a  function  of  its  upper  limit  <f>, 
and  it  therefore  expresses  T  Jg/l  as  a  function  of  <f>.  We  can 
reverse  this  and  represent  <j>  as  a  function  of  T  Jg/l.  This  gives  us 
the  so-called  elliptic  functions. 

<f>  =  am  (T  Jgjl) ;  mod  k  =  sin  \a. 

The  elliptic  functions  are  thus  related  to  the  elliptic  integrals  the 
same  as  the  trigonometrical  functions  are  related  to  the  inverse 
trigonometrical  functions,  for,  as  we  have  seen,  if 

f*      dx 

y  =        /-         2;  .*•  y  -  sin-^;  and  a;  =  siny. 

We  get,  from  (3)  and  (5), 

<£  =  am  T  sjgjl ;  x  =  sin  <f> ;  .\  x  =  sin  am  T  Jg/l, 
according  to  Jacobi's  notation,  but  which  is  now  written,  after 
Gudermann,  x  =  sn  T  Jg/l.  Similarly  the  centrifugal  force,  F,  of 
a  pendulum  bob  of  mass  m  oscillating  like  the  above-described 
pendulum,  is  written  F  —  &mg sin ^a.cnT s/g/l,  where  en  T sjgjl 
is  the  cosine  of  the  amplitude  of  T  Jg/l. 

The  elliptic  functions  bear  important  analogies  with  the  ordin- 
ary trigonometrical  functions.  The  latter  may  be  regarded  as 
special  forms  of  the  elliptic  functions  with  a  zero  modulus,  and 
there  is  a  system  of  formulas  connecting  the  elliptic  functions  to 
each  other.  Many  of  these  bear  a  formal  resemblance  to  the 
ordinary  trigonometrical  relations.     Thus, 

sdHl  +  cn%  =  1 ;  x  =  sn  u ;  en  u  =  Jl  -  x2 ;  etc. 

The  elliptic  functions  are  periodic.  The  value  of  the  period 
depends  on  the  modulus  k.  We  have  seen  that  the  period  of 
oscillation  of  the  pendulum  is  a  function  of  the  modulus.  The 
substitution  equation,  sin  \0  =  sin  Ja .  sin  <f>,  shows  how  sin  J0 
changes  as  <£  increases  uniformly  from  0  to  2tt. 


3  137.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        429 

As  <£  increases  from    0  to  Jtt,  \B  increases  to  +  \a. 

As  <f>  increases  from  Jtt  to    tt,  %6  decreases  to    0. 

As  <f>  increases  from    ir  to  'Jtt,  %6  decreases  to  -£o 

As  <f>  increases  from  ±ir  to  27r,  J0  increases  to     0. 
During  the  continuous  increase  of  <£,  therefore,  \B  moves  to  and 
fro  between  the  limits  ±  \a. 

The  rectification  of  a  great  number  of  curves  furnishes  expres- 
sions which  can  only  be  integrated  by  approximation  methods — 
say,  in  series.  The  lemniscate  and  the  hyperbola  furnish  elliptic 
integrals  of  the  first  class  which  can  only  be  evaluated  in  series. 
In  the  ellipse,  the  ratio  oFJoP*  (Fig.  22,  page  100)  is  called  the  ec- 
centricity of  the  ellipse,  the  "  e  "  of  Ex.  (3),  page  115.     Therefore, 

c  =  ae ;  but,  c2  =  a2  -  b2,  .-.  b2/a2  =  1  -  e2. 
Substitute  this  in  the  equation  of  the  ellipse  (1),  page  100.     Hence, 

/dy\2      (1  -  e2)x2 

Therefore,  the  length,  I,  of  the  arc  of  the  quadrant  of  the  ellipse  is 

This  expression  cannot  be  reduced  by  the  usual  methods  of  inte- 
gration. Its  value  can  only  be  determined  by  the  usual  methods 
of  approximation.  Equation  (7)  oan  be  put  in  a  simpler  form  by 
writing  x  =  a  sin  <£,  where  <f>  is  the  complement  of  the  "eccentric" 
angle  6  (Fig.  152).     Hence, 


fin- 

l  =  a\     sjl  -  e2am2<j>.d<f>. 
Jo 


The  right  member  is  an  elliptic  integral  of  the  second  class, 

which  is  usually  written,  for  brevity's  sake,  E(k,  </>),  since  k  is 
usually  put  in  place  of  our  e.     The  integral  may  also  be  written 


»*-w^y   ■  ■  .(8» 

by  a  suitable  substitution.  We  are  also  acquainted  with  elliptic 
integrals  of  the  third  class, 

n<n>  k>  +>  -  fcl  +  Min^l  -  fctoV  ^  "^  &'  *  *°"  (9) 

where  n  is  any  real  number,  called  Legendre's  parameter.  If  the 
limits  of  the  first  and  second  classes  of  integrals  are  1  and  0, 
instead  of  x  and  0  in  the  first  case  and  £tt  and  0  in  the  second 


430 


HIGHER  MATHEMATICS. 


§137. 


case,  the  integrals  are  said  to  be  complete.  Complete  elliptic  in- 
tegrals of  the  first  and  second  classes  are  denoted  by  the  letters  F 
and  E  respectively. 

The  integral  of  an  irrational  polynomial  of  the  second  degree, 
of  the  type, 

f    / i o    ^    -,  f         X<fa 

I  si  a  +  bx  +  ex2 .  X  .  dx ;  or,  I    ,        _      ==• 
J  J  si  a  +  bx  +  ex2 

(where  X  is  a  rational  function  of  x),  can  be  made  to  depend  on 

algebraic,  logarithmic,  or  on  trigonometrical  functions,  which  can 

be  evaluated  in  the  usual  way.     But  if  the  irrational  polynomial 

is  of  the  third,  or  fourth  degree,  as,  for  example, 


I  J  a  +  bx  +  ex2  +  dxz  +  ex*Xdx ; 

the  integration  cannot  be  performed  in  so  simple  a  manner.  Such 
integrals  are  also  called  elliptic  integrals.  If  higher  powers  than 
x*  appear  under  the  radical  sign,  the  resulting  integrals  are  said  to 
be  ultra- elliptic  or  hyper-elliptic  integrals.  That  part  of  an 
elliptic  integral  which  cannot  be  expressed  in  terms  of  algebraic, 
logarithmic,  or  trigonometrical  functions  is  always  one  of  the  three 
classes  just  mentioned. 

Legendre  has  calculated  short  tables  of  the  first  and  second 
class  of  elliptic  integrals  ;  the  third  class  can  be  connected  with 
these  by  known  formulae.  Given  k  and  x,  F(k,  <f>)  or  E(k}  <f>)  can 
be  read  off  directly  from  the  tables.  The  following  excerpt  will 
give  an  idea  of  how  the  tables  run : 

Numerical  Values  of  F(k,  <f>) ;  sin  a  =  k. 


</>. 

a   -0°. 

a  =  6°. 

a  -  10°. 

a  =  16°. 

a   =  20°. 

a  =  26°. 

41° 
42° 
43° 

0-7156 
0-7330 
0-7505 

0-7160 
0-7335 
0-7510 

0-7173 
0-7348 
0-7524 

0-7193 
0-7370 
0-7548 

0-7222 
0-7401 
0-7681 

0-7258 
0-7440 
0-7622 

f     dx        f 
Example. — Show  that  /    ,  .     -  =  I  —f 
J  vsin  x    J  vi 

\/2  .  sin  \x  =  sin  <p 


dx 


sMJ 


n/1  -\ 


'cos  a;  J  vi  -  3  snrty 

Hint.  The  first  step  follows  from  (6),  page  242. 
by  differentiation,  cos  \xdx  =  s/2  .  cos  <t>d^> ;  2  sin2^*  =  sinfy.    Hence, 
dx  f    \/2".  cos  \x .  dx         f  s/2.  co 

cos  x      J 


,  provided 

Next 


/ 


■/ 


\/2~.  cos  %x  vcosa;      J  \/l  -  sin2$a;  <s/l  -  2  sin2^x' 


etc. 


from  (20)  and  (35),  page  612. 


§  138.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        431 

We  cannot  spare  space  to  go  farther  into  this  matter.  Mascart 
and  Joubert  have  tables  of  the  coefficient  of  mutual  induction  of 
electric  currents,  in  their  Electricity  and  Magnetism  (2,  126,  1888), 
calculated  from  E  and  F  above.  A.  G.  Greenhill's  The  Applica- 
tions of  Elliptic  Functions  (London,  1892),  is  the  text-book  on 
this  subject. 

§  138.    The  Exaot  Linear  Differential  Equation. 

A  very  simple  relation  exists  between  the  coefficients  of  an 
exact  differential  equation  which  may  be  used  to  test  whether  the 
equation  is  exact  or  not.     Take  the  equation, 

x»%+xm  +  x>%  +  x<y-x>  ■     ■    « 

where  X0,  X^,  . . . ,  B  are  functions  of  x.  Let  their  successive 
differential  coefficients  be  indicated  by  dashes,  thus  X,  X",  . . . 

Since  X0 .  d3y/dxz  has  been  obtained  by  the  differentiation  of 
X0 .  d2y/dx2}  this  latter  is  necessarily  the  first  term  of  the  integral 
of  (1).     But, 

dx\*°dxy  ~  A w  +  A  «W 

Subtract  the  right-hand  side  of  this  equation  from  (1). 

(*»  -  *.)§  +  *ffi  +  ■**  -  *       •        •       (2) 

Again,  the  first  term  of  this  expression  is  a  derivative  of 
(Xx  -  X0)dy/dx.  This,  therefore,  is  the  second  term  of  the  in- 
tegral of  (1).     Hence,  by  differentiation  and  subtraction,  as  before, 

(Xt  -  X\  +  X"0)g  +  X#  -  B.        .        .      (3) 

This  equation  may  be  deduoed  by  the  differentiation  of 
(X2  -  X\  +  X"0)y,  provided  the  first  differential  coefficient  of 
(X2  -  X\  +X"{)  with  respeot  to  x,  is  equal  to  X8,  that  is  to  say, 
X2  -  X'\  +  X"0  -  X8 ;  or,  X,  -  X2  +  Xf\  -  X"0  -  0.  (4) 
But  if  this  is  really  the  origin  of  (3),  the  original  equation  (1)  has 
been  reduced  to  a  lower  order,  namely, 

Xo§  +  (*i  -  ^o)|-  +  (*,  -  *\  +  *\yy  =  \Bdx  +  Or    (5) 
This  equation  is  called  the  first  integral  of  (1),  because  the  order 


432  HIGHER  MATHEMATICS.  §  138. 

of  the  original  equation  has  been  lowered  unity,  by  a  process  of 
integration.  Condition  (4)  is  a  test  of  the  exactness  of  a  differential 
equation. 

If  the  first  integral  is  an  exact  equation,  we  can  reduce  it,  in 
the  same  way,  to  another  first  integral  of  (1).  The  process  of 
reduction  may  be  repeated  until  an  inexact  equation  appears,  or 
until  y  itself  is  obtained.  Hence,  an  exact  equation  of  the  nth 
order  has  n  independent  first  integrals. 

Examples.— (1)  Is  a5 .  d3y[dx?  +  15a4 .  dPy/dx*  +  60a3 .  dyjdx  +  60x*y  =  e* 
an  exact  equation?  From  (4),  X3  =  60a;2;  X'2  =  180a2;  X'\  =  180a2; 
X"\  =  60a2.  Therefore,  X%  -  X'2  +  X'\  -  X"\  =  0  and  the  equation  is 
exact.  Solve  the  given  equation.  Ansr.  x5y  =  ex  +  Oxx2  +  G%x  +  C3. 
Hints.  From  (5),  the  first  integral  is  (a5!)2  +  10a4D  +  2003)3/  =  ex  +  Cv 
This  is  exact,  because  the  new  values  of  X  for  the  first  integral  just  obtained 
X2  -  X\  +  X"0  =  0,  since,  20a3  -  40a3  +  20a3  -  0.  For  the  next  first  in- 
tegral, we  have 

X<^  +  (xi-x'o)y  =  fexdx  +  fc^  +  C,.       .        .        (6) 

Hence  (x?>D  +  5x4)y  =  ex  +  Gxx  +  C2.  This  is  exact,  because  the  new  values 
of  X,  namely,  Xx  -  X0  =  0.  Hence,  the  third  and  last  first  integral  is 
aPy  =  jexdx  +  jCjXdx  +  jC2dx  +  C3,  etc. 

(2)  Solve  xd^yjdx3  +  (a2  -  S)d2y/dx2  +  4a .  dyjdx  +  2y  =  0,  as  far  as  pos- 
sible, by  successive  reduction.  The  process  can  be  employed  twice,  the 
residue  is  a  linear  equation  of  the  first  order,  not  exact.  Complete  solution  : 
x-66hhj=  Gx\x~H^dx  +  C2jx-*ehx2dx  +  03. 

There  is  another  quick  practical  test  for  exact  differential  equa- 
tions (Forsyth)  which  is  not  so  general  as  the  preceding.  When 
the  terms  in  X  are  either  in  the  form  of  axm,  or  of  the  sum  of 
expressions  of  this  type,  xmdny/dxn  is  a  perfect  differential  co- 
efficient, if  m  <n.  This  coefficient  can  then  be  integrated  what- 
ever be  the  value  of  y.  If  m  =  n  or  m  >  n,  the  integration  cannot 
be  performed  by  the  method  for  exact  equations.  To  apply  the 
test,  remove  all  the  terms  in  which  m  is  less  than  n,  if  the  re- 
mainder is  a  perfect  differential  coefficient,  the  equation  is  exact 
and  the  integration  may  be  performed. 

Examples.— (1)  Test  a3 .  d^jdx4  +  a2  .  dPyjdx*  +  a  .  dyjdx  +  y  =  0. 
a .  dyjdx  +  y  remains.  This  has  evidently  been  formed  by  the  operation 
D(xy),  hence  the  equation  is  a  perfect  differential. 

(2)  Apply  the  test  to  (a3D4  +  a2Z>3  +a2D  +  2x)y  =  sin  a.  a2,  dyjdx  +  2xy 
remains.  This  is  a  perfect  differential,  formed  from  Dfa^y).  The  equation 
is  exact. 


§  139.      HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.       433 

§  139.  The  Velocity  of  Consecutive  Chemical  Reactions. 

While  investigating  the  rate  of  decomposition  of  phosphine, 
page  224,  we  had  occasion  to  point  out  that  the  aotion  may  take 
place  in  two  stages  : — 

Stage  I.  PH3  =  P  +  3H.     Stage  II.  4P  =  P4 j  2H  =  H2. 

The  former  change  alone  determines  the  velocity  of  the  whole 
reaction.  The  physical  meaning  of  this  is  that  the  speed  of  the 
reaction  which  occurs  during  the  second  stage,  is  immeasurably 
faster  than  the  speed  of  the  first.  Experiment  quite  fails  to  reveal 
the  complex  nature  of  the  complete  reaction.  J.  Walker  illustrates 
this  by  the  following  analogy  (Proc.  Boyal  Soc.  Edin.,  22,  22, 
1898) :  "  The  time  occupied  in  the  transmission  of  a  telegraphic 
message  depends  both  on  the  rate  of  transmission  along  the  conduct- 
ing wire  and  on  the  rate  of  the  messenger  who  delivers  the  telegram ; 
but  it  is  obviously  this  last,  slower  rate  that  is  of  really  practical 
importance  in  determining  the  total  time  of  transmission". 

Suppose,  for  example,  a  substance  A  forms  an  intermediate 
compound  M,  and  this,  in  turn,  forms  a  final  product  B.  If  the 
speed  of  the  reaction  A  =  M,  is  one  gram  per  tW&w  second,  when 
the  speed  of  the  reaction  M  =  B,  is  one  gram  per  hour,  the  ob- 
served "order  "  of  the  complete  reaction 

A  =  B, 

will  be  fixed  by  that  of  the  slower  reaction,  M  =  B,  because  the 
methods  used  for  measuring  the  rates  of  chemical  reactions  are  not 
sensitive  to  changes  so  rapid  as  the  assumed  rate  of  transformation 
of  A  into  M.  Whatever  the  "  order  "  of  this  latter  reaction,  M  =  B 
is  alone  accessible  to  measurement.  If,  therefore,  A  =  B  is  of  the 
first,  second,  or  nth  order,  we  must  understand  that  one  of  the 
subsidiary  reactions :  A  =  M,  or  M  =  B,  is  (i)  an  immeasurably 
fast  reaction,  accompanied  by  (ii)  a  slower  measurable  change  of 
the  first,  second  or  nth  order,  according  to  the  particular  system 
under  investigation. 

If,  however,  the  velocities  of  the  two  reactions  are  of  the  same 
order  of  magnitude,  the  "  order  "  of  the  complete  reaction  will  not 
fall  under  any  of  the  simple  types  discussed  on  page  218,  and 
therefore  some  changes  will  have  to  be  made  in  the  differential 
equations  representing  the  course  of  the  reaction.  Let  us  study 
some  examples. 

BE 


434  H1GHEK  MATHEMATICS.  §  139. 

I.  Two  consecutive  unimolecular  reactions. 

Let  one  gram  molecule  of  the  substance  A  be  taken.  At  the 
end  of  a  certain  time  tf  the  system  contains  x  of  A,  y  of  M,  z  of  B. 
The  reactions  are 

A  =  M;  M  =  B. 

The  rate  of  diminution  of  x  is  evidently 

dx 
-Tt  =  kYx,  (1) 

where  kx  denotes  the  velocity  constant  of  the  transformation  of 
A  to  M.    The  rate  of  formation  of  B  is 

dz 

di  =  k# (2) 

where  k2  is  the  velocity  constant  of  the  transformation  of  M  to  B- 
Again,  the  rate  at  which  M  accumulates  in  the  system  is  evidently 
the  difference  in  the  rate  of  diminution  of  x  and  the  rate  of  increase 
of  z,  or 

■£  =  kix-  KV-  ...         (3) 

The  speed  of  the  chemical  reactions, 
A  =  M  =  B, 
is  fully  determined  by  this  set  of  differential  equations.    When  the 
relations  between  a  set  of  variables  involves  a  set  of  equations  of 
this  nature,  the  result  is  said  to  be  a  system  of  simultaneous 
differential  equations. 

In  a  great  number  of  physical  problems,  the  interrelations  of 
the  variables  are  represented  in  the  form  of  a  system  of  such 
equations.  The  simplest  class  occurs  when  each  of  the  dependent 
variables  is  a  function  of  the  independent  variable. 

The  simultaneous  equations  are  said  to  be  solved  when  each 
variable  is  expressed  in  terms  of  the  independent  variable,  or  else 
when  a  number  of  equations  between  the  different  variables  can  be 
obtained  free  from  differential  coefficients.  To  solve  the  present 
set  of  differential  equations,  first  differentiate  (2), 

dt*    **dt  ~  \ 

Add  and  subtract  kjc2yt  substitute  for  dy/dt  from  (3)  and  for  k$ 
from  (2),  we  thus  obtain 

gl  +  (&!  +  k2)^  -  lcjc^x  +  y)  =  0. 


§  139.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        435 


But  from  the  conditions  of  the  experiment, 

x  +  y  +  z  =  1,  .'.  z  -  1  <=  -  (x  +  y). 
Hence,  the  last  equation  may  be  written, 


d\z 


dt2 


-  +  &  +  *»>^3T^  +  *A<*  -  1)  «  0. 


(4) 


This  linear  equation  of  the  second  order  with  constant  coefficients 
is  to  be  solved  for  z  -  1  in  the  usual  manner  (§  131).  At  sight 
therefore, 

z  -  1  =  Cx« ~ *i*  +  C2e-*2*. 

But  z  =  0,  when  t  =  0, 

.-.  Gl  +  G2  -  -  1. 
Differentiate  (5).     From  (2)  dz/dt  =  0,  when  t  =  0. 
making  the  necessary  substitutions, 

-  Gfa  -  G2h2  =  0. 
From  (6)  and  (7), 

Ci -*./(*i  -*»);  0%  =  "  V(*i  -  h)> 
The  final  result  may  therefore  be  written, 

#1  -  -  k^i  & 


.        (5) 

•        (6) 
Therefore 

.         (7) 


1  = 


_  "*2*  - 

to\    """   ™2  1   ~"      2 


e  "  *i'. 


(8) 


.  Harcourt  and  Esson   have   studied  the  rate  of  reduction  of 
potassium  permanganate  by  oxalic  acid. 

2KMn04  +  3MnS04  +  2H20  =  K2S04  +  2H2S04  +  5Mn02 ; 

Mn02  +  H2S04  +  H2C204  =  MnS04  +  2H20  +  2C02. 
By  a  suitable  arrangement  of  the  experimental  conditions  this 
reaction  may  be  used  to  test  equations  (5)  or  (8). 

Let  xy  y,  z}  respectively  denote  the  amounts  of  Mn207,  Mn02 
and  MnO  (in  combination)  in  the  system.  The  above  workers 
found  that  G1  =  28-5;  C2  =  27;  <?*i  =  -82 ;  e~kz  =  -98.  The 
following  table  places  the  above  suppositions  beyond  doubt. 


t 
Minutes. 

•  -1. 

t 
Minutes. 

«-l. 

Found. 

Calculated. 

Found. 

Calculated. 

0-5 
10 
1-5 
20 
2-5 

25-85 

21-55 

17-9 

14-9 

12-55 

25-9 
21-4 
17-8 
14-9 
12-5 

3-0 
3-5 
4-0 
4-5 
5-0 

10-45 

8-95 
7-7 
6-65 
5-7 

10-4 
9-0 
7-8 
6-6 

5-8 

Example. — We  could  have  deduced  equation  (8)  by  another  line  of  reason- 
ing.   If  x  denotes  the  amount  of  A  transformed  into  M  in  the  time  t,  and  z 

EE* 


436  HIGHER  MATHEMATICS.  §  139. 

the  amount  of  M  transformed  into  B  at  the  time  t,  then,  if  a  denotes  the 
amount  of  A  present  at  the  beginning  of  the  experiment, 

dy  dz 

dt=ki(a-y);  %  =  *&-*)-        t      •      •     P) 

Prom  the  first  equation,  y  *=  a(l  -  e-h*).    Substitute  this  result  in  the  second 

equation,  and  we  get 

dz 

-fa  +  ktf  -  k,a(l  -  e-V)  -  0. 

Prom  Ex.  (5),  page  388,  if  2  =  0,  when  t  =  0,  we  get 

z,c,-v+a.^.V;0=^__o;      .   (10) 

and  we  get,  finally,  an  expression  resembling  (8)  above.  Equation  (8)  has 
also  been  employed  to  represent  the  decay  of  the  radioactivity  excited  in 
bodies  exposed  to  radium  and  to  thorium  emanation. 

II.  Two  bimolecular  consecutive  reactions. 

During  the  saponification  of   ethyl  succinate  in  the  presence 
of  sodium  hydroxide. 

C2H4(COOC2H5)2  +  NaOH  =  C2H5OH  +  02H4 .  COONa .  COOC2H5 ; 
C2H4.COONa.COOH  +  NaOH  =  C2H6OH  +  C2H4(COONa)2 . 
Or, 

A  +  B  =  C  +  M ;  M  +  B  =  C  +  D. 
Let  x  denote  the  amount  of  ethyl  succinate,  A,  which  has  been 
transformed  at  the  time  t;  a  -  x  will  then  denote  the  amount 
remaining  in  the  solution  at  the  same  time.  Similarly,  if  the 
system  contains  b  of  sodium  hydroxide,  B,  at  the  beginning  of 
the  reaction,  at  the  time  t,  x  of  this  will  have  been  consumed  in 
the  formation  of  sodium  ethyl  succinate,  M,  and  y  in  the  formation 
of  sodium  succinate,  D,  hence  b  -  x  -  y  of  sodium  hydroxide,  B, 
and  x  -  y  of  sodium  ethyl  succinate,  M,  will  be  present  in  the 
system  at  the  time  t.  The  rate  of  formation  of  sodium  ethyl  suc- 
cinate, M,  is  therefore 

(ii) 

(12) 


dx 

Tt  -  k^a 

-•)<&■ 

-  *  -  y) ; 

. 

and  the  rate  of  formation  of  sodium  succinate,  D, 

will  be 

Tt  =  *** 

-y)  (b 

-x-y). 

. 

By  division, 

if  h2lkx  =  K, 

dy 

dx      a 

K 

-xy- 

Kx 
a  -  x 

§  139.  HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.  437 
This  equation  has  been  integrated  in  Ex.  (6),  p.  388.     Hence 

y  -  tt$r=f  -  * +  >}•'  •     •     (13) 

dx      ,  .  ,  f ,  (a-x)K  a  Ex    \    „  A. 

•••  *~  k^a  ~  X\b  ~  X  ~  {K-  l)a'-i  +  K^x  -  KZl)-  <14> 
This  can  only  be  integrated  when  we  know  the  numerical  value  of 
E.  As  a  rule,  in  dealing  with  laboratory  measurements,  it  will  be 
found  most  convenient  to  use  the  methods  for  approximate  in- 
tegration since  the  integration  of  (14)  is  usually  impracticable,  even 
when  we  know  the  value  of  E. 

III.  A  unimolecular  reaction  followed  by  a  bimolecular  reaction. 

Let  x  denote  the  amount  of  A  which  remains  untransformed 
after  the  elapse  of  an  interval  of  time  t,  y  the  amount  of  M,  and  z 
the  amount  of  B  present  in  the  system  after  the  elapse  of  the  same 
interval  of  time  t.     The  reaction  is 

A  =  M ;  M  +  B  =  C. 
Hence  show  that  the  rate  of  diminution  of  A,  and  the  rate  of 
diminution  of  M  (or  of  B)  are  respectively 
dx  dz 

-  m~kix;  -  dt  =  k*y*>      -     -    (15> 

the  rate  of  formation  of  M2  is  the  difference  between  the  rate  of 
formation  of  M  by  the  reaction,  and  the  velocity  of  transformation 

of  M  into  C,  by  the  second  reaction  and 

dy 
.'.jt  =  kxx-k2yz.         .         .         .       (16) 

If  x,  y,  z,  could  be  measured  independently,  it  would  be  sufficient 
to  solve  these  equations  as  in  I,  but  if  x  and  y  are  determined 
together,  we  must  proceed  a  little  differently.  If  there  are  a 
equivalents  of  A,  and  of  B  originally  present,  then,  at  the  time  t 
we  shall  have  a-x=a-z  +  y,  or  y  =  z  -  x.  Divide  (16)  by 
the  first  of  equations  (15) ;  substitute  dy  =  dz  -  dx  An  the  result ; 
put  y  =  z  -  x ;  divide  by  z2,  and  we  get 

1    dz      E      E      „ 

^Tx  +  l~x=°>  '         *       <17> 

where  iT  has  been  written  in  place  of  kjkv  The  solution  of  this  equa- 
tion has  been  previously  determined,  Ex.  (3),  page  389,  in  the  form 

Ee-*4cx  -  log  x  +  Ex  -  ^(Exf  +  ...}*  =  1.      (18) 
In  some  of  Harcourt  and  Esson's  experiments,  Cx  =  4-68 ;  &j  =  -69  > 


438 


HIGHER  MATHEMATICS. 


§139. 


k2  =  -006364.  From  the  first  of  equation  (9),  it  is  easy  to  show 
that  x  =  ae  ~  V.  Where  does  a  come  from  ?  What  does  it  mean  ? 
Obviously,  the  value  of  x  when  t  =  0.  Hence  verify  the  third 
column  in  the  following  table  : — 


t 
Minutes. 

X. 

Found. 

Calculated. 

2 
3 
4 
5 

51-9 

42-4 
35-4 
29-8 

51-6 
42-9 
35-4 
29-7 

After  the  lapse  of  six  minutes,  the  value  of  x  was  found  to  be 
negligibly  small.  The  terms  succeeding  log  x  in  (18)  may,  there- 
fore, be  omitted  without  committing  any  sensible  error.  Substi- 
tute x  =  ae~klt  in  the  remainder, 


h  1 

j?(Ci  -  log  a  +  kxt)z  =  1 ;  or  (ff1+  t)z  =  p 

where  G\  =  GJ^  -  (log  a)jhv  Harcourt  and  Esson  found  that 
C\  =  O'l,  and  l/k2  =  157.  Hence,  in  continuation  of  the  preceding 
table,  these  investigators  obtained  the  results  shown  in  the  follow- 
ing table.  The  agreement  between  the  theoretical  and  experimental 
numbers  is  remarkable. 


f 

z 

t 

Minutes. 

t 
Minutes. 

Found. 

Calculated. 

Found. 

Calculated. 

6 

25*7 

25-7 

10 

15-5 

15-5 

7 

22-1 

22-1 

15 

10-4 

10-4 

8 

19-4 

19-4 

20 

7*8 

7'8 

9 

17-3 

17*3 

30 

5-5 

5-2 

The  theoretical  numbers  are  based  on  the  assumption  that  the 
chemical  change  consists  in  the  gradual  formation  of  a  substance 
which  at  the  same  time  slowly  disappears  by  reason  of  its  reaction 
with  a  proportional  quantity  of  another  substance. 

This  really  means  that  the  so-called  u  initial  disturbances  "  in 
chemical  reactions,  are  due  to  the  fact  that  the  speed  during  one 
stage  of  the  reaction,  is  faster  than  during  the  other.  The  magni- 
tude of  the  initial  disturbances  depends  on  the  relative  magnitudes 


§  139.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS         439 

of  kx  and  k2.  The  observed  velooity  in  the  steady  state  depends 
on  the  difference  between  the  steady  diminution  -  dx/dt  and  the 
steady  rise  dz/dt.  If  k2  is  infinitely  great  in  comparison  with  klf 
(8)  reduces  to 

*  =  a(l  -  e  -*!'), 

which  will  be  immediately  recognised  as  another  way  of  writing 
the  familiar  equation 

h,  =  -  log . 

1       t     5  a  -  z 

So  far  as  practical  work  is  concerned,  it  is  necessary  thafc  the 
solutions  of  the  differential  equations  shall  not  be  so  complex  as  to 
preclude  the  possibility  of  experimental  verification. 

IV.  Three  consecutive  bimolecular  reactions. 

In  the  hydrolysis  of  triacetin, 

C3H5 .  A3  +  H  .  OH  =  3A  .  H  +  03H5(OH)3, 

where  A  has  been  written  for  CH3 .  COO,  there  is  every  reason 
to  believe  that  the  reaction  takes  place  in  three  stages  . 

C8H5  .  A3  +  H  .  OH  =  A  .  H  +  C3H5  .  A2  .  OH  (Diacetin)  ; 

C3H5  .  J2  .  OH  +  H  .  OH  =  A  .  H  +  C3H6 .  A(OH)2    (Monoacetin) ; 

C8H5  A  .  (OH)2  +  H  .  OH  =  A  .  H  +  C3H5(OH)3  (Glycerol). 
These  reactions  are  interdependent.  The  rate  of  formation  of 
glycerol  is  conditioned  by  the  rate  of  formation  of  monoacetin  ;  the 
rate  of  monoacetin  depends,  in  turn,  upon  the  rate  of  formation  of 
diacetin.  There  are,  thei  efore,  three  simultaneous  reactions  of  the 
second  order  taking  place  in  the  system. 

Let  a  denote  the  initial  concentration  (gram  molecules  per 
unit  volume)  of  triacetin,  b  the  concentration  of  the  water ;  let  x, 
y,  z,  denote  the  number  of  molecules  of  mono,-  di-  and  triacetin 
hydrolyzed  at  the  end  of  t  minutes.  The  system  then  contains 
a  -  z  molecules  of  triacetin,  z  -y,  of  diacetin,  y  -x,oi  monacetin, 
and  b  -  (x  +  y  +  z)  molecules  of  water.  The  rate  of  hydrolysis 
is  therefore  completely  determined  by  the  equations : 

dx/dt  =  hx(y  -  x)  (b  -  x  -  y  -  z) ;     .         .       (19) 

dy/dt  =  k2(z  -y)  (b-x-y  -  z);     .         .       (20) 

dz/dt  =  k3(a  -  z)  (b  -  x  -  y  -  z)  ■     .        .       (21) 

where  kv  k2,  kz,  represent  the  velocity  coefficients  (page  63)  of  the 

respective  reactions. 


440  HIGHER  MATHEMATICS.  §  139. 

Geitel  tested  the  assumption:  kx  =  k2  =  ks.  Hence  dividing 
(21)  by  (19)  and  by  (20),  he  obtained 

dz/dy  =  (a  -  z)/(z  -  y) ;  dz/dx  =  (a  -  z)/(y  -  x).         (22) 
From  the  first  of  these  equations, 

dy  1  z 

dz       ya  -  z~  a  -  z* 

which  can  be  integrated  as  a  linear  equation  of  the  first  order. 
The  constant  is  equated  by  noting  that  if  a  =  1,  z  =  0,  y  =  0. 
The  reader  might  do  this  as  an  exercise  on  §  125.     The  answer  is 

y  =  z  +  (a  -  z)\og(a  -  *).  .         .       (23) 

Now  substitute  (23)  in  the  second  of  equations  (22),  rearrange 
terms  and  integrate  as  a  further  exercise  on  linear  equations  of  the 
first  order.     The  final  result  is, 

x  =-  z  +  (a  -  z)  log  (a  -  z)  -  ^-^{log  (a  -  z)}*.        (24) 

z 

Geitel  then  assigned  arbitrary  numerical  values  to  z  (say  from 
0*1  to  1*0),  calculated  the  corresponding  amounts  of  x  and  y  from 
(23)  and  (24)  and  compared  the  results  with  his  experimental 
numbers.  For  experimental  and  other  details  the  original  memoir 
must  be  consulted. 

A  study  of  the  differential  equations  representing  the  mutual 
conversion  of  red  into  yellow,  and  yellow  into  red  phosphorus, 
will  be  found  in  a  paper  by  G.  Lemoine  in  the  Ann.  Ghim.  Phys. 
[4],  27,  289,  1872.  There  is  a  series  of  papers  by  R.  Wegscheider 
bearing  on  this  subject  in  Monats.  Ghemie,  22,  749,  1901 ;  Zeit. 
phys.  Chem.,  30,  593,  1899;  34,  290,  1900;  35,  513,  1900;  J. 
Wogrinz,  ib.t  44,  569,  1903 ;  H.  Kuhl,  ib.,  44,  385,  1903.  See 
also  papers  by  A.  V.  Harcourt  and  W.  Esson,  Phil.  Trans.,  156, 
193,  1866;  A.  0.  Geitel,  Journ.  prakt.  Chem.  [2],  55,  429,  1897; 
57,  113,  1898 ;  J.  Walker,  Proc.  Boy.  Soc.  Edin.,  22,  22,  1898. 
It  is  somewhat  surprising  that  Harcourt  and  Esson's  investiga- 
tions had  not  received  more  attention  from  the  point  of  view  of 
simultaneous  and  dependent  reactions.  The  indispensable  differ- 
ential equations,  simple  as  they  are,  might  perhaps  account  for 
this.  But  chemists,  in  reality,  have  more  to  do  with  this  type  of 
reaction  than  any  other.  The  day  is  surely  past  when  the  study 
of  a  particular  reaction  is  abandoned  simply  because  it  "  won't  go  " 
according  to  the  stereotyped  velocity  equations  of  §  77. 


§  140.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        441 

§  1*0.  Simultaneous  Equations  with  Constant  Coefficients. 

By  way  of  practice  it  will  be  convenient  to  study  a  few  more 
examples  of  simultaneous  equations,  since  they  are  so  common  in 
many  branches  of  physics.  The  motion  of  a  particle  in  space  is 
determined  by  a  set  of  three  differential  equations  which  determine 
the  position  of  the  moving  particle  at  any  instant  of  time.  Thus, 
if  X,  Y,  Zf  represent  the  three  components  of  a  force,  F,  acting  on 
a  particle  of  mass  ra,  Newton's  law,  page  396,  tells  us  that 

„         d2x     .,         d2y      _         d*z 

and  it  is  necessary  to  integrate  these  equations  in  order  to  represent 
x,  y,  z  as  functions  of  the  time  t.  The  solution  of  this  set  of 
equations  contains  six  arbitrary  constants  which  define  the  position 
and  velocity  of  the  moving  body  with  respect  to  the  x-}  y-,  and  the 
s-axis  when  we  began  to  take  its  motion  into  consideration. 

In  order  to  solve  a  set  of  simultaneous  equations,  there  must 
be  the  same  number  of  equations  as  there  are  independent  variables. 
Quite  an  analogous  thing  occurs  with  the  simultaneous  equations 
in  ordinary  algebra.  The  methods  used  for  the  solution  of  these 
equations  are  analogous  to  those  employed  for  similar  equations  in 
algebra.  The  operations  here  involved  are  chiefly  processes  of 
elimination  and  substitution,  supplemented  by  differentiation  or 
integration  at  various  stages  of  the  computation.  The  use  of  the 
symbol  of  operation  D  often  shortens  the  work. 

Examples.— (1)  Solve  dxjdt  +  ay  =  0,  dy/dt  +  bx  m  0.  Differentiate  the 
first,  multiply  the  second  by  a.  Subtract  and  y  disappears.  Hence  writing 
ab  =  w2,  x  =  C^e"*  +  C^-™*;  or,  y  =  C2  n/o/o" .  «  ~  "*  -  O^bja.e^.  We 
might  have  obtained  an  equation  in  y,  and  substituted  it  in  the  second. 
Thus  four  constants  appear  in  the  result.  But  one  pair  of  these  constants 
can  be  expressed  in  terms  of  the  other  two.  Thus:  two  of  the  constants, 
therefore,  are  not  arbitrary  and  independent,  while  the  integration  constant 
is  arbitrary  and  independent.  It  is  always  best  to  avoid  an  unnecessary 
multiplication  of  constants  by  deducing  the  other  variables  from  the  first 
without  integration.  The  number  of  arbitrary  constants  is  always  equal 
to  the  sum  of  the  highest  orders  of  the  set  of  differential  equations  under 
consideration. 

(2)  Solve  dx/dt  +  y  =  3x ;  dy/dt  -  y  =  x.  Differentiate  the  first.  Sub- 
tract each  of  the  given  equations  from  the  result.  (D2  -  4D  +  4)a;=0  remains. 
Solve  as  usual,  x  =  {Gx  +  CzfyeP.  Substitute  this  value  of  x  in  the  first  of 
the  given  equations  and  y  =  {Cx  -  02  +  C2£)e2'. 

(3)  The  rotation  of  a  particle  in  a  rigid  plane,  is  represented  by  the  equa- 


442  HIGHER  MATHEMATICS.  §  140. 

tions  dxjdt  =  fxy ;  dy/dt  =  /xx.  To  solve  these,  differentiate  the  first,  multiply 
the  second  by  /*,  etc.  Finally  x  =  G1  cos  fit  +  C2  sin  /d;y  =  G\cos  fit  +  G'2  sin  /nt. 
To  find  the  relation  between  these  constants,  substitute  these  values  in  the 
first  equation  and  -  /xGx  sin  fd  +  fiG2  cos  fit  =  fiG\  cos  fit  +  fiC'2  sin  fit,  or 
C2  =  -  G'2  and  C2  =  C\. 

(4)  Solve  d?x/dt2=  -n*x  d2y]dt2=  -r&y.  Each  equation  is  treated  separ- 
ately as  on  page  400,  thus  x  =  Gl  cos  nt  +  C2  sin  nt;  y  =  G\  cos  nt  +  C2  sin  ni. 
Eliminate  t  so  that 

(0V*  -  Gxy)2  +  (O'rfB  -  O^)2  =  (C^  -  C^'J2,  etc. 
The  result  represents  the  motion  of  a  particle  in  an  elliptic  path,  subject  to  a 
central  gravitational  force. 

(5)  Solve  dy/dx  +  Sy  -  4s  =  5e5*  ;  dz\dx  +  y  -  2x  =  -  3e5*.  Differentiate 
the  first  and  solve  for  dzjdx ;  substitute  this  value  of  dzjdx  in  the  second 
equation.     We  thus  get  a  linear  equation  of  the  second  order : 

dhi      dv 

d^  +  Tx-2y  =  3tfi**  •'•  y  =  G*x  +  °2*~2* +  &*''*> 

when  solved  by  the  usual  method.  Now  differentiate  the  last  equation,  and 
substitute  the  value  of  dy/dx  so  found  in  the  first  of  the  given  equations. 
Also  substitute  the  value  of  y  just  determined  in  the  same  equation.  We 
thus  get  z  m  Gxe*  +  i<V_2a5  -  ffe8*. 

(6)  R  Wegscheider  (Zeit.  phys.  Ghem.,  41,  52,  1902)  has  proposed  the 
equations  dx/dt  =  kjp  -  x  -  y)  \  dyjdt  =  k2(a  -  x  -  y)  (b  -  y),  to  represent 
the  speed  of  hydrolysis  of  sulphonic  esters  by  water.     Hence  show  that 

a  -  (1  +  bk)x  +  $bKW  =  M  +  <?• 
Hint.  Divide  the  one  equation  by  the  other ;  expand  e  ~ Kx  ;  reject  all  but 
the  first  three  terms  of  the  series. 

(7)  J.  W.  Mellor  and  L.  Bradshaw  {Zeit.  phys.  Ghem.,  48,  353,  1904) 

solved  the  set  of  equations 

dX      7  ,        ~    du      ,  .  .     dv      .  . 

-3r=  k^a  -  X);  ^  =  fc2(a;  -  «) ;  ^  =  fc3(*/  -  v) 

with  the  assumption  that  X  =  x  +  y  +  u  +  v;  v  =  kAu ;  w  =  v  =  a;  =  j/  =  0 
when  t  =  0.     Show  that 

u    2(k4  + 1)  V     &  -  *,       +  r^r    /' 

if  6  is  put  in  place  of  2k2k5(k4  +  l)/(&2  +  k3k4).     See  Ex.  (5),  page  390. 

(8)  J.  J.  Thomson  (Conduction  of  Electricity  through  Gases,  Cambridge, 
86,  1903)  has  shown  that  the  motion  of  a  charged  particle  of  mass  m,  and 
charge  e,  between  two  parallel  plates  with  a  potential  gradient  E  between 
them  when  a  magnetic  field  of  induction  H  is  applied  normal  to  the  plates, 
is  given  by  the  equations 

«£-*-*!;  «3-*&     •     •     •     W 

provided  that  there  is  no  resisting  medium  (say  air)  between  the  plates.     To 

solve  these  equations  with  the  initial  conditions  that  x,  y,  x,  y,  are  all  zero, 

put 

dx         dp  dy    dhj      ,  d2p 

p  =  dt>  '''dt=a-bdt;  w=bP>  -'-w=-b^ 


h 


§  140.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        443 

Hence,  from  page  405,  and  remembering  that  *  =  0,  when  t  =  0, 

dx  C,  v 

p  =  C^sm  bt  +  02cos  bt ;  02  =  0  ;  -^  =  Cjsin  6i ;  ,\  «  as  -y  (1  -  cos  M),    (2) 

since,  when  t  =  0,  x  =  0,  and  the  integration  constant  is  equal  to  GJb.    From 

the  third  of  equations  (2),  and  the  second  of  equations  (1), 

dhj  dv 

fip  =  fcCjsin  bt;  .:  ^  =  -  Ojcos  bt  +  Gv 

the  integration  constant  is  equal  to  Cx  when  dyjdt  =  0,  and  t  =  0 ;  again 
integrating,  and  we  get  y  =  Cx(bt  -  sin  bt)/b,  since  y  =  0,  when  t  =  0.  To 
evaluate  the  constant  Cj,  substitute  for  dj  and$,  from  (2)  and  the  above,  in  the 
first  of  equations  (1).  We  find  Gx  =  a/b,  and  consequently,  if  a  =  EmjH'2e, 
and  6  =  Helm. 

x  =  a(l  -  cos  bt) ;  y  =  a(bt  -  sin  6£),     ...         (3) 
Let  us  follow  the  motion  of  a  particle  moving  on  the  path  represented  by 
equations  (3).     Of  course  we  can  eliminate  bt  and  get  one  equation  connect- 
ing x  and  yt  but  it  is  better  to  retain  bt  as  the  calculation  is  then  more 
simple.     When 


M  =  0, 

*■. 

2x, 

&r, 

far, 

57T, 

x  =0, 

2a, 

0, 

2a, 

0, 

2a, 

y  =o, 

air, 

2ax, 

3air, 

4o7r, 

5air, 

Hence,  a;  oscillates  to  and  fro  between  0  and  2a ;  y  too  is  periodic,  repeating 

itself  in  the  time  2ir/6  ;  passing  through  a  distance  2o»  from  the 

origin  every  period.     In  other  words,  the  path  of  an  electron 

moving  under  the  above  conditions  is  that  of  a  cycloid  traced 

by  the  rim  of  a  wheel  of  radius  a  rolling  upon  a  plate  Oyt 

Fig.  153. 

(9)  Two  vessels,  capacity  vx  and  v2,  are  filled  with  the  same 
gas  but  at  different  pressures  px  and  p2  respectively.     Assume 
that  the  vessels  are  connected  by  a  capillary  tube  and  that  the 
quantity  of  gas  which  flows  from  one  vessel  to  the  other  is  pro- 
portional to  the  difference  in  the  squares  of  the  pressures  in       w*2a 
the  two  vessels,  and  to  the  time.     What  are  the  pressures,  xx      Fig.  153. 
and  a?2,  in  each  vessel  at  the  end  of  t  seconds  ?    (Lorentz.)     The  quantity  of 
gas,  dQ,  whioh  flows  through  the  capillary  during  the  infinitely  small  in- 
terval of  time  dt  is  by  hypothesis 

dQ  =  a(x^  -  xfldt (4) 

where  a  is  a  constant.  Let  b  denote  the  quantity  of  gas  in  unit  volume,  bv 
will  therefore  denote  the  amount  of  gas  which  occupies  v  volumes  at  atmos- 
pheric pressure.  If  the  pressure  changes  by  an  amount  dx,  the  quantity  of 
gas,  dQ,  changes  an  amount  bv2dx,  hence, 

dQ  m  bv^dt ;  dQ  =  -  to^M.        ...        (5) 

The  difference  in  sign  shows  that  the  gas  which  leaves  one  vessel  enters  the 
other.  The  temperature  is  of  course  supposed  to  remain  constant.  From 
(4)  and  (5), 

dx,  a  ,    '  ■      dx9         a  .    „ 


444  HIGHER  MATHEMATICS.  §  141. 

But  the  total  mass  of  gas  remains  constantly  equal  to  ac,  say 

.-.  V&  +  v^2  =  c ;  .-.  c  =  p1vl  +  #2v2,  ■  •  •  (7) 
by  Boyle's  law.  Multiply  the  first  of  equations  (6)  with  x^oxv^  and  the 
second  by  x^^ ;  subtract  the  latter  from  the  former ;  divide  by  x£ ;  sub- 
stitute x  =  x1(x2  and 

dx  ac.  „      -i 

remains.    Solve  this  equation  in  the  usual  way,  and  we  get 

2       g  a;  -  1        6   +  °  '  °  '  *     8  (^  -  a^)  (^  +  p2)      bvxv2 
From  this  equation  and  the  first  of  equations  (7),  it  is  possible  to  caloulate 
xx  and  x2  at  any  time  t. 

(10)  If  two  adjacent  circuits  have  currents  Gx  and  0&  then,  according  to 
the  theory  of  electromagnetic  induction, 

^  +  L*w  +  B*G*  =  *• ;  M~d  +  L^  +  B&  -  *" 

where  i21}  E2,  denote  the  resistances  of  the  two  circuits,  Lx,  L2,  the  co- 
efficients of  self-induotion,  Ex,  E2,  the  electromotive  forces  of  the  respective 
circuits  and  M  the  coefficient  of  mutual  induction.  All  the  coefficients  are 
supposed  constant. 

First,  solve  these  equations  on  the  assumption  that  EX  =  E2=Q.  Assume 
that  Gx=aemt ;  and  G2=bemt,  satisfy  the  given  equations.  Differentiate  each 
of  these  variables  with  respect  to  t,  and  substitute  in  the  original  equation 

aMm  +  b(L2m  +  B2)  =  0 ;  bMm  +  a(Lxm  +  Bx)  =  0. 
Multiply  these  equations  so  that 

{LXL2  +  M2)m2  +  {LXB2  +  B^m  +  BXB2  =  0. 
For  physical  reasons,  the  induction  LXL2  must  always  be  greater  than  M. 
The  roots  of  this  quadratic  must,  therefore,  be  negative  and  real  (page  354), 
and 

Ci  =  aie  ~  ■¥«  or,  atf  -  "»*  ;  C2  =«  bxe  ~  *#,  or,  V  ~  m*. 
Hence,  from  the  preceding  equation, 

a^Mrn^  +  bxL2mx  +  bxB2  =  0 ;  or  Oj/ftj  =  -  (L2mx  +  R2)lMm\  '■> 
similarly,  a^j^  =  -  Mm2l(Lxm2  +  Bx).      Combining  the  particular  solutions 
for  Gx  and  02,  we  get  the  required  solutions'. 

Ox  =  Oje  ~  mi*  +  a#  -  ™# ;  G2  =  bxe  -  m\*  +  b^e  ~  ™*. 
Second,  if  Ex  and  E2  have  some  constant  value, 
Gx  =  E1/B1  +  axe  -  ***  +  a#  -  ***  ;  G2  =  E2\B2  +  bxe  ~  •"*  +  b#  ~  m*, 
are  the  required  solutions. 


§  141.  Simultaneous  Equations  with  Yariable  Coefficients. 

The  general  type  of  simultaneous  equations  of  the  first  order, 


is 


Pxdx  +  Qxdy  +  BYdz  =  0 ;      \ 

P2dx  +  Q^dy  +  B2dz  =  0, . . .  J  *        '         '         ( *' 


§  1.41.     HOW  TO  SOLVE  DIFFEKENTIAL  EQUATIONS.        445 

where  the  coefficients  are  functions  of  x,  y,  z.  These  equations 
can  often  be  expressed  in  the  form l 

dx      dy  _  dz 

T--Q-"R> (J) 

which  is  to  be  looked  upon  as  a  typical  set  of  simultaneous  equa- 
tions of  the  first  order.  If  one  pair  of  these  equations  involves 
only  two  differentials,  the  equations  can  be  solved  in  the  usual 
way,  and  the  result  used  to  deduce  values  for  the  other  variables, 
as  in  the  second  of  the  subjoined  examples. 

When  the  members  of  a  set  of  equations  are  symmetrical,  the 
solution  can  often  be  simplified  by  taking  advantage  of  a  well- 
known  theorem2  in  algebra — ratio.  According  to  this, 
dx  ,  dy  _  dz  Idx  +  mdy  +  ndz  _  Vdx  +  m'dy  +  n'dz 
P  ==  Q  ~  B  "  IP  +  mQ  +  nB  "  I'P  +  m'Q  +  n'B  =  ■♦•>>8) 
where  I,  m,  n,  V ,  m',  n',. .  .  m&y  be  constants  or-  functions  of 
x,  y,  z.  Since  I,  m,  n,...  are  arbitrary,  it  is  possible  to  choose 
I,  m,  n, . . .  so  that 

IP  +  mQ  +  nB  =  0  ;  I'P  +  m'Q  +  n'B  =  0 ; . . .  (4) 

Idx  +  mdy  +  ndz  =  0,  etc.      .         .  (5) 

The  same  relations  between  x,  y,  z,  that  satisfy  (5),  satisfy  (2) ; 

and  if  (4)  be  an  exact  differential  equation,  equal  to  say  du,  direct 

integration  gives  the  integral  of  the  given  system,  viz., 

u  =  G1   .       \         .        .        .       (6) 
where  C^ denotes  the  constant  of  integration. 
In  the  same  way,  if 

Vdx  +  m'dy  +  n'dz  =  0, 
is  an  exact  differential  equation,  equal  to  say  dv,  then,  since  dv  is 
also  equal  to  zero, 

v-C* (7) 

is  a  second  solution.    These  two  solutions  must  be  independent. 

Examples.— (1)  B^  way  of  illustration  let  us  solve  the  equations 
dx  dy  dz 

y  -  z  ~  z  -  x  ~  x-  y 


1  The  proof  will  come  later,  page  584. 

2  The  proof  is  interesting.    Let  dx'P  =  dy/Q  =  dz/E  =  k,  say  ;  then,  dx  =  Pk 
dy  =  Qk;  dz  =  Bk;  or,  Idx  =  IPk  ;  mdy  =  mQk ;  ndz  =  nBk.     Add  these 
suits,  Idx  +  mdy  +  ndz  =  k(lP  +  mQ  +  nB) 


re- 


Idx  +  mdy  +  ndz  _h_dx_Ay__dz 
IP  +  mQ  +  nB   ~     ~  ~P~  ~  Q  ~  B' 


446  HIGHER  MATHEMATICS.  §  141. 

Here  P  =  y  -  z;  Q  =  z-x',  R=x-y.     Since,  as  in  (4), 

y-z+z-x+x-y=0\  l  =  m  =  n  =  l; 
and  as  in  (6), 

x(y  -  z)  +  y(z  -  x)  +  z{x  -  y)  =  0 ;  I'  =  x ;  m'  =  y;  n'  =  z. 
For  the  first  combination,  therefore 

dx  +  dy  +  dz  =  0 ;  or,  x  +  y  +  z  =  Cx ;      .        .        .      (8) 
and  for  the  second  combination 

xdx  +  ydy  +  zdz  =  0 ;  .\  x2  +  y2  +  z2  =  C2  .  .  .  (9) 
The  last  of  equations  (8)  and  (9)  define  x  and  y  as  functions  of  z,  and  also 
contain  two  arbitrary  constants,  the  conditions  necessary  and  sufficient  in 
order  that  these  equations  may  be  a  complete  solution  of  the  given  set  of 
equations.     Equations  (8)  and  (9)  represent  a  family  of  circles. 

(2)  Solve  dx/y  =  dyjx  =  dzjz.  The  relation  between  dx  and  dy  contains 
k  and  y  only,  the  integral,  y2  -  x2  =  Cv  follows  at  once.  Use  this  result  to 
eliminate  x  from  the  relation  between  dy  and  dz.     The  resultf  is,  p.  349, 

dzjz  =  dylJ{y2  -  Cx) ;  or,  y  +  J(y2  -  Gx)  =  G2z. 
These  two  equations,  involving  two  constants  of  integration,  constitute  a 
complete  solution. 

(3)  Solve  dx/ (mz  -  ny)  =  dy(nx  -  le)  =  dz/(ly  -  mx).  HereP=  mz  -  ny\ 
Q  =  nx  -  lz;  B  =  ly  -  mx.  I,  m,  n  and  x,  y,  z  form  a  set  of  multipliers 
satisfying  the  above  condition.     Hence,  each  of  the  given  equal  fractions  is 

equal  to 

Idx  +  mdy  +  ndz 


and  to 


Accordingly, 


l(mz  -  ny)  +  m(nx  -  lz)  +  n(ly  -  mx) ' 

xdx  +  ydy  +  zdz 
x(mz  -  ny)  +  y(nx  -  lz)  +  z(ly  -  mx)' 


Idx  +  mdy  +  ndz  =  0  ;  xdx  +  ydy  +  zdz  =  0. 
The  integrals  of  these  equations  are 

u  =  Ix  +  my  +  nz  =  Cx;  v  =  x2  +  y2  +  z2  =  C2, 
which  constitute  a  complete  solution. 

dx  _  dy  _  dz      t    _  xdx  +  ydy  +  zdz 

(4)  Solve  x*  _  yi  _  #  ~2xy-2x~z'  •"•  ~  <r(z2  -  y2  -  z2)  +  2xy2  +  2xzv 
dz      2xdx  +  2ydy  +  2zdz 

•'•  j  -  — 32  +  y2  +  ;g2 — ;  •••  log(*2  +  V2  +  *2)  =  log  *  +  log  c2; 

consequently,  x2  +  y2  +  z2  =  C^z  is  the  second  solution  required. 

It  is  thus  evident  that  equation  (5)  must  be  integrable  before 
the  given  set  of  simultaneous  equations  can  be  solved.  The 
criterion  of  integrability,  or  the  test  of  the  exactness  of  an 
equation  containing  three  or  more  variables  is  easily  deduced. 
For  instance,  let 

Pdx  +  Qdy  +  Bdz  =  0 ;  .-.  du  =  ^-dx  +  ^-dy  +  ^~dz.    (10) 

The  second  of  equations  (10)  is  obviously  exact,  and  equivalent  to 
the  first  of  equations  (10),  since  both  are  derived  from  u  =  Cv 
Hence  certain  conditions  must  hold  in  order   that   the   first  of 


oP         7)u        ~dQ 

ftP 

W  = 

_()W 

T?U    \ 

m(*Si' 

*V 

oQ         7)u        DB 

ou 

~bu 

^Tz  =B^  +  ^;- 

=  iV 

■9t*> 

oB         ou        oP 

(oB 

~bP\ 

T?U 

ou 

^x  ~  FDz  +  P-bz  »  ' 

'•  ^te  " 

'   oz)  = 

-pTz- 

B^\ 

§  141.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        44? 

equations  (10)  may  be  reduced  to  the  exact  form  of  the  second. 
As  indicated  above,  there  must  exist  a  function  of  x,  y,  and  zt 
say,  fx,  such  that 

7)u  liu  Tiu 

^=fxP;^=lxQ;o7  =  ^B-  '         '       (11) 

Let  the  student  now  differentiate  each  of  equations  (11),  first  with 
respect  to  y  and  z  ;  second  with  respect  to  z  and  x ;  and  third  with 
respect  to  x  and  y,  the  result  will  be 

OX 

Multiply  the  three  equations  on  the  right,  in  turn,  by  B,  P,  and 
Q  respectively,  and  add  the  results  together.  The  result  gives  us 
the  relation  which  must  hold  between  the  coefficients  P,  Q,  and  B 
of  the  first  of  equations  (10)  in  order  that  it  may  have  an  integral 
of  the  form  u  =  C.     We  must  have,  in  fact, 

<8-fW(S-£W|-S)-<>- 

If  equation  (10)  be  not  exact  it  can  be  made  exact  by  means  of  an 
integrating  factor. 

Examples. — (1)  Given  (y  +  z)dx  +  (z  +  x)dy  +  (x  +  y)dz  =  0,  show  that 
the  condition  of  integrability  is  satisfied.  To  integrate  an  exact  equation  of 
this  kind,  first  suppose  that  z  is  temporarily  constant,  and  integrate.  Thus, 
we  get 

(y  +z)dx  +  (z  +  x)dy  =  0  ;  {y  +  z)  (z  +  x)  =  C  .  .  (13) 
The  integration  constant  obviously  includes  the  variable  z\  let  C  =f(z). 
To  determine  the  form  of  this  function,  differentiate  (y  +  z)  (z  +  x)  =  f(z) 
with  respect  to  x,  y,  and  z,  and  compare  the  result  with  the  given  equation. 
We  get 

{y  +  z)dx  +  (e  +  x)dy  +  (x  +  y)dz  +  2zdz  =  ~^dz ; 

.-.  2zdz  -  df(z)  =  0 ;  or,  f(z)  =  z*  +  G2; 

•*•  iV  +  z)  («  +  a)  =  z2  +  Co ;  or,  xy  +  yz  +  zx  =  GM 

is  the  required  solution.    The  same  result  could  have  been  obtained  more 

quickly,  in  this  particular  case,  by  expanding  the  given  equation  and  so 

getting 

(xdy  +  ydx)  +  (ydz  +  zdy)  +  (zdx  +  xdz)  =  0 ;  .-.  xy  +  yz  +  zx  m  C2. 

(2)  Integrate  yzdx  +  xzdy  +  xydz  =  0.     Divide  by  xyz,  and 

dx      dy      dz 

—  =  -7  =  —  =  0 ;  .*.  log  x  +  log  y  +  log  z  =  log  C ;- .-.  xyz  =  C. 

n        y        • 


448  HIGHER  MATHEMATICS.  §  142. 

(3)  Integrate  xydx  -  zxdx  -  yHz  —  0.  Ansr.  x\y  -  log  a  =  C.  Hint. 
Divide  by  1/xy2  and  the  equation  becomes  exact. 

(4)  If  (ydx  +  xdy)  (a  -  z)  +  xydz  =  0,  show  that  xy  =  G(z  -  a).  Hint. 
Proceed  as  in  Ex.  (1),  making  z  =  constant,  and  afterwards  showing  that 
vy  =  f{z),  and  then  that  f(z)  =  0(z  -  a). 


§  142.    Partial  Differential  Equations. 

Equations  obtained  by  the  differentiation  of  functions  of  three 
or  more  variables  are  of  two  kinds  : 

1.  Those  in  which  there  is  only  one  independent  variable) 
such  as 

Pdx  +  Qdy  +  Edz  =  Sdt, 
which   involves   four  variables— three  dependent   and   one  inde- 
pendent.   These  are  called  total  differential  equations. 

2.  Those  in  which  there  is  only  one  dependent  and  two  or 
more  independent  variables,  such  as, 

tJ)z       ^z          ~bz        - 

PS  +  %  +  BTt  "  °' 

where  z  is  the  dependent  variable,  x,  y,  t  the  independent  variables. 
These  equations  are  classed  under  the  name  partial  differential 
equations.  The  former  class  of  equations  are  rare,  the  latter  very 
common.  Physically,  the  differential  equation  represents  the  rela- 
tion between  the  dependent  and  the  independent  variables  when 
an  infinitely  small  change  is  made  in  each  of  the  independent 
variables.1 

In  the  study  of  ordinary  differential  equations,  we  have  always 
assumed  that  the  given  equation  has  been  obtained  by  the  elimina- 
tion of  constants  from  the  original  equation.  In  solving,  we  have 
sought  to  find  this  primitive  equation.  Partial  differential  equa- 
tions, however,  may  be  obtained  by  the  elimination  of  arbitrary 


1  The  reader  will,  perhaps,  have  noticed  that  the  term  "  independent  variable  "  is 
an  equivocal  phrase.  (1)  If  u  =J\z),  u  is  a  quantity  whose  magnitude  changes  when 
the  value  of  z  changes.  The  two  magnitudes  u  and  z  are  mutually  dependent.  For 
convenience,  we  fix  our  attention  on  the  effect  which  a  variation  in  the  value  of  z  has 
upon  the  magnitude  of  u.  If  need  be  we  can  reverse  this  and  write  z  =f(u),  so  that 
u  now  becomes  the  "  independent  variable".  (2)  If  v  =f(x,  y),  x  and  y  are  "  inde- 
pendent variables "  in  that  x  and  y  are  mutually  independent  of  each  other.  Any 
variation  in  the  magnitude  of  the  one  has  no  effect  on  the  magnitude  of  the  other,  x 
and  y  are  also  "  independent  variables  "  with  respect  to  v  in  the  same  nse  that  *  has 
just  been  supposed  the  ' '  independent  variable  "  with  respect  to  u. 


^  143.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        449 

functions  of  the  variables  as  well  as  of  constants.     For  example,  if 

it  =  f{ax  +  by), 
be  an  arbitral y  funct'on  of  x  and  y,  we  get,  as  in  Euler's  theorem 
page  75, 

£  =  Saf(ax>  i  by*) ';  g  =  8ft/(«*»  +  by*) ;  ,.  *g  -  ag  =  0, 
where  the  arbitrary  function  has  disappeared. 
Examples.— (1)  li  u  =  flat  +  a),  show  that 

dt  ~  adx    dt2  ~  a  dx* 
Here  dufdt  =  af'(at  +  x) ;  'dufdx  =  /'(a*  +  x),  etc.     Establish  the  result  by 
giving  flat  +  x)  some  specific  form,  say,  flat  +  x)  =  at  +  x ;  and  sin  (at  +  x) 

(2)  Eliminate  the  arbitrary  function  from  the  thermodynamic  equation 

OP       rOp  dp 

(3)  Remembering  that  the  object  of  solving  any  given  differential  is  to 
find  the  primitive  from  which  the  differential  equation  has  been  derived  by 
the  elimination  of  constants  or  arbitrary  functions.  -  Show  that  z=f1(x)  +f2{y) 
is  a  solution  of  'dPzfdx'dy  —  0.     Hint.  Eliminate  the  arbitrary  function. 

(4)  Show  that  z  =  fx(x  +  at)  +  f2(x  -  at)  is  a  solution  of  d2zldt2=a,2d2z/dx\ 

An  arbitrary  function  of  the  variables  must  now  be  added  to 
the  integral  of  a  partial  differential  equation  instead  of  the  constant 
hitherto,  employed  for  ordinary  differential  equations.  If  the 
number  of  arbitrary  constants  to  be  eliminated  is  equal  to  the 
number  of  independent  variables,  the  resulting  differential  equa- 
tion is  of  the  first  order.  The  higher  orders  occur  when  the 
number  of  constants  to  be  eliminated,  exceeds  that  of  the  inde- 
pendent variables. 

.  If  u  a*  f(x,  y),  there  will  be  two  differential  coefficients  of  the 
first  order  ;  three  of  the  second ;  . . .  Thus, 

"du   ~du  §  ~d2u   ~d2u     ~b2u 
~dx'  ~dy  *  ^x2'  7)y2'  7)x7iy ' 

§  133.    What  is  the  Solution  of  a  Partial  Differential 
Equation  ? 

Ordinary  differential  equations  have  two  classes  of  solutions 
— the  complete  integral  and  the  singular  solution.  Particular 
solutions   are   only   varieties  of    the   complete    integral.      Three_ 

FF 


450  HIGHER  MATHEMATICS.  §  143. 

classes  of  solutions  can  be  obtained  from  some  partial  differential 
equations,  still  regarding  the  particular  solution  as  a  special  case 
of  the  complete  integral.  These  are  indicated  in  the  following 
example. 

The  equation  of  a  sphere  in  three  dimensions  is, 

X2  +  y2  +  Z2  =  r2j  ,         #         #         (1) 

where  the  centre  of  the  sphere  coincides  with  the  origin  of  the 
coordinate  planes  and  r  denotes  the  radius  of  the  sphere.  If  the 
centre  of  the  sphere  lies  somewhere  on  the  cc^-plane  at  a  point 
(a,  b),  the  above  equation  becomes 

(x  -  af  +  (y  -  bf  +  z2  =  r2.  .  .  (2) 
When  a  and  b  are  arbitrary  constants,  each  or  both  of  which  may 
have  any  assigned  magnitude,  equation  (2)  may  represent  two 
infinite  systems  of  spheres  of  radius  r.  The  centre  of  any  mem- 
ber of  either  of  these  two  infinite  systems — called  a  double  infinite 
system — must  lie  somewhere  on  the  rc?/-plane. 
Differentiate  (2)  with  respect  to  x  and  y. 

x " a  +  z^c =  0;  y  ~ b  +  %  =0-       •     (3) 

Substitute  for  x  -  a  and  y  -  b  in  (2).     We  obtain 

•1©' *©"♦'}-"•  •  •  « 

Equation  (2),  therefore,  is  the  complete  integral  of  (4).      By 

assigning  any  particular  numerical  value  to  a  or  6,  a  particular 

solution  of  (4)  will  be  obtained,  such  is 

(x  -  l)2  +  (y  -  79)2  +  z>  =  r2.        .        .        (5) 

If  (2)  be  differentiated  with  respect  to  a  and  b, 

<>  7) 

^{(x  -  af  +  (y  +  bf  +  z*  =  r2} ;  ^{{x  -  af  +  (y  -  bf  +  s2  =  r2}, 

or,  x  -  a  =  0,  and  y  -  b  =  0. 

Eliminate  a  and  b  from  (2), 

z  =  ±  r (6) 

This  result  satisfies  equation  (4),  but,  unlike  the  particular  solution, 
is  not  included  in  the  complete  integral  (2).  Such  a  solution  of  the 
differential  equation  is  said  to  be  a  singular  solution. 

Geometrically,  the  singular  solution  represents  two  plane  sur- 
faces touched  by  all  the  spheres  represented  by  equation  (2).  The 
singular  solution  is  thus  the  envelope  of  all  the  spheres  represented 


§  143.     HOW  TO  SOLVE  D1FFEKENTIAL  EQUATIONS.         451 

by  the  complete  integral.  If  AB  (Fig.  97)  represents  a  cross  sec- 
tion of  the  #2/-plane  containing  spheres  of  radius  r,  CD  and  EF 
are  cross  sections  of  the  plane  surfaces  represented  by  the  singular 
solution. 

If  the  one  constant  is  some  function  of  the  other,  say, 

a  =  b, 
(2)  may  be  written 

(x  -  af  +  (y  +  of  +  z2  =  r2.         .        .        (7) 
Differentiate  with  respect  to  a.     We  find 

a  =  i(x  +  y). 
Eliminate  a  from  (7).     The  resulting  equation 
x2  +  y2  +  2z2  -  2xy  =  2ra, 
is  called  a  general  integral  of  the  equation. 

Geometrically,  the  general  integral  is  the  equation  to  the 
tubular  envelope  of  a  family  of  spheres  of  radius  r  and  whose 
centres  are  along  the  line  x  =  y.  This  line  corresponds  with  the 
axis  of  the  tube  envelope.  The  general  integral  satisfies  (4)  and 
is^also  contained  in  the  complete  integral. 

Instead  of  taking  a  =  b  as  the  particular  form  of  the  function 
connecting  a  and  b,  we  could  have  taken  any  other  relation,  say 
a  =  ^b.  The  envelope  of  the  general  integral  would  then  be  like 
a  tube  surrounding  all  the  spheres  of  radius  r  whose  centres  were 
along  the  line  x  =  \y.  Had  we  put  a2  -  b2  =  1,  the  envelope  would 
have  been  a  tube  whose  axis  was  an  hyperbola  x2  —  y2  =  1. 

A  particular  solution  is  one  particular  surface  selected  from  the 
double  infinite  series  represented  by  the. complete  solution.  A  general 
integral  is  the  envelope  of  one  particular  family  of  surfaces  selected 
from  those  comprised  in  the  complets  integral.  A  singular  solution 
s  an  envelope  of  every  surface  included  in  the  complete  integral.1 

Theoretically  an  equation  is  not  supposed  to  be  solved  com- 
pletely until  the  complete  integral,  the  general  integral  and  the 
singular  solution  have  been  indicated.  In  the  ideal  case,  the 
complete  integral  is  first  determined  ;  the  singular  solution  ob- 
tained by  the  ehmination  of  arbitrary  constants  as  indicated  above  ; 
the  general  integral  then  determined  by  eliminating  a  and  f(a). 

Practically,  the  complete  integral  is  not  always  the  direct  ob- 

1 G.  B.  Airy's  little  book,  An  Elementary  Treatise  on  Partial  Differential 
Equations,  London,  1873,  will  repay  careful  study  in  connection  with  the  geometrical 
interpretation  of  the  solutions  of  partial  differential  equations. 


452  HIGHER  MATHEMATICS.  §  144. 

ject  of  attack.  It  is  usually  sufficient  to  deduce  a  number  of 
particular  solutions  to  satisfy  the  conditions  of  the  problem  and 
afterwards  to  so  combine  these  solutions  that  the  result  will  not 
only  satisfy  the  given  conditions  but  also  the  differential  equation. 
Of  course,  the  complete  integral  of  a  differential  equation 
applies  to  any  physical  process  represented  by  the  differential 
equation.  This  solution,  however,  may  be  so  general  as  to  be  of 
little  practical  use.  To  represent  any  particular  process,  certain 
limitations  called  limiting  conditions  have  to  be  introduced. 
These  exclude  certain  forms  of  the  general  solution  as  impossible.1 
We  met  this  idea  in  connection  with  the  solution  of  algebraic 
equations,  page  363. 

§  1M.   The  Linear  Partial  Equation  of  the  First  Order. 

Let  u  =  Gv  .        .         .         .        (1) 

be  a  solution  of  the  linear  partial  equation  of  the  first  order  and 
degree,  namely  of 

->bz      ~bz       _ 

PU+%-^  •  •  •  (2) 
where  P,  Q,  and  B  are  functions  of  x,  y,  and  z ;  and  G1  is  a  con- 
stant. Now  differentiate  (1)  with  respect  to  x,  and  y  respectively, 
as  on  page  44,  or  Ex.  (5),  page  74. 

bu      bu   bz  _  n  e   bu      bu   bz 

~dx       bz  '  bx      ,      '    by       ~dz'  by  ~     '  '  ^  ' 

Now  solve  the  one  equation  for  bzfbx,  and  the  other  for  bz/by,  and 
substitute  the  results  in  (2).     We  thus  obtain 
-fiu       r$u       ^bu 

p^+%  +  B^  =  °-  •  ■  •    •  ■■■(*) 

Again,  let  (1)  be  an  integral  of  the  equation 

bx  _  "by  _  bz 

p~~~Q"~ir  •      *      '      (5) 

The  total  differential  of  u  with  respect  to  x,  y,  and  z,  is 

bu  _        bu  _        bu  , 

^ax+Tyay+Tzaz~o-        .      .      (6) 

and  since,  by  equations,  dx  =  kP ;  dy  =  kQ;  dz  =  kB,  page  445 

(footnote),  we  have 

bu.^        bu~        bu^       n  '    % 

TxP  +  TyQ  +  TzR-(i'     ...         (7) 


1  For  examples,  see  the  end  of  Chapter  VIII.  ;  also  page  460,  and  elsewhere. 


~du      ~du    ~dz       „.,  ^/~dv  .  Dv    ~bz\_Du 

Dx     Dz  '  Dx 


§  144.     HOW  TO  SOLVE  DIFFEKENTIAL  EQUATIONS.        453 

which  is  identical  with  (4).  This  means  that  every  integral  of  (2) 
satisfies  (5),  and  conversely.  The  general  integral  of  (2)  will 
therefore  be  the  general  integral  of  (5). 

What  has  just  been  proved  in  connection  with  u  =  G1  also 
applies  to  the  integral  v  =  C2  of  (7),  page  445.  If  therefore  we 
can  establish  a  relation  between  u  and  v  such  that 

u  =  f(v) ;  or,  <t>(u,  v)  =  0,  .         .         (8) 

this  arbitrary  function  will  be  a  solution  of  the  given  equation. 
This  is  known  as  Lagrange's  solution  of  the  linear  differential 
equation  j  equations  (5)  are  called  Lagrange's  auxiliary  equa- 
tions. 

We  may  now  show  that  any  equations  of  form  (8)  will  furnish 
a  definite  partial  equation  of  the  linear  form  (2).  Differentiate 
Equations  (8),  say  the  first,  with  respect  to  each  of  the  inde- 
pendent variables  x  and  y.     We  get 

/Dv    Dv    Dz\,Du    Du   Dz  /Dv    Dv    Dz\ 

~J  {  )\Dx    Dz  '  Dx/'  Dy    Dz  '  Dy  ~  J  ^  \Dy    Dz'DyJ' 
By  division  and  rearrangement  of  terms,  f(v)  and  the  terms  con- 
taining the  product  of  Dz/Dx  with  Dz/Dy  disappear,1  and  we  get 
dx  dy  dz 

Du    Dv       Du    Dv        Du    Dv        Du    Dv  ~  Du    Dv        Du    Dv'       ^  ' 
~dy  '  Dz      Dz'  Dy      Dz  '  Dx      Dx'  ~dz      dx '  Dy      Dy'  Dx 
This  equation  has  the  same  form  as  Lagrange's  equation 
dx      dy      dz  Dz  Dz 

-P=-Q  =  B>  and%+  %  =  B> 
hence,  if  u  =  f(v)  is  a  solution  of  (2),  it  is  also  a  solution  of  (5). 

Examples. — (1)  Solve  E.  Clapeyron's  equation  (Journ.  de  VEcole  Roy. 
PolyL,  14,  153,  1834), 

*^  +  %-  ~% <10> 

well  known  in  thermodynamics.  Here,  P  =  p,  Q  =  p,  B  =  -  p/ap,  and  La- 
grange's auxiliary  equations  assume  the  form, 

§£  =  ^1  -JL (n) 

P  p  ap 

From  the  first  pair  of  these  equations  we  get  logp  -  logp  =  log  Cl ;  conse- 
quently pip  =  Gv     From  the  first  and  last  of  equations  (11),  we  have 

1  When  the  reader  has  read  Chapter  XI.  he  will  write  the  denominator  in  the 
form  of  a  "  Jacobian". 


454  HIGHER  MATHEMATICS.  §  145. 

is  the  second  solution  of  (10).     The  complete  solution  is  therefore 

Q^-JLlogp+ff-P 


aP  \p, 

(2)  Solve  y  .  dz/dx  -  x  .  dz/dy  =  0  ;  here,  P  =  y,  Q  =  -  x,  R  =  0, 

dx       dy       dz  ,  _        •       , 

.*.  —  = —  ~rT'>  .'.  dz  =  0,  and  xdx  +  ydy  =  0. 

y       —  x       u 

.-.  x2  +  y2  =  C2 ;  and  «  =  C2  ;  or,  z  =  f(x2  +  y2). 

(3)  Solve  xz .  dzjdx  +  yz .  dz/dy  =  xy.  Here,  P  =  1/y,  Q  =  1/s,  £  =  l/#. 
The  auxiliary  equations  are  therefore  ydx  =  xdy  =  zdz.  From  the  first  two 
terms  we  get  y\x  =  Cx ;  and  from  the  multipliers  I  =  y ;  m  =  x  ;  n  =  -  2z, 
as  on  page  445  (4),  we  get 

Idx  +  mdy  +  ndz  =  0 ;  .*.  ydx  +  xdy  =  2zdz ;  or,  z2  -  xy  =  C2, 
from  (5),  page  445.     This  is  the  second  solution  required.     Hence,  the  com- 
plete solution  is  z2  =  xy  +  f(x/y) ;  or,  <p(z2  -  xy,  x/y)  =  0. 

(4)  Moseley  (Phil.  Mag.  [4],  37,  370,  1869)  represents  the  motion  of  im- 
perfect fluids  under  certain  conditions  by  the  equation 

"dz       'dz 

Vx'  +  Vy-™*  •'•  *  =  emyKx  ~  *)' 

§  155.    Some  Special  Forms. 

For  the  ingenious  general  methods  of  Charpit  and  G.  Monge, 
the  reader  will  have  to  consult  the  special  text-books,  say,  A.  E. 
Forsyth's  A  Treatise  on  Differential  Equations,  London,  1903. 
There  are  some  special  variations  from  the  general  equation  which 
can  be  solved  by  "  short  cuts  ". 

I.  The  variables  do  not  appear  directly.     The  general  form  is 
/~dz    ~dz\ 

;w  V  =        '      '      •      •      L 

The  solution  is 

z  =  ax  +  by  +  G, 

provided  a  and  b  satisfy  the  relation 

f(a,  6)  =  0;  or  b=f(a). 

The  complete  integral  is,  therefore, 

z  =  ax  +  yf{a)  +  G.        .         .         .         (1) 

(*dz\2  fdz\2 
dx)  +  \dy)  =  m2' Sh°W  z==ax  +  bV  +  d  provided 
a2  +  b2  =  m2.  The  solution  is,  therefore,  z  =*  ax  +  y  sj(m2  -  a2)  +  C.  For 
the  general  integral,  put  C  =  f(a).  Eliminate  a  between  the  two  equations, 
z  =  ax  +  sl{m2  -  a2)y  +  f(a)  ;  and  x  -  a{m2  -  a2)~iy  +f'{a)  =  0,  in  the  usual 
way.  The  latter  expression  has  been  obtained  from  the  former  by  differ- 
entiation with  respect  to  a. 


§  145.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        455 

(2)  Solve  zxzy  «=  1.  Ansr.  z  =  ax  +  yja  +  C.  Hint,  ab  =  1,  etc.  Note. — 
We  shall  sometimes  write  for  the  sake  of  brevity,  p  =  'dzl'dx—zx;  q='dzj'dy=zy. 

(3)  Solve  a(zx  +  zv)  =  z.  Sometimes,  as  here,  when  the  variables  do  ap- 
pear in  the  equation,  the  function  of  x,  which  occurs  in  the  equation,  may 
be  associated  with  dz/dx,  or  a  function  of  y  with  dzjdy,  by  a  change  in  the 
variables.  We  may  write  the  given  equation  ap\z-\-aq\z~\.  Put  dz[z  =  dZ; 
(Iy/a  =  dY,  dx\a—dX,  hence,  d)Z '/dY +  dZ [dX=l,  the  form  required.  Ansr. 
Z  =  aX  +  Z(l  -  a)  +  C  ;  where,  Z  =  log  z ;  Y  =  y\a\  X  =  x\a,  etc. 

(4)  Solve  xhi2  +  y2u2  =  z2.  Put  X  -  log  x,  Y  =  log  y,  Z  =  log  z.  Pro- 
ceed as  before.     Ansr.  z  =  CxaysJ(1~a2). 

If  it  is  not  possible  to  remove  the  dependent  variable  z  in  this 
way,  the  equation  will  possibly  belong  to  the  following  class : — 

II,  The  independent  variables  x  and  y  are  absent.  The  general 
form  is, 

/     "dz  tz\ 

TS?W" 

Assume  as  a  trial  solution,  that 


0.  ...        II. 


~bz  _     ~dz 

~dy  ■      ~dx 
Let  ~dzfbx  be  some  function  of  z  obtained  from  II.,  say  ux  =  cj>(z). 
Substitute  these  values  in 

dz  =  zxdx  +  zydy. 

We  thus  get  an  ordinary  differential  equation  which  can  be  readily 

integrated. 

f  dz 
dz  =  <f>(z)dx  +  a${z)dy.     .\  x  +  ay  =    — rr  +  C.  (2) 

Examples.— (1)  Solve  z2z  +  zy2  =  4.     Here  put  dz/dy  =  adz/dx, 

.-.  {a2  +  z)  {dz/dx)2  =  4.     sJa^T~z .  dz/dx  =  2, 

.-.  x  +  ay  +  C=fe{a2  +  z)tdz=l(a?  +  z)™.     Ansr.  2(a2  +  z)3  =  3(x  +  ay  +  C)2. 

4 

(2)  Solve  p(l  +  q2)  =  q(z  -  a).   Ansr.  ^  (z  -  a)  =  (bx  +  y  +  C)2  +  4.  Hint. 

Put  q  =  bp,  etc.     The  integration  and  other  constants  are  collected  under  C. 

(3)  Moseley  (Phil.  Mag.,  [4],  37,  370,  1869)  has  the  equation  of  the  motion 
of  an  imperfect  fluid 

fiz      "dz 

Let  |*  =  a|2;  .-.  f-2  +  ag*  =  W;  ...  f!  =  J~-  ;  ...  %  =  £2"     by  sub- 
dy        dx'        ox        ox  dx      1  +  a         dy      1  +  a'     J 

stitution  in  the  original  equation.     From  (3), 

mz    ,         amz    ,      dz    -*»,„,       ,  m 

d*  =rr^x  +  rr^  ^ ;  T=fT^  +  a^} ;  •'• log* =  m{x  +  a^  +  °- 


456  HIGHER  MATHEMATICS.  §  145. 

If  z  does  not  appear  directly  in  the  equation,  we  may  be  able 
to  refer  the  equation  to  the  next  type. 

III.  z  does  not  appear  directly  in  the  equation,  but  x  and  'bzfbx 
can  be  separated  from  y  and  ~bzfiy.     The  leading  type  is 

#D=4-D-     •  •  m 

Assume  as  a  trial  solution,  that  each  member  is  equal  to  an  arbi- 
trary constant  a,  so  that  zx,  and  zy  can  be  obtained  in  the  form, 

zx  =  f^x,  a);  zy=  f2(y,  a) j  dz  =  zxdx  +  zydy, 
then  assumes  the  form 

z  =  lfx{x,  a)dx  +  Sf2(y,  a)dy  +  G.       .        .         (3) 

Examples. — (1)  Solve  zy  -  zx  +  x  -  y  =  0.  Put  dz/dx  -x  =  'dzj'dy  -y  =  a. 
Write  zx  =  x  +  a;  zy  =  y  +  a; 

.-.  dz  =  (x  +  a)dx  +  {y  +  a)dy,  z  =  %(x  +  a)2  +  %(y  +  a)2  +  C. 

(2)  Assume  with  S.  D.  Poisson  (Ann.  de  Chim.,  23,  337, 1823)  that  the 
quantity  of  heat,  Q,  contained  in  a  mass  of  gas  depends  upon  the  pressure, 
p,  and  its  density  p,  so  that  Q  =f{p,  p).  According  to  the  well-known  gas 
equation,  p  =  Rp(l  +  a9) ;  if  _p  is  constant, 

dp  _        fy  a   dp  _     Rp 

de~~  IT^;  a    '  de  ~  TTVe' 

if  p  is  constant.     Prom  (10)  and  (7),  page  80,  the  specific  heats  at  constant 
pressure,  and  constant  volume  (i.e.,  p  =  constant)  may  be  written 

Cp  ~  \dp)P{ve)p-  ~  {TpJpT+Ve'  and  °-  WJXdejr  WJ.VTVe 

Assuming,  with  Laplace  and  Poisson,  that  y  =  CpjGv  is   constant,  we  get, 
by  division. 

This  differential  equation  comes  under  (3).     Put 

dQ  _  j^.  dQ  _  _  <z 
'dp      yp '  dp  p ' 

.-.dQ  =  a(-.-£  -  -£y,   or,  Q  =  ^  logp  -  a  log  p  -t  G. 

If  tp  is  an  independent  function, 


-(f>-? 


*(Q)» 


where  \p  is  the  inverse  function  of  <p.  If  it  be  assumed  that  the  quantity  of 
heat  contained  in  any  gas  during  any  change  is  constant,  ^(Q)  will  remain 
constant.     Otherwise  expressed, 

py 

—  =  constant ;  or,  pv"Y  =  constant. 

since  volume,  v,  varies  inversely  as  the  density,  p.  This  relation  was  deduced 
another  way  on  page  258. 


§  146.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        457 

(4)  E.  Clapeyron's  equation,  previously  discussed  on  page  453,  may  be 
solved  by  the  method  of  the  separation  of  the  variables. 


dQ       dQ         P        dQ      1  dQ 
pdp  +  pdp-~  pa>  •'•  dp  +  c '  dp  ~ 

c 

ap 

3oyle's  law. 

dQ       c             dQ 
.  .  "^r-  +  —  =  A  :  "ts~  =  Ac. 
dp      ap         'dp 

Hence,  by  integration  and  substitution  in  (3),  we  get 

Q  =  Ap  +  AcP  +  C-  £logp;  or,  Q  =/(^)  -  i^ogp, 

by  collecting  all  the  integration  functions  under  the  symbol /(.  .  .),  and  sub- 
stituting for  c  from  Boyle's  law.  Of  course  f(p/p)  can  only  be  evaluated  when 
the  relation  between  p  and  p  is  known.  C.  Holtzmann  assumed  that  this 
function  could  be  written  =  A  +  BT,  where  A  and  B  were  constants,  T  the 
absolute  temperature. 

IV.  Analogous  to  Clairaut's  equation.     The  general  type  is 
~bz  ~dz  rfbz    ^z\ 

z  =  ^-x  +  yry+f{rx-^)-   •     ■    IY- 

The  complete  integral  is 

z  =  ax  +  by  +  /(a,  b).  .         .         (4) 

Examples. — Solve  the  following  equations : 

(1)  z=zxx  +  zyy  +  zxzy.     Ansr.  z=ax  +  by  +  ab.     Singular  solution  z  =  -xy. 

(2)  z  =  z^c  +  zyy  +  r  ^/(l  +  ex2  +  zy2).  Ansr.  z  =  ax  +  by  +  r  \/l  +  a2+6^ 
Singular  solution,  x*  +  y*  +  z2  =  r2.  The  singular  solution  is,  therefore,  a 
sphere;  r,  of  course,  is  a  constant. 

(3)  z=zxx  +  zyy-nZ/zxzy.  Ansr.  z=ax  +  by-n\/ab.  Singular  solution, 
z  =  (2  -  n)  {xyyiV-  «>. 

There  are  no  general  methods  for  the  solution  of  partial  differ- 
ential equations,  and  it  is  only  possible  to  perform  the  integration 
in  special  cases.  The  greatest  advances  in  this  direction  have  been 
made  with  the  linear  equation.  Linear  equations  are  often  en- 
countered in  physical  mathematics. 

§  146.  The  Linear  Partial  Equation  of  the  Second  Order. 

Suppose  an  elastic  medium  (gas)  to  be  confined  in  a  tube  of  unit 
sectional  area ;  let  E  denote  the  coefficient  of  compressional 
elasticity  of  the  gaseous  medium  ;  and  p  a  force  which  will  produce 
a  compression  du  in  a  layer  of  the  gas  dx  thick,  then,  since 

Stress  =  elasticity  x  strain, 


458  HIGHER  MATHEMATICS.  §  146. 

as  in  Hooke's  well-known  law — ut  tensio  sic  vis — we  get 

Again,  the  layer  dx  will  be  moved  forwards  or  backwards  by  the 
differences  of  pressure  on  the  two  sides  of  this  layer.  Let  this 
difference  be  dp.     Hence,  by  differentiation  of  p  and  du,  we  get 

*P-&&      ....        (2) 

Let  p  denote  the  density  of  the  gas  in  the  layer  dx,  then,  the  mass  m, 

m  =  pdx. 

Now  the  pressure  which  moves  a  body  is  measured,  in  dynamics, 
as  the  product  of  the  mass  into  the  acceleration,  or 

,    ,         d2u        d2u. 
dp/m=w=  p^dx. 

The  equation  of  motion  of  the  lamina  is 

*'•  dt2  ~  P  dx2'  *        #        '        (3) 

This  linear  homogeneous  partial  differential  equation  represents 
the  motion  of  stretched  strings,  the  small  oscillations  of  air  in 
narrow  (organ)  pipes,  and  the  motion  of  waves  on  the  sea  if  the 
water  is  neither  too  deep  nor  too  shallow.  Let  us  now  proceed  to 
the  integration  of  this  equation. 

There  are  many  points  of  analogy  between  the  partial  and  the 
ordinary  linear  differential  equations.  Indeed,  it  may  almost  be 
said  that  every  ordinary  differential  equation  between  two  variables 
is  analogous  to  a  partial  differential  of  the  same  form.  The  solu- 
tion is  in  each  case  similar,  but  there  are  these  differences  : 

First,  the  arbitrary  constant  of  integration  in  the  solution  of 
an  ordinary  differential  equation  is  replaced  by  a  function  of  a 
variable  or  variables. 

Second,  the  exponential   form,   Cemx,  of  the  solution   of   the 

ordinary  linear  differential  equation  assumes  the  form  e     by<f>(y). 


The  expression,  e    sv<p(y),  is  known  as  the   symbolic  form   of  Taylor's 
theorem.      Having  had  considerable  practice  in  the  use  of  the  symbol  of 

<-\  -p. 

operation  D  for  ^-,  we  may  now  use  D'  to  represent  the  operation  ^-.     By 


§  146.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        459 

Taylor's  theorem. 


<p(y  +  mx)  =  <p(y)  +  mx  -^-  +  -^y .  ~^2    + 


d<p(y)  .  mW  d2<p(y) 

jrtu;  — 
where  x  is  regarded  as  constant. 

t       /  d        mW    d* 

.-.  <{>(y  +  mx)  =  ^1  +  mx^y  +  -jj-  ^  + 

The  term  in  brackets  is  clearly  an  exponential  series  (page  285),  equivalent  to 

e     6y,  or,  writing  D  for  «-, 

<p(y  +mx)  =  e™D'<t>(y).  ....      (4) 

Now  convert  equation  (3)  into 

W  "  ^         •         •         •         •       (5) 

by  writing  <z2  =  i^/o.     This  expression  is  sometimes  called  d'Alem- 

bert's  equation.     Instead  of  assuming,  as  a  trial  solution,  that 

y  =  e™*,  as  was  the  case  with  the  ordinary  equation,  suppose  that 

u  =  f(x  +  mt),      .  .  (6) 

is  a  trial  solution.     Differentiate  (6),  with  respect  to  t  and  x>  we 

thus  obtain, 

7)u  ~bu  ~b2u 

Tt  =  mf(x  +  mt);  ^  =  f(x  +  mt);  ^  =  mf\x  +  mt) ; 

-572  =  mT(z  +  mi) ;  ^2  -  /'(*  +  m0- 

Substitute  these  values  in  equation  (5)  equated  to  zero,  and  divide 
out  the  factor  f"(x  +  mt).     The  auxiliary  equation, 

m2  -  a2  =  0  .  .  .  .  (7) 
remains.  If  m  is  a  root  of  this  equation,  f"(x  +  mt)  =  0,  is  a 
part  of  the  complementary  function.  Since  +  a  are  the  roots  of 
(7),  then 

u  =  e-°tD'f1(x)  +  eatD'f2(x).  .        .       (8) 

From  (4)  and  (6),  therefore, 

u  =  fx(x  -  at)  +  f2(x  +  at)  .  .  .  (9) 
Since  +  a  and  -  a  are  the  roots. of  the  auxiliary  equation  (7),  we  can 
write  (5)  in  the  form, 

(D  +  aD')  {D  -  aD')u  =  0.  .         .        .       (10) 

d2z      d*z 
Examples.— (1)  If  ^  -  ^  =  0,  show  e  m  fx{y  +  x)  +  f.2(y  -  x). 

d2z  ?Pz  ^z 

(2)  If  W>  "  *dxdy  +  43^2  =  °'  show  z  =  f^V  +  2x)  +My  +  2x)' 

(3)  If  2  ^  -  3^  -  2^-2  =  0,  show  0  =  M2y  -  x)  +  f2(y  +  2x). 

In  the  absence  of  data   pertaining  to   some  specific  problem, 
we  cannot  say  much  about  the  undetermined  functions  fx(x  +  at) 


460  HIGHER  MATHEMATICS.  §  146. 

and  f2(x  -  at)  of  (9).  Consider  a  vibrating  harp  string,  where  no 
force  is  applied  after  the  string  has  once  been  put  in  motion.     Let 

x  =  I  -  AB  (Fig.  154)  denote  the  length  of 
the  string  under  a  tension  T ;  and  m  the 
mass  of  unit  length  of  the  vibrating 
string.  In  the  equation  of  motion  (5), 
in  order  to  avoid  a  root  sign  later  on, 
FlQ  154#  a2  appears  in  place  of  T/m.     Further,  let 

u  PM  represent  the  displacement  of  any 
part  of  the  string  we  p^ase,  and  let  the  ordinate  of  one  end  of  the 
string  be  zero.  Then,  whatever  value  we  assign  to  the  time  t,  the 
ends  of  the  string  are  fixed  and  have  the  limiting  condition  u  =  0, 
when  x  =  0 ;  and  u  =  0,  when  x  =  L 

•••  /iO0  +  M  ~at)  =  0;  0  +  at)  +  fjl  -  at)  =  0,  (11) 
are  solutions  of  d'Alembert's  equation  (5).  From  the  former,  it 
follows  that 

fl(at)  must  always  be  equal  to  —  /2(  —  at)  .         (12) 

But  at  may  have  any  value  we  please.  In  order  to  fix  our  ideas, 
suppose  that  we  put  I  +  at  for  at  in  the  second  of  equations  (11)  * 
then,  from  (12), 

Mat  +  21) -Mat).  .  .  .  (13) 
The  physical  meaning  of  this  solution  is  that  when/1(.  . .)  is  increased 
or  diminished  by  21,  the  value  of  the  function  remains  unaltered. 
Hence,  when  at  is  increased  by  21,  or,  what  is  the  same  thing,  when 
t  is  increased  by  2lja,  the  corresponding  portions  of  the  string  will 
have  the  same  displacement.  In  other  words,  the  string  performs  at 
least  one  complete  vibration  in  the  time  2l/a.  We  can  show  the  same 
thing  applies  for  4Z,  61.  .  .  .  Hence,  we  conclude  that  d'Alembert's 
equation  represents  a  finite  periodic  motion,  with  a  period  of  oscil- 
lation. 


at 


21;  or,  t-  |;  or,  t  =  2lyJ~-   ■         •       (14) 

Numerical  Example. — The  middle  C  of  a  pianoforte  vibrates  .264  times 
per  second,  that  is,  once  every  -^  second.  If  the  length  of  the  wire  is  2£ 
feet,  and  one  foot  of  the  wire  weighs  0-002  lbs.,  find  the  tension  T  in  lbs.  Now 
mass  equals  the  weight  divided  by  g,  that  is  by  32.     Hence, 

*  -  «Vsl?;  2§i  -  5Va^=  •••  T  = 108  lbs- 

Equation  (5),  or  (9),  represents  a  wave  or  pulse  of  air  passing 
through  a  tube  both  from  and  towards  the  origin.     If  we  consider 


§  146.    HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        4G1 

a  pulse  passing  from  the  origin  only, 

u  =  f{x  +  at) 
is  the  solution  of  the  differential  equation.    By  differentiation  with 
regard  to  x,  and  with  regard  to  t,  we  have  already  shown,  Ex.  (1), 

page  449,  that 

du  du 

s  =  j  fix  +  at),  -gj-  af(x  +  at). 

The  first  of  these  equations  represents  the  rate  of  expansion  or 
contraction ;  the  second,  the  velocity  of  a  particle.  The  velocity  of 
the  wave  is,  by  division, 

dx  ■  IS  m 

which  is  Newton's  formula  for  the  velocity  of  sound  (Newton's 
Principia,  ii.,  Prob.  43-50).  Newton  made  E  represent  the  iso- 
thermal elasticity,  p  ;  Laplace,  the  adiabatic  elasticity  yp  of 
page  114. 

When  two  of  the  roots  in  equation  (7)  are  equal  to,  say,  a. 
We  know,  page  401,  that  the  solution  of 

(D  -  afz  -  0,  is  z  =  eT'^x  +  G2), 
by  analogy,  the  solution  of 

(D  -  aD'yz  -  0,  is  z  -  (T^ixf^y)  +  f2(y)h 

or,  z  =  xfx(y  +  ax)  +  f2(y  +  ax).       .         .         (15) 

Examples.— (1)  Solve :  (D3  -  D'D'  -  DD'*  +  D'3)z  =  0. 

Ansr.  z  =  xfx(y  -  x)  +  /3(y  -  a)  +  f3(y  +  x). 

'cpz       n    &Z  &Z 

(2)  ^2  +  2?tfdy  +  dy*  =  0-   Ansr-  *  =  */i(y  +  x)  +Mv  +  x)' 

If  the  equation  *be  non-homogeneous,  say, 
l2z  Wz  ~b2z  ^z  'dz 

Ao^  +  ^igjty  +  A*tyi  +  ^  +  ^  +  V  =  0,      (16) 

and  it  can  be  separated  into  factors,  the  integral  is  the  sum  of  the 
integrals  corresponding  to  each  symbolic  factor,  so  that  each  factor 
of  the  form  D  -  mD',  appears  in  the  solution  as  a  function  of 
y  +  mx,  and  every  factor  of  the  form  D  -  mD'  -  a,  appears  in 
the  solution  in  the  form  z  =  eaxf(y  +  mx). 

Examples.-(1)  Solve  ^-^+^-  +  ^-=0. 
Factors,  (D  +  D')  (D  -  D'  +  l)z  =  0.     Ansr.  z  =  fx(y  -  x)  +  e  -  %(y  +  x). 
(2)  Solve  — _^+^-^-=o. 

Factors,  (D  +  1)  (D  -  D>  =  0.     Ansr.  z  =  e-*fx{y)  +  f2(x  +  y). 
It  is,  however,  not  often  possible  to  represent  the  solutions  of 


462  HIGHER  MATHEMATICS.  §  146. 

these  equations  in  this  manner,  and  in  that  case  it  is  customary  to 
take  the  trial  solution, 

z  =  e«*+0y.  .  .  .  .  (17) 
Of  course,  if  a  is  a  function  of  j3  we  can  substitute  a  =  f(fi)  and  so 
get  rid  of  /?.     Now  differentiate  (17)  so  as  to  get 

~bz  ~dz  ~d2Z  02Z  1)2Z 

5S  -  az;  Ty  =  ^'  5^  "  a^;  W  ~  az'  5p  "  ?"• 

Substitute  these  results  in  (16).      We  thus   obtain  the  auxiliary 

equation 

(V2  +  Ai*P  +  A2@2  +  A<sa  +  Afi  +  Ab)z  =  0-    •       (18) 
This  may  be  looked  upon  as  a  bracketed  quadratic  in  a  and  (3. 

Given  any  value  of  (3,  we  can  find  the  corresponding  value  of  a ;  or 
the  value  of  /3  from  any  assigned  value  of  a.  There  is  thus  an 
infinite  number  of  particular  solutions.  Hence  these  important 
rules  : 

I.  If  uv  u2,  u3, . . . ,  are  particular  solutions  of  any  partial  dif- 
ferential equation,  each  solution  can  be  multiplied  by  an  arbitrary 
constant  and  each  of  the  resulting  products  is  also  a  solution  of  the 
equation. 

II.  The  sum  or  difference  of  any  number  of  particular  solutions 
is  a  solution  of  the  given  equation. 

It  is  usually  not  very  difficult  to  find  particular  solutions,  even 
when  the  general  solution  cannot  be  obtained.  The  chief  difficulty 
lies  in  the  combining  of  the  particular  solutions  in  such  a  way, 
that  the  conditions  of  the  problem  under  investigation  are  satisfied. 
In  order  to  fix  these  ideas  let  us  study  a  couple  of  examples  which 
will  prepare  the  way  for  the  next  chapter. 

Examples.— (1)  Solve  (D2-D')z=0.  Here  a2-£=0;  .-.  0  =  a2.  Hence 
(17)  becomes  z  =  Ce<*z  +  a2y.  Put  a  =  £,  a  =  1,  a  =  2, . .  .  and  we  get  the  par- 
ticular solutions  eW*  +  V\  ex  +  y,  e2x  +  4^  .  . . 

.-.  z  =  Cj«Kto  +  y">  +  a#x  +  y  +  ojp*  +  *y  +  . . . 

Now  the  difference  between  any  two  terms  of  the  form  e"*  +  M,  is  included  in 
the  above  solution,  it  follows,  therefore,  that  the  first  differential  coefficient 
of  e"*  +  Py,  is  also  an  integral,  and,  in  the  same  way,  the  second,  third  and 
higher  derivatives  must  be  integrals.     Since, 
DeaX  +  a*y  =  (x  +  2ay)e(lX  +  a*y  ;  DH0*  +  a*y  =  {{x  +  2ay)  +  2y}eaX  +  a2^  j 
Dseax  +  a*,  =  ^x  +  2ay)  +  §y(x  +  2ay))eaX  +  a2y  ;  etc., 
we  have  the  following  solution  : — 

*  =  d(a;  +  2ay)eax  +  ^  +  C2{{x  +  2ay)  +  2y]e*x  +  °*y  +  . . . 
If  a  =  0,  we  get  the  special  case, 

z  =  Cxx  +  C2(x2  +  2y)  +  C3(x3  +  6xy)  +  ... 


§  147.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        463 

(2)  Solve  |^  -  ||3  -  3^  +  3||  =  0.     Put  *  =  OS*  +  &  ;  and  we  get 
(a  -  j8)  (a  +  0  -  8)  =  0.     .'.  0  =  a,  and  0  =  3  -  a. 
.-.  •  =  C^*  +  ^>  +  &G4&  ~  *>  =/i(2/  +  a;)  +  e^f2(x  -  y). 

The  processes  for  finding  the  particular  integrals  are  analogous 
to  those  employed  for  the  particular  integrals  of  ordinary  differ- 
ential equation,  I  shall  not  go  further  into  the  matter  just  now,  but 
will  return  to  the  subject  in  the  next  chapter.  Partial  differential 
equations  of  a  higher  order  than  the  second  sometimes  occur  in 
investigations  upon  the  action  of  magnetism  on  polarized  light ; 
vibrations  of  thick  plates,  or  curved  bars  ;  the  motion  of  a  cylinder 
in  a  fluid  ;   the  damping  of  air  waves  by  viscosity,  etc. 


§  147.  The  Approximate  Integration  of  Differential  Equations. 

There  are  two  interesting  and  useful  methods  for  obtaining  the 
approximate  solution  of  differential  equations  : 

I:  Integration  in  series.  When  a  function  can  be  developed  in 
a  series  of  converging  terms,  arranged  in  powers  of  the  independent 
variable,  an  approximate  value  for  the  dependent  variable  can  easily 
be  obtained.  The  degree  of  approximation  attained  obviously 
depends  on  the  number  of  terms  of  the  series  included  in  the 
calculation.  The  older  mathematicians  considered  this  an  under- 
hand way  of  getting  at  the  solution,  but,  for  practical  work,  it  is 
invaluable.  As  a  matter  of  fact,  solutions  of  the  more  advanced 
problems  in  physical  mathematics  are  nearly  always  represented 
in  the  form  of  an  abbreviated  infinite  series.  Finite  solutions  are 
the  exception  rather  than  the  rule. 

Examples. — (1)  It  is  required  to  find  the  solution  dyjdx  =  y,  in  series. 
Assume  that  y  has  the  form 

y  =  a0  +  a^x  +  a^x2  +  a^  +  ... 
Differentiate,  and  substitute  for  y  and  y  in  the  given  equation, 
(a,  -  a0)  +  {2a2  -  a^x  +  (3a3  -  a2)x2  +  . . .  =  0. 
If  x  is  not  zero,  this  equation  is  satisfied  when  the  coefficients  of  x  become 
zero.     This  requires  that 

1         1  XI 

ai  —  ao  J  <^  —  2ai  ~  2ao »  az  —  qa2  =  3  jao »  •  •  • 

Hence,  by  substitution  in  (1),  we  obtain 

y  =  a0(l  +  x  +  2-jX2  +  ^}f  +  . . .  J  =  a<p*. 

Put  a  for  the  arbitrary  constant  so  that  the  final  result  is  y  —  aex.     That  this 
is  a  complete  solution,  is  proved  by  substitution  in  the  original  equation.     We 


404  HIGHER  MATHEMATICS.  §  147. 

must  proceed  a  little  differently  with  equations  of  a  higher  order.     Take  as  a 
second  example 

(2)  Solve  dy\dx  +  ay  +  bx2  =  0  in  series.     By  successive  differentiation  of 
this  expression,  and  making  y  =  y0  when  x  =  0  in  the  results,  we  obtain 

|)o=-^;(g)o=^;(S)o=-^-^;... 

By  Maclaurin's  theorem, 

=Vo  -  aVox + \<&yv&  -  U^3yo + 2bW + •  •  • ; 

c=  y0(l  -ax  +  \a?x2  -...)-  26a  - 3  {\a?x*  -  ^aV  +  ...); 

=y0e~ax  +  2ba-3  (e~ ax -1  +  ax- £a2<c2), 
by  making  suitable  transformations  in  the  contents  of  the  last  pair  of  brackets. 
Hence  finally  y=C1e~ax  -  2ba  ~  3  (1  -  ax  +  |a¥) .    Verify  this  by  the  method 
of  §  125,  page  387. 

(3)  Solve  d2yjdx2-a2y=0  in  series.      By  successive  differentiation,  and 
integration 

da?'  a  dx '  dx*~  adx**  '"  V^Wo       Vo '  ' ' "' 
when  the  integrations  are  performed  between  the  limits  x  =  x,  and  x  =  0,  so 
that  y  becomes  yQ  when  x  =  0.     From  Maclaurin's  theorem,  (1)  above,  we  get 
by  substitution 

(dy\  x       „     x2  (dy\  x* 

y=yo+\Tx)ol.+  ay»Tl+a*  \Tx)0'3\  +  '--> 

f      a2x2    aV  1      1  fdy\   (ax    asx*  } 

By  rearranging  the  terms  in  brackets  and  putting  the  constants  yQ  =  A,  and 
y0/a  =  B,  we  get, 

y=A{($  +  %ax  +  %a2x2+. .  .)  +  {i~%ax  +  ia2x2-  . . .)+  B(  .  .  .  ))• 
= IA  (eax  +  e  ~  ax)  +  B(e™  -  e  ~  ax)  =  C^e™  +  C2e~  ax. 
Sometimes  it  is  advisable  to  assume  a  series  with  undetermined  indices 
and  to  evaluate  this  by  means  of  the  differential  equation,  as  indicated  in  the 
next  example. 

(4)  Solve 

§-*!-cr/  =  ** (2) 

(i)  The  complementary  function.     As  a  trial  solution,  put  y  =  aQxm.     The 
auxiliary  equation  is 

m(m  -  ljaosc™ ~2  -  (m  +  c)aQxm  =  0.  .        .        (3) 

This  shows  that  the  difference  between  the  successive  exponents  of  x  in  the 
assumed  series,  is  2.     The  required  series  is,  therefore, 

y  =  a0a^»  +  a1xm  +  2  +  ...  +  an  +  1xm  +  2n~2+  anxm+2n  =  0,   .        (4) 
which  is  more  conveniently  written 

y=^anx™  +  **=0 (5) 

In  order  to  completely  determine  this   series,  we  must  know   three  things 
about  it.     Namely,  the  first  term ;  the  coefficients  of  x ;  and  the  different 


§  147.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.         465 

powers  of  x  that  make  up  the  series.     By  differentiation  of  (4),  we  get 

#  =  2°°  {m  +  2n)anxm  +  2n> l ;  y  =  2°°  (m  +  2ri)  (m  +  2n  -  l)anxm  +  *•  - 2. 
By  substitution  of  this  result  and  (4)  in  equation  (2),  we  have 

2  0°°{(w  +  2n)  (m  +  2n  -  l)anxm  +  **  ~  2  -  (w  +  2w  +  c)ana;m  +  ^  =  0,  (6) 
where  n  has  all  values  from  zero  to  infinity.  If  (5)  is  a  solution  of  (2),  equa- 
tion (6)  is  identically  zero,  and  the  coefficient  of  each  power  of  x  must  vanish. 
Hence,  by  equating  the  co-efficients  of  xm+*n,  and  of  xm+2n  -  2  to  zero,  we  have 

(m  +  2n)  (m  +  2n  -  VjCLnX™  +  2n  -  2  _  (w  +  2n  +  c)anxm  +  2n  =  0  ; 
and  replacing  n  by  n  -  1  in  the  second  term,  we  get 

(m  +  2n)  (m  +  2n  -  l)an  -  (m  +  2n  -  2  +  c)an  _  1  =  0 ;  .  (7) 
since  (m  +  2n)  (m  +  2n  -  1)  =  0,  when  n  =  0,  m(m  -  1)  =  0;  consequently, 
m  =  0,  or  m  =  1 ;  for  succeeding  terms  n  is  greater  than  zero,  and  the  relation 
between  any  two  consecutive  terms  is 

m  +  2n  -  2  +  c 
"*-  {m  +  2n)  {m  +  2n  -  l)a»-i-        ...        (8) 

This  formula  allows  us  to  calculate  the  relation  between  the  successive  co- 
fficients  of  x  by  giving  n  all  integral  values  1,  2,  3,  :  .  .  Let  a0  be  the  first  term. 
First,  suppose  m  =  0,  then  we  can  easily  calculate  from  (8), 
_    c  _  c  +  2a       c{c  +  2) 

ai  "  l .  2a° '  ^  "~   3  . 4    '  =      4  !      ao  5  •  •  • 

.-.  r2  =  a0{l  +  4f  +  0(0  +  2)1-,  +.-.]■•  .        .        (9) 

Next,  put  m  =  1,  and,  to  prevent  confusion,  write  6,  in  (8),  in  place  of  a. 

c  +  2t&  -  1 
°w-2n(ln  +  l)°»-i' 

proceed  exactly  as  before  to  find  successively  bv  62,  63,  .  .  . 

.-.  F9  =  60{*  +  (c  +  l)^j  +  (c  +  1)  (c  +  3)|~,  +  . . .  }.      .        (10) 

The  complete  solution  of  the  equation  is  the  sum  of  these  two  series,  (9) 
and  (10) ;  or,  if  we  put  Cxyx  =  Flf  C.2y2  =  Fa, 

V  =  Cxyx  +  Cqy2, 
which  contains  the  two  arbitrary  constants  Cx  and  02. 

(ii.)  The  particular  integral.  By  the  above  procedure  we  obtain  the  com- 
plementary function.  For  the  particular  integral,  we  must  follow  a  somewhat 
similar  method.  E.g.,  equate  (7)  to  x2  instead  of  to  zero.  The  coefficient  of 
xm-zt  in  (3)  becomes 

m{m  -  l)a0x™  -  a  =  x9, 
A  comparison  of  the  exponents  shows  that 

m  -  2  =  2  ;  and  m(m  -  l)a0  =  1 ;  .\  m  =  4 ;  o0  =  x\. 
From  (8),  when  m  =  4, 

2  +  2n  +  c 
an  ~  2>  +  2)  (2n  +  S)0*"1' 
Substitute  successive  values  of  n  =  1,  2,  3,  ...  in  the  assumed  expansion,  and 
we  obtain 

Particular  integral  =  a0xm  +  axxm  +  2  +  aacc™  +  4  +  . . . , 
where  a0,  Oj,  a^  .  .  .  and  m  have  been  determined. 


46G  HIGHER  MATHEMATICS.  §  147. 

(5)  The  following  velocity  equations  have  been  proposed  for  the  catalytic 
action  of  an  enzyme  upon  salicine  (J.  W.  Mellor's  CJiemical  Statics  and  Dyn- 
amics, London,  380,  1904)  :— 

dy  dx 

-j-t  =  kx(a  -  x-  y)(c  -  y) ;    -^  =  k2y.  .        .        .        (11) 

From  Maclaurin's  theorem, 

fdx\  /d*x\    f2 

x  =  x°  +  {di)0t+\dr*)02-i  + (12> 

Hence,  when  x  =  0,  and  y  =  0,  equations  (11)  furnish 

(§)0  =  (k*y)°  =  °  ;  •'•  (l)o  =  k^  J  (^2)o  =  klKaC  =  A'  "^       (13) 
By  differentiation  of  the  first  of  equations  (11),  we  get 

(g)r-^--,-w|-,l(c-„(|  +  f).,  .     ,», 

and  from  the  second  of  equations  (11),  (13),  and  (14), 

-ftp  J   =  ~  ^i2Mc(a  +  c)  =  B,  say.     .        .        .        (15) 
Again,  differentiating  (14), 

d*y         f/dx    dy\dy  d*y      dyfdx     dy\  (d?x     d*y\\ . 

W  =  H\di+dt)di  ~(a~x  -v)w  +  dt[dt+Tt)  +  <c  ~  V\W  +  d&)f  ' 

\S)    =  2fciRa2c2  +  ki3a2c{a  +  c)  -  k^k^ac   -  k*ac{a  +  c)}; 
=  ZkfaW  +  k?a?c{a  +  c)  -  k^Kac*  +  k^ac*(a  +  c) 
=  k?ac{a  +  c)2  +  2fc13a2c2  -  k2k2ac\ 

(dAx\ 
-^j  )   =  k*k2ac{a  +  c)  +  ZkfktfLW  -  k^k^ac*  +  C,  say.         .        (16) 

Consequently,  from  (12),  (13),  (15),  and  (16),  and  collecting  the  constants  to- 
gether, under  the  symbols  A,  B,  C,  .  .  .  we  get 

J2       -r,t3       „#  A    t       B    t2       G    ts 

x  =  A2i  +  B3-i  +  cri  +  '->  -'•y  =  k2'i  +  hi'2i  +  k2'sT+"'  <17> 

We  have  expressed  x  and  y  in  terms  of  t  and  constants. 

A  great  number  of  the  velocity  equations  of  consecutive  chemical 
reactions  are  turned  out  by  the  integral  calculus  in  the  form  of  an 
infinite  series.  If  the  series  be  convergent  all  may  appear  to  be 
well.  But  another  point  must  here  be  emphasized.  The  constants 
in  the  series  are  evaluated  from  the  numerical  data  and  the  agree- 
ment between  the  calculated  and  the  observed  results  is  quoted  in 
support  of  a  theory.  As  a  matter  of  fact  the  series  formula  is  quite 
empirical.  Scores  of  hypotheses  might  be  suggested  which  would 
all  furnish  a  similar  relation  between  the  variables,  and  "  best 
values  "  for  the  constants  can  be  determined  in  the  same  way. 
Of  course  if  it  were  possible  to  evaluate  the  constants  by  independ- 
ent processes,  and  the  resulting  expression  gave  results  in  harmony 
with  the  experimental  material,  we  might  have  a  little  more  faith 


§  147.     HOW  TO  SOLVE  DIFFERENTIAL  EQUATIONS.        467 

in  the  theory.  These  remarks  in  no  way  conflict  with  the  dis- 
cussion on  page  324.  There  the  constants  were  in  questions,  here 
we  speak  of  the  underlying  theory. 

But  we  are  getting  beyond  the  scope  of  this  work.  I  hope 
enough  has  been  said  to  familiarize  the  student  with  the  notation 
and  ideas  employed  in  the  treatment  of  differential  equations  so 
that  when  he  consults  more  advanced  books  their  pages  will  no 
longer  appear  as  "  unintelligible  hieroglyphs  ".  For  more  extensive 
practical  details,  the  reader  will  have  to  take  up  some  special  work 
such  as  A.  E.  Forsyth's  Differential  Equations,  London,  1903; 
W.  E.  Byerly's  Fourier's  Series  and  Spherical  Har?nonics,  Boston, 
1895.  H.  F.  Weber  and  B.  Eiemann's  Die  Partiellen  Diffeiential- 
Gleichungen  der  Mathematischen  Physik,  Braunschweig,  1900-1901, 
is  the  text-book  for  more  advanced  work.  A.  Gray  and  G.  B. 
Mathews  have  A  Treatise  on  Bessel's  Functions  and  their  Applica- 
tion to  Physics,  London,  1895. 

II.  Method  of  successive  approximations.  This  method  re- 
sembles, in  principle,  that  used  for  the  approximate  solution  of 
numerical  equations,  page  358.  When  some  of  the  terms  of  the 
given  equation  are  small,  solve  the  equation  as  if  these  terms  did 
not  exist.  Thus  the  equation  of  motion  for  the  small  oscillations 
of  a  pendulum  in  air, 

d20  fdd\2  d26 

dp+<F0=  a\dt)  '  becomes  dfi  +  ^  =  °'      •        (18) 

provided  the  right  member  a(dO/dt)2  is  small.     Solving  the  second 

of  these  equations  by  the  method  of  page  401,  we  get  0  =  -  r  cos  qt. 

If  so  then  a(dO/dt)2  must  be  ar2q2sm2qt.     Substituting  this  in  the 

first  of  equations  (18),  and  remembering  that  1  -  cos  2x  m  2  sin2#, 

page  612,  we  get 

d26       .„      ar2q2/  -     \  d20         /       ar2\         aq2r2      n 

-^  +  q*Q  =  -£-  (1  -  cos  2qtJ  j  or,  ^  +  2 y  ~  ^ )  =  "     2    C0S  2^' 

This  gives  $  =  \ar2  +  (r  -  far2)  cos  qt  +  \ar2  cos  2qt,  when  solved, 
as  on  page  421,  with  the  conditions  that  when  t  =  0,  0  =  r,  and 
dO/dt  =  0. 

Example. — A  set  of  equations  resembling  those  of  Ex.  (5),  of  the  preced- 
ing set  of  examples  is  solved  in  Technics,  1,  514,  1904,  by  the  method  of  suc- 
cessive approximation,  and  under  the  assumption  that  kx  and  a  are  small  in 
comparison  with  ft2  and  c.  Hint.  Differentiate  the  second  of  equations  (11) ; 
multiply  out  the  first ;  and  on  making  the  proper  substitution 
1  d?x      kJdx\*      kJ  \dx      ,    , 

5  d?  =  k2{Tt)  +  kXx  -c-a)dt  +  k^a  ~x)"       '       (19) 

GG  * 


468  HIGHER  MATHEMATICS.  §  147. 

Neglect  terms  in  kjk^,  and  d2(x  -  a)/dt2  +  q2(x  -  a)  =  0  remains,  if  <f  be  put  in 

place  of  k^k^     Hence,  x  -  a  =  -  a  cos  qt,  since  x  and  y  are  both  zero  when 

t  =  0.     Differentiate  and  substitute  the  results  in  (19).     We  get 

d2x  fc, 

-372  +  <Z2(&  -  o)  =  -  ^(a  cos  qt  -  c)aq  sin  gtf  +  T~a222  sin2g£ ; 

d2a;  a2qk,  a2q2k, 

•"*  ^  +  22^  -  a)  =  23sin  2*  -  ~^sin  22*  +  "2ft""1  *  cos22*)> 

which  can  be  solved  by  the  method  of  page  421.     The  complete  (approximate) 

solution — particular  integral  and  complimentary  function — is 

q2t  cos  qt      cPk-L  sin  2  qt      a2kx  cos  2qt      a2kx 
x-a  =  +acoSqt- g + ^ + ^- +  ^-. 

This  solution  is  not  well  fitted  for  practical  work.     It  is  too  cumbrous. 


CHAPTEE  VIII. 

FOURIER'S   THEOREM. 

"  Fourier's  theorem  is  not  only  one  of  the  most  beautiful  results  of 
modern  analysis,  but  may  be  said  to  furnish  an  indispensable 
instrument  in  the  treatment  of  nearly  every  recondite  question 
in  modern  physics.  To  mention  only  sonorous  vibrations,  the 
propagation  of  electric  signals  along  a  telegraph  wire,  and  the 
conduction  of  heat  by  the  earth's  crust,  as  subjects  in  their 
generality  intractable  without  it,  is  to  give  but  a  feeble  idea  of 
its  importance." — Thomson  and  Tait. 

§  148.  Fourier's  Series. 

Sound,  as  we  all  know,  is  produced  whenever  the  particles  of  air 
are  set  into  a  certain  state  of  vibratory  motion.  The  to  and  fro 
motion  of  a  pendulum  may  be  regarded  as  the  simplest  form  of 
vibration,  and  this  is  analogous  to  the  vibration  which  produces  a 
simple  sound  such  as  the  fundamental  note  of  an  organ  pipe. 
The  periodic  curve,  Fig.  52,  page  136,  represented  by  the  equations 
y  =  sin  x ;  or  y  =  cos  x,  is  a  graphic  representation  of  the  motion 
which  produces  a  simple  sound. 

A  musical  note,  however,  is  more  complex,  it  consists  of  a 
simple  sound — called  the  fundamental  note — compounded  with  a 
series  of  auxiliary  vibrations — called  overtones.  The  periodic  curve 
of  such  a  note  departs  greatly  from  the  simplicity  of  that  represent- 
ing a  simple  sound.  Fourier  has  shown  that  any  periodic  curve 
can  be  reproduced  by  compounding  a  series  of  harmonic  curves 
along  the  same  axis  and  having  recurring  periods  1,  \y  £,  \, . . .  th 
of  the  given  curve.  The  only  limitations  are  (i)  the  ordinates  must 
be  finite  (page  243) ;  (ii)  the  curve  must  always  progress  in  the  same 
direction.  Fourier  further  showed  that  only  one  special  combina- 
tion of  the  elementary  curves  can  be  compounded  to  produce  the 
given  curve.     This  corresponds  with  the  fact  observed  by  Helm- 

469 


470  HIGHER  MATHEMATICS.  §  149. 

holtz  that  the  same  composite  sound  is  always  resolved  into  the 
same  elementary  sounds.  A  composite  sound  can  therefore  be  re- 
presented, in  mathematical  symbols,  as  a  series  of  terms  arranged, 
not  in  a  series  of  ascending  powers  of  the  independent  variable,  as 
in  Maclaurin's  theorem,  but  in  a  series  of  sines  and  cosines  of 
multiples  of  this  variable. 

Fourier's  theorem  determines  the  law  for  the  expansion  of  any 
arbitrary  function  in  terms  of  sines  or  cosines  of  multiples  of  the 
independent  variable,  x.  If  f(x)  is  a  periodic  function  with  respect 
to  time,  space,  temperature,  or  potential,  Fourier's  theorem  states 
that 

f(x)  =  A0  +  axsin  x  +  a2sin  2x+  ...  +  ^cos  x  +  62cos  2x  + . . .       (1) 

This  is  known  as  Fourier's  series.  It  is  easy  to  show,  by 
plotting,  as  we  shall  do  later  on,  that  a  trigonometrical  series  like 
that  of  Fourier  passes  through  all  its  changes  and  returns  to  the 
same  value  when  x  is  increased  by  2w.  This  mode  of  dealing  with 
motion  is  said  to  be  more  advantageous  than  any  other  form  of 
mathematical  reasoning,  and  it  has  been  applied  with  great  success 
to  physical  problems  involving  potential,  conduction  of  heat,  light, 
sound,  electricity  and  other  forms  of  propagation.  Any  physical 
property — density,  pressure,  velocity — which  varies  periodically 
with  time  and  whose  magnitude  or  intensity  can  be  measured, 
may  be  represented  by  Fourier's  series. 

In  view  of  the  fact  that  the  terms  of  Fourier's  series  are  all 
periodic  we  may  say  that  Fourier's  series  is  an  artificial  way  of 
representing  the  propagation  or  progression  of  any  physical  quality 
by  a  series  of  waves  or  vibrations.  "It  is  only  a  mathematical 
fiction,"  says  Helmholtz,  "  admirable  because  it  renders  calcula- 
tion easy,  but  not  necessarily  corresponding  with  anything  in 
reality." 

§  159.    Evaluation  of  the  Constants  in  Fourier's  Series. 

Assuming  Fourier's  series  to  be  valid  between  the  limits  x  =  +  7r 
and  x  =  -  7r,  we  shall  now  proceed  to  find  values  for  the  co- 
efficients A0i  av  a2, .  .  .  ,  bv  b2,  . .  .  ,  which  will  make  the  series 
true. 

I.  To  find  a  value  for  the  constant  A0.  Multiply  equation  (1) 
by  dx  and  then  integrate  each  term  between  the  limits  x  *=  +  it  and 
x  =  -  7T.     Every  term  involving  sine  or  cosine  term  vanishes,  and 


§  149.  FOUKIER'S  THEOREM.  471 

2irAQ  =  J  _  f(x) .  ^ ;  or,  A0  =  ^J  _  /(a?) .  £te,        .         (2) 

remains.  Therefore,  when  /(a?)  is  known,  this  integral  can  be 
integrated.1 

I  strongly  recommend  the  student  to  master  §§  74,  75,  83  before 
taking  up  this  chapter. 

II.  To  find  a  value  for  the  coefficients  of  the  cosine  terms,  say 
bn,  where  n  may  b9  any  number  from  1  to  n.  Equation  (1)  must 
not  only  be  multiplied  by  dx,  but  also  by  gome  factor  such  that  all 
the  other  terms  will  vanish  when  the  series  is  integrated  between 
the  limits  +  7r,  6ncos  nx  remains.  Such  a  factor  is  cosnx.dx. 
In  this  case, 

r+7r 

I     cos2rac .  dx  —  bnTrt 

(page  211),  all  the  other  terms  involving  sines  or  cosines,  when 
integrated  between  the  limits  ±  7r,  will  be  found  to  vanish.  Hence 
the  desired  value  of  bn  is 


If+TT 


cosnx.dx.     .  .        (3) 

This  formula  enables  any  coefficient,  bv  b2,  . . . ,  bn  to  be  obtained. 
If  we  put  n  =  0,  the  coefficient  of  the  first  term  A0  assumes  the 
form, 

4>-i&o & 

If  this  value  is  substituted  in  (1),  we  can  dispense  with  (2),  and 

write 

f{x)  m  ±bQ  +  a^sin  x  +  ^cos  x  +  a2s:n  2x  +  b2cos  2x  +  . . .    (5) 
III.  To  find  a  value  for  the  coefficients  of  the  sine  terms,  say  an. 

As  before,  multiply  through  with  amnxdx  and  integrate  between 

the  limits  +  v.     We  thus  obtain 

an  =  -      f(x)  sin  nx  .dx.     .        .         .         (6) 

T?  J  —  n 

Examples. — (1)  Problems  like  these  are  sometimes  set  for  practice.  Put 
-jp  =  x,  in  (1),  and  develop  the  curve  Fig.  155.  \W 

Note  T  is  a  special  value  of  t.     The  series  to  be  _j__ 

developed  is  it— — t- — »j 

f(t)  =  A0+  oj  sin  ~r  +  . . .  +  bx  cos  ~  +  . . .  Fig.  155. 

1 1  have  omitted  details  because  the  reader  should  find  no  difficulty  in  working 
out  the  results  for  himself.  It  is  no  more  than  an  exercise  on  preceding  work — page 
211. 


time 


472 


HIGHER  MATHEMATICS. 


§149. 


To  evaluate  A0,  multiply  by  the  periodic  time,  i.e.,  by  T,  as  in  (^ 
between  the  limits  0  and  T.     From  page  211. 


and  integrate 


1    FT  v 

4o  -  t\    f®dt  =  AveraSe  neight  of  f{t)  =  ^ . 


y  +  3  sm 


6wt      1    .     IOtt*  \ 

~T  +sBmT~+'")' 


For  the  constants  of  the  cosine  and  sine  terms,  multiply  respectively  by 
cos  (2-7rtlT)dt,  and  by  sin  (2irtjT)dt  and  integrate  between  the  limits  0  and 
T.     The  answer  is 

,„,       y      27/  .     2tt*      1 

Remember  sin  2nir  is  zero  if  n  is  odd  or  even  ;  cos  2mr  is  +  1  if  n  is  even,  and 
-  1  if  ?t  is  odd  or  even.     The  integration  in  this  section  can  all  be  done  by 
the  methods  of  §§  73  and  75.     Note,  however,  jx  sin  nxdx  =  n~2(sin  nx  -  nx 
,cos  nx),  on  integration  by  parts. 

(2)  In  Fig.  156,  the  straight  lines  sloping  downwards  from  right  to  left 

fit) 


have  the  equation  f(t)  =  At,  where  m  is  a  constant.  When  t  =  T,  f(t)  =  7,  so 
that  7=  raT,  or  m  =  VjT;  .-.  f(t)  =  Yt\T.     Hence  show  that 

yrr  y  2  f'Vt   .     2w*  7 

Ao  =  -rji]    t-dt=2>ai=T      2rSm  ~T    t  =  ~* » 
and  also 

'l,'        7/tt         .     2irt       1     .      4tt£      1  6ir*  \ 

/W=^2-sm'2r"28in~2r~3Bm  ~T   ~  "'J' 
(3)  In  Fig.  157,  you  can  see  that  AQ  is  zero  because 

1  fT  2  CT  2irt 

Aq=TJ0  /(0^  =  Average  height  of  f(t)  =  0;  a1=  ^ J   f(t)  sin  -^dt. 

Now  notice  that  f(t)  =  mt  between  the  limits  -  JTand  +  £T;  and  that  when 
t  =  £T,f(t)  =  a  so  that  a  =  \mT;  and  m  =  4a/!F;  while  between  the  limits 
t  =  IT,  and  *  =  f  T,  /(*)  =  2a(l  -  2t;  T) ;  hence, 


2   r+WAat    .     2»*         2  ftr      / 
^=Tj.iT'TSmT-dt+TjiT2a{ 


-I 


.    27ri  SJ      8a 
sm  -Tn-rf^  =  -3. 


In  a  similar  manner  you  can  show  that  air"  the  even  a's  vanish,  and  all  the 
b's  also  vanish.     Hence 


8a/  .     2tt« 
/(*)=TV8in"2r 


1    .     6tt£ 
k  sm 


mt         1      .     10*4  \ 


There  are  several  graphic  methods  for  evaluating  the  coefficients  of  a 
Fourier's  series.  See  J.  Perry,  Electrician,  28,  362,  1892  ;  W.  B.  Woodhouse, 
the  same  journal,  46,  987,  1901 ;  or,  best  of  all,  O.  Henrici,  Phil.  Mag.  [5],  38, 
110,  1894,  when  the  series  is  used  to  express  the  electromotive  force  of  an 
alternating  current  as  a  periodic  function  of  the  time. 


§150. 


FOURIER'S  THEOREM. 


473 


§  150.  The  Development  of  a  Function  in  a  Trigonometrical 

Series. 

I.  The  development  of  a  trigonometrical  series  of  sines.    Suppose 
it  is  required  to  find  the  value  of 

f(x)  =  x, 
in  terms  of  Fourier's  theorem.     From  (2),  (3)  and  (6), 


^n  =  ~~  I     &  •  cos  nx  .  dx  =  0  ;  an  =  -       x .  si 

W  J  -  n  T  J  -  ir 

according  as  n  is  odd  or  even  ; 


sin  nx  .dx  =  +  -, 
—  % 


Ao~  2 


x  .  dx  =  -7—  (tt2 

4.7TV 


ir*)«0. 


(7) 


Hence  Fourier's  series  assumes  the  form 

x  =  2(sin  x  -  \  sin  2x  +  J  sin  Sx  -  . . .), 

which  is  known  as  a  sine  series ;   the  cosine  terms  have  dis- 
appeared during  the  integration. 

By  plotting  the  bracketed  terms  in.  (7)  we  obtain  the  series 
of  curves  shown  in  Fig.  158.  Curve  1  has  been  obtained  by- 
plotting  y  =  sin  x;  curve  2,  by  plotting  y  =  %  sin  2x  \  curve  3,  from 


4/  y 

Fig.  158.— Harmonics  of  the  Sine  Curve. 

y  =  J  sin  Sx.  These  curves,  dotted  in  the  diagram,  represent  the 
overtones  or  harmonics.  Curve  4  has  been  obtained  by  drawing 
ordinates  equal  to  the  algebraic  sum  of  the  ordinates  of  the  pre- 


474 


HIGHEK  MATHEMATICS. 


§150. 


ceding  curves.     The  general  form  of  the  sine  series  is 

f(x)  =  d^sina;  +  a2sin  2x  +  a3sm  3x  +  .  . .,        .         (8) 
where  a  has  the  value  given  in  equation  (6). 

II.  The  development  of  a  trigonometrical  series  of  cosines.     In 
illustration,  let 

f{x)  =  x2, 
be  expanded  by  Fourier's  theorem.     Here 

1  r+  n  _  4 

b n  =  —  I      x2 .  cos  nx .  dx  =  +  —s, 
n      ttJ-tt  n2' 

according  as  n  is  odd  or  even.     Also, 

an  =  -\      #2sin  nx 

Hence, 

x2  =  K7T2  -  4fcos#  -  O2cos2ic  +-O2cos3ic  -  .  .A  .         (9) 

By  plotting  the  first  three  terms  enclosed  in  brackets  on  the  right 
side  of  (9),  we  obtain  the  series  of  curves  shown  in  Fig.  159.  The 
general  development  of  a  cosine  series  is 

f(x)  =  lb0  £  b^osx  +  62cos2#  +  ...,.         .       (10) 
where  b  has  the  values  assigned  in  (3).     As  a  general  rule,  any  odd 


.dx  =  0;   ^0=cH      V^  =  ^J7r3-(-7r)3|=  7T2. 


Fig.  159. — Harmonics  of  the  Cosine  Curve. 

function  of  x  will  develop  into  a  series  of  sines  only,  an  even  function 
of  x  will  consist  of  a  series  of  cosines.  An  even  function  of  x  is 
one  which  retains  its  value  unchanged  when  the  sign  of  the  vari- 
able, x,  is  changed.  E.g.,  the  sign  of  x2  is  the  same  whether  x  be 
positive  or  negative;  cos  a?  is  equal  to  cos(-  x),  page  611,  and 
therefore   x2   and   cos  a;    are    even   functions   of   x.       If   f(x)    is 


§150. 


FOURIEK'S  THEOREM 


475 


an  even  function  of  x,  f(x)  =  /(-#).     An   odd   function  of  x 

changes  its  magnitude  when  the  sign  of  the  variable  x  is  changed- 
Thus,  x,  x3, . .  .  and  generally  any  odd  power  of  an  odd  function, 
since  sin  x  =  -  sin  ( -  x),  sin  a;  is  an  odd  function  of  x  ;  generally 
if  f(x)  is  an  odd  function  of  x,  f{x)  =  -  /( -  x).  In  (8)  f(x)  is  an 
odd  function  of  x,  and  in  (10),  f(x)  is  an  even  function  of  x  between 
the  limits  -  tt  and  +  tt. 

Examples. — (1)   Develop  unity  in  a  series  of  sines  between  the  limits 

x  =  tt  and  x  =  0.      Here  f(x)  =  1.      Now   perform  the   integrations  with 

n  =  1,  2,  3, . . .  and  you  will  see  that 

2  /V  2  2  4 

On  =  —  /    sin  nxdx  =  — (1  -  costtir)  =  — (1  -  (-  l)n)  =  — ,  or,  0, 
irj  o  nirK  '       mrx         v        '  s       tt?r'      '     ' 

according  as  n  is  odd  or  even.     Hence,  from  (8), 

1  =  -(since  +  qsin  3x  +-g sin 5x  +  ..  A  .        .       (11) 

The  first  three  terms  of  this  series  are  plotted  in  Fig.  160  in  the  usual  way. 
(2)  Show  that  f or  x  =  ±  tt 

2sinh7r|7l     1  1  \      /l  2    .  W"    M 

e*= i  (  2-o"cosa;  +  5cos2a;  +  ««  •  )  +  (  osmx_  gSin2a;  +  ...  J  >.     (12) 


Fig.  160. — Harmonics  of  the  Sine  Development  of  Unity. 

(3)  x  sin  x  =  1  -  \ cos x  -  \ cos  1x  +  \ cos Sx  -  ft cos ±x  +  ... 
Establish  this  relation  between  the  limits  tt  and  0.     If  x  =  -^r,  then 


(4)  Show       x  =  5- (  cos  x  +  q-cos  3a?  +  Hkcos  5x  +  . . 

bn  =-\    x  cos  nx  .  dx  =  -5- (cos  mr  -  1)  =  ——  {(  -  1)»  -  1}. 


(13) 


(H) 


476 


HIGHER  MATHEMATICS. 


§150 


(5)  Show  that  if  c  is  constant, 

4c/  1  1  \ 

c  =  — (  sina?  +  ^smSa;  +  gsm5aj  +  . . .  ).  .        .       (15) 

III.  Comparison  of  the  sine  and  the  cosine  series.  The  sine  and 
cosine  series  are  both  periodic  functions  of  x,  with  a  period  of  2tt. 
The  above  expansions  hold  good  only  between  the  limits  x  =  +  ir, 
that  is  to  say,  when  x  is  greater  than  -  tt,  and  less  than  +  71- . 
When  x  =  0,  the  series  is  necessarily  zero,  whatever  be  the  value 
of  the  function.  As  a  matter  of  fact  any  function  can  be  re- 
presented both  as  a  sine  and  as  a  cosine  series.  Although 
the  functions  and  the  two  series  will  be  equal  for  all  values 
7r  and  x  =  0,  there  is 


of  x  between  x  = 


y 

\ 

K 

/ 

\/  oc 

-7T 

0 

y 

n 

difference  between 
the  sine  and  cosine 
developments  for 
other  values  of  x. 
For  instance,  com- 
pare the  graph  of  x 
when  developed  in 
series  of  sines  and 
series  of  cosines 
Pig.  161. — Diagrammatic  Curve  of  the  Cosine  Series,    between  the  limits 

x  =  0  and  x  =  tt,  as  shown  in  Figs.  158  and  159  above.  Plot 
these  equations  for  successive  values  of  x  between  +  7r,  etc.  In  the 
case  of  the  cosine  curve  we  get  the  lines  shown  diagrammatically 
in  Fig.  161.     By  tracing  the  curves  corresponding  with  still  greater 

values  of  tt,  we  get 
the    dotted    lines 
shown  in  the  same 
figure.      For   the 
sine  curve  we  get 
lines  shown   dia- 
grammatically  in 
Fig.  162.  Note  the 
isolated       points, 
Fig.  162. — Diagrammatic  Curve  of  the  Sine  Series.        of    page    171,    for 
x  =  ±  ir,  y  =  0  ;  x  =  +  37r,  y  =  0  ; .  . .  Both  these  curves  coincide 
with  y  =  x  from  x  =  0  to  x  =  tt,  but  not  when  x  is  less  than  -  7r, 
and  greater  than  +  7r. 

Instead  of  taking  the  final  results  bodily  from  the  text-books,  we 
had  better  get  some  practice  in  the  work  by  following  up  the  regular 
proofs. 


I 

^ 

/ 

/ 

• 

/ 

X' 

/   ac 

/*            -7r 

0 

IT 

/ 

/ 

/ 

/ 

/ 

/ 

r 

/ 

§  151.  FOURIER'S  THEOREM.  477 

§  151.    Extension  of  Fourier's  Series. 

Up  to  the  present,  the  values  of  the  variable  in  Fourier's  series 
only  extend  over  the  range  +  v.  The  integration  may  however 
be  extended  so  as  to  include  all  values  of  x  between  any  limits 
whatever. 

I.  The  limits  are  x  =  +  c,  x  =  -  c.  Let  f(x)  be  any  function 
in  which  x  is  taken  between  the  limits  -  c  and  +  c.  Change  the 
variable  from  x  to  cz/7r,  so  that  z  =  -n-x/c.     Hence, 

/(*) =/£*)•   •    •    •    •  (ifi) 

When  x  changes  from  -  c  to  +  c,  z  changes  from  -  -k  to  +  ir,  and, 
therefore,  for  all  values  of  x  between  -  c  and  +  c,  the  function 
([cz/tt)  may  be  developed  as  in  Fourier's  series  (5),  or 

f(~z)  =  %b0  +  ^cos  z  +  opsins  +  £2cos  2z  -f  <x2sin  2z  +  . . .     (17) 

where, 

bn  =  —  I      f(~z)  cos nz . dz ;  an  =  -\      f( -z)  sin nz  .dz.    (18) 

This  development  (17)  is  true  from  +  ?r  to  -  tt.  From  (16),  there- 
fore, 

„   x         ^i  r  irX  .     ttX        .  2irX 

&  Kx)  =  2bo  +  bicoa  V  +  aiam~c~  +  b2C08~  +  •  •  •  (19) 
The  coefficients  a  and  b  are  the  same  as  in  (17),  and  consequently 
(19)  holds  good  throughout  the  range  ±  c.     From  (18),  we  have 


1  f +  C  ,,    x         nirX  ,  1  f +  °  „    N   .     nirX  . 

n  =  c  J  _  /^cos  ~c~dx  ■  a"  =  c  J  _  /^)sm  ~dx' 


(20) 


Hence  the  rule:  Any  arbitrary  function,  whose  period  ranges  from 
-  c  to  +  c,  so  that  T  =  2c,  can  be  represented  as  a  series  of  trigono- 
metrical functions  with  periods  T,  ^T,  JT '. . . 

Examples. — (1)  From  (19),  show  that  the  sine  series,  from  x=0  to  x=c, 
is 

..  .              .     irX              .     2irX              .     SirX  .-. 

f(x)  =  OjSin h^sin +  038111 + (21) 

2  f.,  >  .   nirx , 

an=-    f(x)sm dx (22) 

CJ  0  c 

And  for  the  cosine  series,  from  x  =  0  to  x  =  c,  we  have 

,.  .       1,  ,  7rCC       ,  2irX  ._•; 

f(x)  =  2  °  +  6lC0S T  +   2C0S ~~c~  +  '•'       '       •       *23) 

&n  =  |j'/(x)cos^.         .         .  .         .  (24) 


478                              HIGHER  MATHEMATICS.  §  151. 

(2)  Prove  the  following  series  for  values  of  x  from  x  =  0  to  x  =  c : — 

4c/   .     irX        1     .      SttX       1     .      birX               \  .-„, 

c  =  —  sin —  +  osm +xsm —  +  ...)•  .        .        (25) 

v\        c        3           c         5           c               /  *     ' 

2m.c/  .    ttx      1    .     2-Tra;      1    .     Swx  \ 

mx  =  _/sm__._sin_  +  _sm  —  -... ).  .        .         (26) 


W;r       4ttl/        irOJ        1  SirX 


4tmf      xx       1         3ttx       1    .     5kx  \  re% 

__^003_  +  _00s_.  +  ^_.  +  ...j.       (27) 

Hint.  (26)  is  f(x)  =  mx  developed  in  a  series  of  sines ;  (27)  the  same  function 
developed  in  a  series  of  cosines. 

2c/         itX        1  2irX  \ 

(3)  If  f(x)  =  x  between  +  c;  x  =  —(sin ^  sin +  •••).  (28) 

II.  The  limits  are  +  oo  and  -  oo.  Since  the  above  formulae 
are  true,  whatever  be  the  value  of  c,  the  limiting  value  obtained 
when  c  becomes  infinitely  great  should  be  true  for  all  values  of  x. 
Let  us  look  closer  into  this,  and  in  order  to  prevent  mistakes  in 
working,  and  to  show  that  equations  (20)  have  been  integrated,  we 
may  write,  as  indicated  on  page  232, 

bn  =  c[j^)c0S^?H_c;  an  =  c[f^sin^^]_c; 
but  it  is  more  convenient  to  put  A.  in  place  of  x  to  denote  that  the 
expression  has  been  integrated.     Accordingly,  we  get 

2.  M+\^\  W"^  lf+%/^       •       W77-X 

n  =  c J  _  /(  }  °0S  ~c~ dk ;  an  =  c)_  /(A)  Sm  Tdk-  <29) 
Substitute  these  values  of  av  a2,  .. .,  b0,  bv  . .  .,  in  (19),  and  we 
get,  by  the  series  of  trigonometrical  transformations,  (24),  (13),  (6), 
page  609  et  seq., 

m  -  J[i£/(A)dA  +  j^/W^cosfdA  +  •  •  •) 

+  £]{/W  sin  £  sin  ^dK  +...}]; 
lf+c,/  V7    fl        /        irk  ttX         .     ttA.    .     7nZ\  1 

=  cJ./W^l2  +  Vcos  T  cos  T  +  sm  T  sin  TJ  +  •  •  7 ; 
=  ^£/W^{|  +  C0S^A  -  x)  +  cos  "f(X '"  *)  +  "•  •  •} ' 
=  ^      /(A)dAJl  +  2cos^(X  -  aj)  +  2cosy(\  -  a?)  +.  .  .1; 

=  ^/W^l1  +  C°S  c(X  "  X)  +  C0S  (  "  ~c){X  ~  X)  +  '  '  '  \  ; 

1    f+C,/     x    -,      f7**  07T.  .  7T  IF,.  . 


7T 

+  —cos 
c 


§  151.  FOURIER'S  THEOREM.  4^9 

As  g  is  increased  indefinitely,  the  limiting  value  of  the  term  in 
brackets  is  |      cos  — (A.  -  x)d — .     Let  a  =  — ,  n  being  any  integer, 

J-  00  G  G  .         c 

f(x)  assumes  the  form 

1    f  +  oo  f+  00 

f(x)  =  yA        fWdA         COSa(\  -  X)da,.  (30) 

for  all  values  of  x.      The  double  integral  in  (30)  is   known  as 
Fourier's  integral. 

It  is  sometimes  convenient  to  refer  to  the  following  alternative 
way  of  writing  Fourier's  series  : 

f(x)=-\j(\).d\+-2^n  =  i™s—03s-v-f(\).d\,     (31) 

true  for.. any. value  of  x  between  0  and  c. 

Example. — Find  an  expression  equal  to  v  when  x  lies  between  0  and  a, 
and  equal  to  zero,  when  x  lies  between  a  and  b.     Here  f(\)  =  v,  from  A.  =  0  to 

A  =  a,  and  f(\)  =  0,  from  \  =  a  to  \  •=  b;  c  =  b;  cos  ~r-f(\) .  d\,  becomes 

fa      nir\ ,  vb    ,    mra     __    "     -  .     ,  . 

v  I   cos  —r-d\,  or,  —  sin  — j—.     Hence  the  required  expression  is, 

va      2vf       ira         -kx       1    .     2ira  2irx  \ 

/(*)  =  T  +  ~7\*inT'005T  +  2 ain  IT- cos ~b~  +  •  •  '/■ 

when  x  =  a,  this  expression  reduces  to  £v. 

JIT.  Different  forms  of  Fourier's  integral.     Fourier's  integral 
may  be  written  in  different  equivalent  forms.     From  page  241, 

J+  *  ro  r°° 

cos  xdx  =  I       cos  xdx  +  I    cos  xdx  ; 
-  oo  J  -  oo  J0 

ro  ro  ro 

cos#6&e  =  I     cos(-  x)d(-  x)  =  -  1    eosa^a?; 

J  —   00  J  oo  J  oo 

/•+  oo  p 

I       cos  xdx  =  2  I    oosxdx. 

Hence,  we  may  write  in  place  of  (30), 

\AX)  =  ~  I    °°fWd\  [    cos  a(A.  -  x)da,  .         .       (32) 
U  *"  "  °*      — ^°~~ — — - 

where  the  integration  limits  in  (32)  are  independent  of  a  and  A,  and 
therefore  the  integration  can  be  performed  in  any  order.     Again, 
if  f{x)  is  an  even  or  an  odd  function,  (32)  can  be  simplified, 
(i)  Let  f{x)  be  an  odd  function  of  x,  page  475,  so  that 

f(x)  =  -  f(-  X),  or,  -/(*)-/(-*). 


480  HIGHER  MATHEMATICS.  §  151. 

then,  by  means  of  the  trigonometrical  transformations  of  page  611, 
and  the  results  on  page  241, 

!(•+<»  r+co  .  lf+00       f+°° 

fix)  =  -\      fiX)dX\      cos  a(X  -  x)da  =  - 1       da  I      fiX)  cos  a(A  -  x)dX ; 
if  J  - «  Jo  ^  J  o         J  - » 

1  f+  °°     ro  r  °° 

=  -l      da  I      /(A)cosa(A  -  #)dA.  +  I  /(A)cosa(A.  -  #)d\  ; 

1  f  +  »      fo  f00 

f(x)  =  -\      da\  fi-\)oosa(-\-x)d(-\)+\    /(A)cosa(X  -  x)dX\ 

^  J  0  J  oo  Jo 

-I     /»00  /  —  00  -V  -00 

=  -     daj  -     /(X)cosa(A  +  a)dAV  +     /(X)cosa(X  -  a?)dA; 

=  -l    da  I    /(A.KC0Sa(A  -  X)  -  C0Sa(X  +  x)  \dX\ 

2  f  °°      f  °° 

=  - 1   da  I  /(A.)  sinaX .  sin  ax  ,dX; 

2  poo  -oo 

•,  fix)  =  - 1  f(X)dX  I   sin  aX  .  sin  ax  .da, .         .       (33) 
''"Jo  Jo 

which  is  true  for  all  odd  functions  of  fix)  and  for  all  positive  values 
of  x  in  any  function. 

(ii)  Let  fix)  be  an  even  function  of  x,  page  474,  so  that 

We  can  then  reduce  (32)  in  the  same  way  to  the  Fourier's  integral 

2-oo  -00 

fix)  =  - 1  fiX)dX  I    COS  aX  .  cos  ax .  da,   .  .       (34) 

which  is  true  for  all  values  of  x  when  fix)  is  an  even  function  of  x 
and  for  all  positive  values  of  x  in  any  function. 

Although  the  integrals  of  Fourier's  series  are  obtained  by  inte- 
grating the  series  term  by  term,  it  does  not  follow  that  the  series 
can  be  obtained  by  differentiating  the  integrated  series  term  by 
term,  for  while  differentiation  makes  a  series  less  convergent, 
integration  makes  it  more  convergent.  In  other  words,  a  con- 
verging series  may  become  divergent  on  differentiation.  This 
raises  another  question — the  convergency  of  Fourier's  series.  In 
the  preceding  developments  it  has  been  assumed  : 

(i)  That  the  trigonometrical  series  is  uniformly  convergent. 

(ii)  That  the  series  is  really  equal  to  fix). 

Elaborate  investigations  have  been  made  to  find  if  these  as- 
sumptions can  be  justified.  The  result  has  been  to  prove  that  the 
above  developments  are  valid  in  every  case  when  the  function  is 
single-valued  and  finite  between  the  limits  +  -n-;  and  has  only 
a  finite  number  of  maximum  or  minimum  values  between  the 


§152. 


FOURIER'S  THEOREM. 


481 


limits  x  =  +  7t.  The  curve  y  =  f{x)  need  not  follow  the  same  law 
throughout  its  whole  length,  but  may  ba  made  up  of  several 
entirely  different  curves.  A  complete  representation  of  a  periodic 
function  for  all  values  of  x  would  provide  for  developing  each  term  as 
a  periodic  series,  each  of  which  would  itself  be  a  periodic  function, 
and  so  on. 

An  adequate  discussion  of  the  conditions  of  convergency  of  Fourier's  series 
must  be  omitted.  W.  E.  Byerly's-in-  Elementary  Treatise  on  Fourier's  Series, 
etc.,  is  one  of  the  best  practical  works  on  the  use  of  Fourier's  integrals  in 
mathematical  physics.  J.  Fourier's  pioneer  work  TlUorie  analytiqufr  de  la 
Chaleur,  Paris,  1822,  is  perhaps  as  modern  as  any  other  work  on  this  subject ; 
see  also  W.  Williams,  Phil.  Mag.  [5],  42,  125,  1896  ;  Lord  Kelvin's  Collected 
Papers ;  and  Riemann-Weber's  work  (Z.c),  etc. 


§  152.  Fourier's  Linear  Diffusion  Law. 

Let  AB  be  any  plane  section  in  a  metal  rod  of  unit  sectional 
area  (Fig.  163).  Let  this  section  at  any  instant  of  time  have  a 
uniform  temperature — equi-thermal  surface — and  let  the  tempera- 
ture on  the  left  side  of  the  plane  AB  be  higher  than  that  on  the 
right.  In  consequence,  heat  will  flow  from  the  hot  to  the  cold 
side,  in  the  direction  of  the  arrow,  across  the  surface  AB.  Fourier 
assumes,  (i)  The  direction  of  the  flow  is  perpendicular  to  the  section 
AB ;  (ii)  The  rate  of  flow  of  heat  across  any  given  section,  is  pro- 
portional to  the  difference  of  temperature  on  the  two  sides  of  the 
plate. 

Let  the  rate  of  flow  be  uniform,  and  let  6  denote  the  tempera- 
ture of  the  plane  AB.  The  rate  of  rise  of  temperature  at  any  point 
in  the  plane  AB,  is  dO/ds — 
the  so-called  "temperature 
gradient.  The  amount  of 
heat  which  flows,  per 
second,  from  the  hot  to  the 
cooler  end  of  the  rod,  is 
-  a .  dO/ds,  where  a  is  a 
constant  denoting  the  heat 
that     flows,     per     second, 


0 

Fig.  163. 

through  unit  area,  when  the  tempertaure  gradient  is  unity.  Con- 
sider now  the  value  of  -  a .  d$/ds  at  another  point  in  the  plane 
CD,  distant  Ss  from  AB ;  this  distance  is  to  be  taken  so  small, 
that  the  temperature   gradient  may  be  taken  as  constant.     The 

HH 


482  HIGHER  MATHEMATICS.  §  152. 

temperature  at  the  point  s  +  $s,  will  be  f  0  -  -t-Ss  j,  since  -  -j-  is 

the  rate  of  rise  of  temperature  along  the  bar,  and  this,  multiplied 
by  8s,  denotes  the  rise  of  temperature  as  heat  passes  from  the 
point  s  to  s  +  hs.  Hence  the  amount  of  heat  flowing  through  the 
small  section  ABCD  will  be 

d/        d0<s\  d2<9 

"  ads\°-Tshs)'aM  +  adi^  *  '  (34) 
will  denote  the  difference  between  the  amount  of  heat  which  flows 
in  at  one  face  and  out  at  the  other.  This  expression,  therefore, 
denotes  the  amount  of  heat  which  is  added  to  the  space  ABCD 
every  second.  If  a  denotes  the  thermal  capacity  of  unit  volums, 
the  thermal  capacity  of  the  portion  ABCD  is  (1  x  Ss)a-.  Hence 
the  rate  of  rise  of  temperature  per  unit  area  is  a(d6/dt)os.  There- 
fore, 

aaT^  "  ^  '         '       (35> 

Put  a/a-  =  k  ;  this  equation  may  then  be  written, 

1    dO  _  d20 

k'  dt~  M'  •  '  *  '  (36) 
where  k  is  the  diffusivity  of  the  substance.  Equation  (36)  re- 
presents Fourier's  law  of  linear  diffusion.  It  covers  all  possible 
cases  of  diffusion  where  the  substances  concerned  are  in  the  same 
condition  at  all  points  in  any  plane  parallel  to  a  given  plane.  It 
is  written  more  generally 

1    dV_dW 

K'dt~dx2 (37) 

If  we  had  studied  the  propagation  of  the  "disturbance"  in 
three  dimensions,  instead  of  the  simple  case  of  linear  propaga- 
tion— in  one  direction — equation  (37)  would  have  assumed  the 
form, 

1    dV  __  dW      d?V      <PV 

k  '  dt~  dx>+  dy2  +  dz2 '         '        '        (38) 

Lord  Kelvin  calls  V  the  quality  of  the  substance  at  the  time  t,  at  a 
distance  x  from  a  fixed  plane  of  reference.  The  differential  equa- 
tion (37),  therefore,  shows  that  the  rate  of  increase  of  quality  per 
unit  time,  is  equal  to  the  product  of  the  diffusivity  and  the  rate 
of  increase  of  dV/ds,  i.e.  quality  per  unit  of  space.  The  quality 
depends  on  the  subject  of  the  diffusion.      For  example,  it  may 


§  153.  FOURIER'S  THEOREM.  483 

denote  one  of  the  three  components  of  the  velocity  of  the  motion 
of  a  viscous  fluid,  the  density  or  strength  of  an  electric  current  per 
unit  area  perpendicular  to  the  direction  of  flow,  temperature,  the 
potential  at  any  point  in  an  isolated  conductor,  or  the  concentration 
of  a  given  solution.  Ohm's  law  is  but  a  special  case  of  Fourier's 
linear  diffusion  law.  Fick's  law  of  diffusion  is  another.  The  trans- 
mission of  telephonic  messages  through  a  cable,  and  indeed  any 
phenomenon  of  linear  propagation,  is  included  in  this  law  of 
Fourier. 

§  153.    Application  to  the  Diffusion  of  Salts  in  Solution. 

Fill  a  small  cylindrical  tube  of  unit  sectional  area  with  a  solution 
of  some  salt  (Fig.  164).  Let  the  tube  and  contents  be  submerged 
in  a  vessel  containing  a  great  quantity  of  water,  so  that  the  open 
end  of  the  cylindrical  vessel,  containing  the  salt  solution,  dips  just 
beneath  the  surface  of  the  water.  Salt  solution  passes  out  of  the 
diffusion  vessel  and  sinks  towards  the  bottom  of  the  larger  vessel. 
The  upper  brim  of  the  diffusion  vessel,  therefore,  is  assumed  to  be 
always  in  contact  with  pure  water.  Let  h  denote  the  height  of  the 
liquid  in  the  diffusion  tube,  reckoned  from  the  bottom  to  the  top. 
The  salt  diffuses  according  to  Fourier's  law, 

dV        d?V 

Tt  =  KWx*~      ....        (1) 

which  is  known  as  Fick's  law  of  diffusion  of  substances  in 
solution. 

/.  To  find  the  concentration,  V,  of  the  dissolved  substance  at 
different  levels,  x,  of  the  diffusion  vessel  after  the  elapse  of  any 
stated  interval  of  time,  t.  This  is  equivalent  to  finding  a  solution 
of  Fick's  equation,  which  will  satisfy  the  conditions  under  which 
the  experiment  is  conducted.  These  so-called  limiting  condi- 
tions ara :  (i)  when 
**~""-*  dV 

*  =  °>^  =  °;      •     •     •     (2) 

dV/dx  =  0  means  that  no  salt  goes  out  from,  and  no  salt  enters  the 
solution  at  this  point,     (ii)  when 

x  =  h,  7=0;  .         .         .         (3) 

and  (iii)  when 

*  =  0,  7  =  70.  .        .        .        (4) 

The  reader  must  be  quite  clear  about  this  before  going  any  further. 


484  HIGHER  MATHEMATICS.  §  153. 

What  do  V,  x  and  t  mean  ?     V0  evidently  represents  the  concen- 

,     tration  of  the  salt  solution  at  the  beginning 

of  the  experiment ;   V  is  the  concentration 

of   the   solution   expressed   in,    say,    gram 

—x=o    m0^ecu^es  °f  sa^  Per  ntre  °f  solution,  at  a 

distance  x  from  the  bottom  of   the  inner 

vessel  (Fig.  164)  at  the  time  t ;  at  the  top 

of   the    diffusion   vessel,    obviously   x  =  h, 

and  V  is  zero,  because  there  the  water  is 

Fig.  164.  pure  ;  the  first  condition  means  that  at  the 

bottom  of  the  diffusion  vessel,  the  concentration  may  be  assumed 

to  be  constant  during  the  experiment. 

First  deduce  particular  solutions.      Following  the  method  of 
page  462,  assume  tentatively  that 

y  =  eax  +  at       #  .  #  #  (5) 

is  a  solution  of  (1),  when  a  and  ft  are  constants.     Substitute  this  in 
(1),  and  we  get 

ft  =  KO?.  ....  (6) 

Hence,  if  (6)  is  true,  (5)  is  a  solution  of  (1)  whatever  be  the  value 
of  a.     Hence  it  is  true  when  a  =  t/x. 

'.'     Y  _   QiyJC  +  fit    —    QipX  +  Kcfit    _    QifjiX  -  KfjPt^ 

Consequently 

V  =  e  -  «f-2te^x ;  and  V  =  e  -  K^le  ~  i** 

are  both  solutions  of  (1).     Hence  the  sum  and  difference 

V  =  ie- ^(e1**  ±  e- l»x) 

are  also  solutions  of  (1) ;  and  from  Euler's  sine  and  cosine  series, 

page  285, 

V  =  ae  - K^1  cos  \xx ;  &ndV=be-  •**  sin  px,       .         (7) 

where  a  and  b  are  constants,  as  well  as 

V  =  (a  cos  fix  +  b  sin  fxx)e  -  «**.  \      .         .         (8) 

are  solutions  of  the  given  equation.     It  remains  to  fit  the  constants 

a  and  b  in  with  the  three  given  conditions. 

Condition  i,  when  x  =  0,  dV/dx  =  0.      Differentiate  (8)  with 

respect  to  x,  and  we  get 

dV 

-r-  =  (  -  fxa  sin  jxx  +  pb  cos  fix)e  -  ^K 

Now  when  x  =  0,  sinfix  vanishes,  and  when  x  =  0,  cos/xo;  =  1. 
Consequently  b  must  be  zero,  if  dV/dx  is  to  be  zero  when  x  —  0. 
Hence  (5),  i.e.  (8),  satisfies  the  first  condition. 


§  153.  FOURIER'S  THEOREM.  485 

Condition  ii,  when  x  =  h,  V  =  0.    In  order  that  (8)  may  satisfy 
the  second  condition,  we  must  have  cos  \xh  =  0,  when  x  =  h.     But 

cos  \tt  =  cos  |?r  =  . . .  =  cos \{%fl  -  l)?r  =  0, 
where  tt  =  180°  and  n  is  any  integer  from  1  to  oo.     Hence,  we 
must  have  fJi  =  \tt  ;  yfo  =  j*r  ;  . .  . ,  or 

JT^  %       _    37T  _    57r  (2ft    —    1)7T 

/X  =  2^;   ~M;   ~  2/T;  "•  =       2A        ; 
in  order  that  cos  fxh  may  vanish.      Substitute  these  values  of  fx 
successively  in  (8)  and  add  the  results  together ;  we  thus  obtain 

f  -  y  -  ® '"'cos  5  +  ty  "  ®"'Wjf  +  ...  to  inf.,      (9) 

which  satisfies  two  of  the  required  conditions,  namely  (1)  and  (2). 
Condition  iii,  when  t  =  0,  V  =  70,  we  must  evaluate  the  coeffi- 
cients ax,  a2, . . .  in  (9),  in  such  a  way  that  the  third  condition  may 
be  satisfied  by  the  particular  solution  (8),  or  rather  (9).  This  is 
done  by  allowing  for  the  initial  conditions,  when  t  =  0,  in  the 
usual  way.     When  t  =  0,  V  =  VQ.     Therefore,  from  (9), 

_  7TX  S7TX 

V0  =  ajcos  ^  +  a2cos-o/T  + (10) 

is  true  for  all  values  of  x  between  0  and  h.     Hence,  Q.n~--~7    1  ooo  — 
2F0f*      irx.  2V0Ch      Bttx.  4F0  '      ,„ 

ai  =  -r}0cosZhdx>  a>  =  -Fj0cos^^-"^  =  (2^:T>r-  (n^^ 

These  results  have  been  obtained  by  equating  each  term  of  (9)  to  ~rr 
zero,  and  integrating  between  the  limits  0  and  h.     Substituting 
these  values  of  a0,  av  . . .  in  (9),  we  get  a  solution  satisfying  the      — 
limiting  conditions  of  the  experiment.      If  desired,  we  can  write 
the  resulting  series  in  the  compact  form, 


2n-T  cos—^-ttx,       (12) 

where  the  summation  sign  between  the  limits  n  =  oo  and  n  =  0 
means  that  n  is  to  be  given  every  positive  integral  value  0,  1,  2,  3 
...  to  infinity,  and  all  the  results  added  together. 

If  we  reckon  h  from  the  top  of  the  diffusion  vessel  x  =  0  at  the 
top,  and, -at  the  bottom,  x  =  h,  hence  it  follows,  by  the  same 
method,  that 

We  could  have  introduced  a  fourth  condition  dV/dx  =  0,  when 


486  HIGHER  MATHEMATICS.  §  153. 

x  —  h,  but  it  would  lead  to  the  same  result,  as  we  shall  see  in  one 
of  the  subjoined  examples,  viz.,  Ex.  (1). 

mgd&  Examples. — (1)  T.  Graham's  diffusion  experiments  (Phil.  Trans.,  151, 188, 
/ '-',  1861).  A  cylindrical  vessel  152  mm.  high,  and  87  mm.  in  diameter,  contained 
0*7  litre  of  water.  Below  this  was  placed  0*1  litre  of  a  salt  solution.  The 
fluid  column  was  then  127  mm.  high.  After  the  elapse  of  a  certain  time, 
successive  portions  of  100  c.c,  or  £  of  the  total  volume  of  the  fluid,  were 
removed  and  the  quantity  of  salt  determined  in  each  layer.  Here  x  =  0 
at  the  bottom  of  the  vessel,  and  x  =  H  at  the  top ;  x  =  h  at  the  surface 
separating  the  solution  from  the  liquid  when  t  =  0.  The  vessel  has  unit 
irea.  The  limiting  conditions  are :  At  the  end  of  a  certain  time  t,  (i)  when 
v  =  0,  dV/dx  =  0  ;  and  (ii)  when  x  =  H,  dV/dx  =  0 ;  (iii)  when  t  =  0,  V=  V0 
between  x  =  0  and  x  =  h ;  (iv)  when  t  =  0,  V  =  0  between  x  =  h  and  x  =  H. 
To  adapt  these  results  to  Fourier's  solution  of  Fick's  equation,  first  show  that 
(6)  is  a  particular  integral  of  Fick's  equation.  Differentiate  (8)  with  respect 
to  x  and  show  that  for  the  first  condition  we  must  have  b  zero,  and  condition 
(i)  is  satisfied.  For  condition  (ii),  sin  /xH  must  be  zero  ;  but  sin  ?iir  is  zero ; 
hence  we  can  put 

fiH  =nir;  or,  fx  =  -gr, 

where  n  has  any  value  0,  1,  2,  3, . . .  Adding  up  all  the  particular  integrals,  we 
have 

-KX    -(1)**  2irX    -fi£W 

V  =  a0  +  ajcos  -H-e      ^u'       +  a2cos  -jnre      v>/i/      +  . . ., 

where  the  constants  a0,  a^,  a2, . . .  have  to  be  adapted  to  conditions '  (iii)  and 
(iv).     For  condition  (iii),  when  t  =  0,  V  =  V0,  consequently, 

irX  2irX 

Vo  =  a0  +  a^os  -g  +  OjjCos  -g   +  . . ., 

from  x  =  0,  to  x  =  h.  For  condition  (iv),  substitute  V  =  0,  from  x  =  h  to 
x  =  H.     Hence,  from  (4),  page  241, 

In  the  same  way  it  can  be  shown  that 

2  CB  ■        n*xn        2VQfh      mrxJnirx\       2V0    .    rnrh 
"»=Hj0  V«coa  -Wdx  =  1^)  Ocos  -Wd\-W)  =  15?  sm  ~H' 
Hence,  taking  all  these  conditions  into  account,  the  general  solution  appears 
in  the  form, 

which  is  a  standard  equation  for  this  kind  of  work.  In  Graham's  experi- 
ments, h  =  IH.  Hence  the  concentration  V  in  any  plane,  distant  x  units 
from  the  bottom  of  the  diffusion  vessel,  is  obtained  from  the  infinite  series : 

-      V0      2V0    *  =  °°1    .     nn         mrx    -  C^Y<t  /1Kt 

V=-^-  +  — ^2         -sin —.cos -^-e      \H)     .  .      (15) 

An  infinite  series  is  practically  useful  only  when  the   series  converges 
rapidly,  and  the  higher  terms  have  so  small  an  influence  on  the  result  that 


§  153.  FOURIER'S  THEOREM.  487 

all  but  the  first  terms  rftay  be  neglected.  This  is  often  effected  by  measuring 
the  concentration  at  different  levels  x,  so  related  to  If  that  costyiirx/H)  reduces 
to  unity  ;  also  by  making  t  very  great,  the  second  and  higher  terms  become 
vanishingly  small. 

(2)  H.  F.  Weber's  diffusion  experiments  (Wied.  Ann.,  7,  469,  536,  1879; 
or  Phil.  Mag.,  [5],  8,487,  523,  1879).  A  concentrated  solution  of  zinc  sulphate 
(0*25  to  0*35  grm.  per"  c.c.  of  solution)  was  placed  in  a  cylindrical  vessel  on 
the  bottom  of  which  was  fixed  a  round  smooth  amalgamated  zinc  disc  (about 
71  cm.  diam.).  A  more  dilute  solution  (0*15  to  O20  grm.  per  c.c.)  was  poured 
over  the  concentrated  solution,  and  another  amalgamated  zinc  plate  was 
placed  just  beneath  the  surface  of  the  upper  layer  of  liquid.  It  is  known  that 
if  Vly  V2  denote  the  respective  concentrations  of  the  lower  and  upper  layers 
of  liquid,  the  difference  of  potential  E,  due  to  these  differences  of  concentra- 
tions, is  given  by  the  expression 

E  =  A(V2  -  VJ{1  +  B(V2  +  V,)},  .  .  .  (16) 
where  A  and  B  are  known  constants,  B  being  very  small  in  comparison  with  A. 
This  difference  of  potential  or  electromotive  force,  can  be  employed  to  deter- 
mine the  difference  in  the  concentrations  of  the  two  solutions  about  the  zinc 
electrodes.  To  adapt  these  conditions  to  Fick's  equation,  let  hx  be  the  height 
of  the  lower,  h2  of  the  upper  solution,  therefore  \  +  h2  =  H.  The  limiting 
conditions  to  be  satisfied  for  all  values  of  t,  are  dV/dx  =  0,  when  x  =  0,  and 
dV\dx  =  0,  when  x  =  H.  The  initial  conditions  when  t  =  0,  are  V  =  V2t  for 
all  values  of  x  between  x  =  h  and  x  =  H.  From  this  proceed  exactly  as  ir 
Ex.  (3),  and  show  that 

a0  = ^ ,  an= - .-  sin  -^  , 

and  the  general  solution 

F=  FA+ry,  _  2JS-5J— x  .n  ^    ^  -  iff 

-a  ir  n  =i  n         i±  a  v    ' 

This  equation  only  applies  to  the  variable  concentrations  of  the  boundary 
layers  x  =  0  and  x  =  H.  It  is  necessary  to  adapt  it  to  equation  (16).  Let 
the  layers  x  =  0  and  x  =  H,  have  the  variable  concentrations  V  and  V" 
respectively. 

FA  +  V2\      2(V2  -  V,)  (  ._  *V  -IB*  .  1  _:     2^V  "  S3*' 

V      =  XT  ~ 


-V,)(.     rh,    -  m*t  ,   1    .     2irfc,    -  TR't  } 

r,+  r  =  2IA^.MI^)^sin^e-W'  +  l  et0.  ...}. 

In  actual  work,  H  was  made  very  small.  After  the  lapse  of  one  day  (t  =  1), 
the  terms.  £sin  iirhj/H,  etc.,  and  isin  birhJH,  etc.,  were  less  than  -&$. 
Hence  all  terms  beyond  these  are  outside  the  range  of  experiment,  and  may, 
therefore,  be  neglected.  Now  h  was  made  as  nearly  as  possible  equal  to  %H, 
in  order  that  the  term  $sin  SirhJH,  etc.,  might  vanish.     Hence, 


488  HIGHER  MATHEMATICS.  §  153. 

^ r -  ^^  -  ffiziJ  sin  £.,  -  4# 

Now  substitute  these  values  of  V"-V  and  V'+V  in  (16),  observing  that 
t>  T*  T7!,  feu  ^2>  sin  i^^  sin  1^  and  H,  are  all  constants,  and  that  V%  and  F2 
become  respectively  V  and  F'  when  t  =  0.  The  difference  of  potential  E, 
between  the  two  electrodes,  due  to  the  difference  of  concentration  between 
the  two  boundary  layers  V  and  V"  is 

E=A{V'-V'){l  +  B(V"+F')}=Ale       #2  +  Bxe       m  ,     .        (18) 

where  A1  and  Bx  are  constant.  Since  I?  is  very  small  in  comparison  with  A, 
the  expression  reduces  to 

E  =  ALe       m, (19) 

in  a  very  short  time.  This  equation  was  used  by  Weber  for  testing  the 
accuracy  of  Fick's  law.  The  values  of  the  constant  7r'2/c/IT2,  after  the  elapse 
of  4,  5,  6,  7,  8,  9,  10  days  were  respectively  0-2032,  0-2066,  0-2045,  0'2027» 
0-2027,  0-2049,  0-2049.  A  very  satisfactory  result.  See  also  W.  Seitz,  In- 
augural-Dissertation, Leipzig,  1897. 

II.  To  find  the  quantity  of  salt,  Q,  which  diffuses  through  any 
horizontal  section  in  a  given  time,  t.     Differentiate  (7)  with  respect 

to  x,  multiply  the  result  through  with  xdt,  so  as  to  obtain  -  K~=-dt. 

ax 

If  x  represents  the  height  of  any  given  horizontal  section,  then 
-  xq-j-dt  will  represent  the  quantity  of  salt  which  passes  through 

this  horizontal  plane  in  the  time  dt ;  q  represents  the  area  of  that 
section.     Let  the  vessel  have  unit  sectional  area,  then  q  =  1. 

Integrate  between  the  limits  0  and  t.  The  result  represents  the 
quantity  of  salt  which  passes  through  any  horizontal  plane,  x,  of 
the  diffusion  vessel  in  the  time  t,  or, 

III.  To  find  the  quantity  of  salt,  Qv  which  passes  out  of  the 
diffusion  vessel  in  any  given  time,  t.  Substitute  h  =  x  in  (20). 
The  sine  of  each  of  the  angles  \ir,  |tt,  . . .,  \{2n  +  1)  is  equal  to 
unity.     Therefore, ' 


§  153.  FOURIER'S  THEOREM.  489 

IV.  To  find  the  value  of  k,  the  coefficient  of  diffusion.  Since 
the  members  of  series  (21)  converge  very  rapidly,  we  may  neglect 
the  higher  terms  of  the  series.  Arrange  the  experiment  so  that 
measurements  are  made  when  x  =  h,  ^h,  \h, . . .,  in  this  way 
sin  7rx/2h,  ...  in  (20)  become  equal  to  unity.  We  thus  get  a  series 
resembling  (21).  Substitute  for  the  coefficient  and  we  obtain,  by 
a  suitable  transposition  of  terms, 

Q*     U       -aft        w,    (Q*      A 

270    2h(h       ttx    n/irx\      4Fn 

•••  —  l^fe"1}  (22) 

There  are  several  other  ways  of  evaluating  k  besides  this.     See 
page  198,  for  instance. 

V.  To  find  the  quantity  of  salt,  Q2,  which  remains  in  the 
diffusion  vessel  after  the  elapse  of  a  given  time,  t.  The  quantity 
of  salt  in  the  solution  at  the  beginning  of  the  experiment  may  be 
represented  by  the  symbol  Q0.  Q0  may  be  determined  by  putting 
t  =  0  in  (20)  and  eliminating  aimrx^h, ...  as  indicated  in  IV. 
2hf  1  \ 

Qq  =  ~^\ai  ~  3aa  +  •  •  -) » and  Q2  =  Qo  -  Ql  '» 

,Q^%e-^-\a.yO^  +  ..).  (23) 

Example. — A  solution  of  salt,  having  a  concentration  V0,  is  poured  into 
a  tube  up  to  two-thirds  of  its  height,  the  rest  of  the  tube  is  filled  with  pure 
solvent.     Find  Q2.     From  (9), 


-270/1 


2Vn  [lh       nirx  4yo        nv 

cos  -^dx  =  —  sin  T, 


where  n  =  1,  3,  5, . . .  and  h  denotes  the  height  of  the  tube.     Hence 
a,  =  —  sin  60°  = °  ;  a2  ^  0 ;  a3  =  -  - — = K 


*2 


From  (23),  we  have 

g2 ,  t^(«-(£)'«.  -  y(%Y"  -  £."<#*  i/C^V  +  •••). 

VI.  If  the  diffusion  vessel  is  divided  into  m  layers,  to  find  the 
quantity  of  salt,  Qr,  between  the  (r  -  l)th,  and  the  rth  layer.  The 
quantity  of  salt  in  the  rth  layer  dx  thick  is  obviously  dQr ;  and  since 
V  is  the  concentration  of  the  sal'  in  the  plane  x  units  distant  from 


490 


HIGHER  MATHEMATICS. 


§153. 


the  bottom  of  the  diffusion  vessel, 


dQr  =  Vdx 


Vdx. 


(r—  \)H 


(24) 


The  value  of  V  given  in  (10)  or  (11)  is  substituted  in  (24)  and  the 
integration  performed  in  the  usual  way. 

(1)  Returning  to  Graham's  experiments,  Ex.  (1),  page  486, 


Examples 
show  that 


«'=f    ft 

j  (r-m\  o 


0    .     ritr         flirX 
-\ sin  -s-  cos  -^fti 


(¥)*-) 


dx 


using  both  (24)  and  (15),  and  neglecting  the  summation  symbol  pro  tern. 
integration,  therefore 

n*(r-l)\     -ftrt'l 


On 


&-  m  +  is?  sm  t lsm  ~m~  - sm 


.-.£,-    8 
from  (39),  page  612. 


-  Vw    ™2*-2 


rtir   .     7i7r        nir(2r-l) 
sin-^-  sm  ~ —  cos 


(S) 


8   Ui"  2m  "^        2m 
Restoring  the  neglected  summation  symbol,  and  re- 
membering that  in  Graham's  experiments  m  =  8,  we  have 

rX      82    »  =  «  i      .    nir     .    nw         (2r-l)mr  -  Cwj***y 
Q+^n  =  1  ~2. sin  ^. sin  B. cos        lg 


Gr  = 


)■ 


(25) 


i.i  w2*Di".  8 

where  %V0H  multiplied  by  the  cross  section  of  the  vessel  (here  supposed 
unity)  denotes  the  total  quantity  of  salt  present  in  the  diffusion  vessel.  Put 
QQ  bs  \VqH.  Unfortunately,  a  large  number  of  Graham's  experiments  are  not 
adapted  for  numerical  discussion,  because  the  shape  of  his  diffusion  vessels, 
even  if  known,  would  give  very  awkward  equations.  A  simple  modification 
in  experimental  details  will  often  save  an  enormous  amount  of  labour  in  the 
mathematical  work.  The  value  of  Qr  depends  upon  the  value  of  cos  ^nir(2r  - 1). 
If  the  diffusion  vessel  is  divided  into  8  equal  parts — layers — r  has  the  values 
1,  2,  3, . . .  8.  Now  set  up  a  table  of  values  of  cos  ^  nir(2r  - 1)  for  values  of 
r  from  1  to  8  ;  and  for  values  of  n  =  1,  2,  3,  4, . . .  We  get 


rth  layer 

n  =  l 

n  =  2 

w  =  3 

n  =  4 

r  =  1 

+  cos  TV 

+  cos  i?r 

+  cos  tV 

+  COS  £ir 

r  =  2 

+  cos  tVt 

+  cos  -§*■ 

-  COS  ^n 

-     COS  ^7T 

r  =  3 

+    COS  ^TT 

—    COS  f  7T 

-  cos  tVt 

-   COS^7T 

r  =  4 

+  cos  tvt 

—   COS  |7T 

-    COS^TT 

+  COS  \ir 

r  =  5 

—    COS  ^7T 

-    COS  ^7T 

+  cos  TVr 

+  COS  \lT 

r  =  6 

-    COS^TT 

—  COS  fir 

+   COS  ^TT 

-  COS  \ir 

r  =  7 

—  COS  ^ir 

4-  COS  f  7T 

+    COS  ^TT 

-    COS  ^7T 

r  =  8 

—    COS  ^7T 

+   COS  ^7T 

-   COS  f\7r 

+   COS^7T 

By  calculating  up  corresponding  values  of  Qv  Q4,  Q5,  Q8  from  (25),  and  taking 
their  sum  it  will  be  found  that  the  first  three  terms  of  the  trigonometrical 


§  153.  FOURIER'S  THEOREM.  491 

series  cancel  out,  and  that  the  succeeding  terms  are  negligibly  small.  Ac- 
cordingly, we  get 

Qi+Qa+Q*+Q8=IQo>     •      •'  -       •       (26) 

a  result  in  agreement  with  J.  Stefan's  experiments  (Wien  Akad.  Ber.,  79,  ii. 
161,  1879),  on  the  diffusion  of  sodium  chloride,  and  other  salts  in  aqueous 
solution.  When  t  =  14  days,  series  (23)  is  so  rapidly  convergent  that  all  but 
the  first  two  terms  may  be  neglected.     Consequently, 

„  „  „    /l  64      .        IT      .         IT  7T  37T       -    (^YKt\ 

Qi+  Q2=  Qo\J  +  ^  sin  g  sm  ^  cos  ^  cos  j^e       ^a/      J  (27) 

remains.  Qv  Q2,  Q0,  H,  t,  can  all  be  measured,  e  and  -n  are  known  constants, 
hence  k  can  be  readily  computed. 

(2)  A  gas,  A,  obeying  Dalton's  law  of  partial  pressures,  diffuses  into  an- 
other gas,  B,  show  that  the  partial  pressure  p  of  the  gas  A,  at  a  distance  xt  in 
the  time 

^■-M  /Oft 

[n  Loschmidt's  diffusion  experiments  (Wien  Akad.  Ber.,  61,  367,  1870;  62, 
468,  1870)  two  cylindrical  tubes  were  arranged  vertically,  so  that  communica- 
tion could  be  established  between  them  by  sliding  a  metal  plate.  Each  tube 
was  48-75  cm.  high  and  2-6  cm.  in  diameter  and  closed  at  one  end.  The  two 
tubes  were  then  filled  with  different  gases  and  placed  in  communication  for  a 
certain  time  t.  The  mixture  in  each  tube  was  then  analyzed.  Let  the  total 
length  L  of  the  tubes  joined  end  to  end  be  97*5  cm.  It  is  required  to  solve 
equation  (28)  so  that  when  t  =  0,  p  =  p0,  from  x  =  0  to  x  =  $L  =  I;  p  =  0, 
from  x  =  $L  to  x  =  L  ;  dpjdx  =  0,  when  x  =  0,  and  x  =  L,  for  all  values 
of  U    Note  pQ  denotes  the  original  pressure  of  the  gas.     Hence  show  that 

Pox2ftvfl  =  Bl    •     *™         nirX    ~  (t)  *  (9Q\ 

»=  77+^-^2,        -  sin  -o-  cos  -=-  e      xw      •  •        \z^) 

r       2        it      n  =  i  n  2  L 

The  quantity  of  gas  Qx  and  Q2  contained  in  the  upper  and  lower  tubes,  after 
the  elapse  of  the  time  t,  is,  respectively, 

Q1=q.]pdx\Q2=Cijipdx,    ....        (30) 

where  I  =  £L,  and  q  is  the  sectional  area  of  the  tube.     Hence  show  that 

from  which  the  constant  k  can  be  determined.    If  the  time  is  sufficiently  long, 

Jr-V©'- <81> 

where  D  and  8  respectively  denote  the  sum  and  difference  of  the  quantities 
of  gas  contained  in  the  two  vessels.  Loschmidt  measured  D,  S,  t,  and  a  and 
found  that  the  agreement  between  observed  and  calculated  results  was  very 
close.  O.  A.  von  Obermayer's  experiments  (Wien  Akad.  Ber.,  83,  147,  749. 
1882 ;  87, 1,  1883)  are  also  in  harmony  with  these  results. 


,492  HIGHER  MATHEMATICS.  §  153. 

VII.  To  find  the  concentration  of  the  dissolved  substance  in 
different  parts  of  the  diffusion  vessel  when  the  stationary  state  is 
reached.  After  the  elapse  of  a  sufficient  length  of  time,  a  state 
of  equilibrium  is  reached  when  the  concentration  of  the  substance 
in  two  parts  of  the  vessel  is  maintained  constant.  This  occurs  if 
the  outer  vessel  (Fig.  164)  is  very  large,  and  the  liquid  at  the 
bottom  of  the  inner  vessel  is  kept  saturated  by  immediate  contact  £  \\ 
with  solid  salt.     In  this  case,  I  Of    -,.   * 

Integrate  the  latter,  and  we  get 

V  =  ax  +  b,       .         .        .         .       (32) 

where  a  and  b  are  constants  to  be  determined  from  the  experi- 
mental data,  as  described  in  §  108.  (32)  means  that  the  concen- 
tration diminishes  from '  below  upwards  as  the  ordinates  of  a 
straight  line,  in  agreement  with  Fick's  experiments. 

Exampes. — (1)  Adapt  (38),  page  482,  to  the  diffusion  of  a  salt  in  a  funnel- 
shaped  vessel.  For  a  conical  vessel,  q  =  -imfix1,  where  the  apex  of  the  cone 
is  at  the  origin  of  the  coordinate  axes,  m  is  the  tangent  of  half  the  angle 
included  between  the  two  slant  sides  of  the  vessel.  Fick  has  made  a  series 
of  crude  experiments  on  the  steady  state  in  a  conical  vessel  with  a  circular 
base  (funnel-shaped),  and  the  results  were  approximately  in  harmony  with 
the  equations 

The  apex  of  the  cone  was  in  contact  with  the  reservoir  of  salt,  hence  when 
x  =  0,  V  =  V0,  and  when  x  =  h,  the  height  of  the  cone,  V  =  0.  This  enables 
C1  and  C2  to  be  evaluated.  A.  Fick,  Fogg.  Ann.,  94,  59,  1855  ;  or  Phil.  Mag., 
[4],  10,  30,  1855. 

(2)  An  infinitely  large  piece  of  pitchblende  has  two  plane  faces  so  arranged 
that  the  cc-axis  is  perpendicular  to  the  faces.  Owing  to  the  generation  of  heat 
by  the  internal  changes  there  is  a  Continuous  outflow  of  heat  through  the 
faces  of  the  plate.  In  the  steady  state,  the  outflow  of  heat,  -  k(d26jds^)Ss  per 
sq.  cm.,  is  equal  to  the  rate  of  generation  of  heat  per  sq.  cm.,  say  to  q8s. 
Hence  show  d?6jds*=  -  qjk.  If  the  slab  be  100  cm.  thick  and  the  faces  be  kept 
at  0°C. :  and  if  q  be  120o000th  units,  and  k  =  0-005  (R.  J.  Strutt,  Nature,  68, 6y 
1903)  show  that  the  temperature  in  the  middle  of  the  slab  will  be  £°C.  hotter 
than  the  faces.  Hint.  First  integrate  the  above  equation.  The  limiting  con- 
ditions are-  6  =  0,  when  5  =  0,  and  when  s  =  100. 

as  1 

.«.  o=  -   wr(s  -  100) ;   and  6  =  g,, approximately, 

at  the  middle  of  the  slab  where  s  =  50. 


§  154.  FOURIER'S  THEOREM.  493 

§  15$.  Application  to  Problems  on  the  Conduction  of  Heat. 

The  reader  knows  that  ordinary  and  partial  differential  equa- 
tions differ  in  this  respect:  W.hil p.  ordinary- differential . equations 
have  only  a  finite  number  of  independent  particular  integrals,  partial 
differential  equations  have  an  infinite  number  of  such  integrals. 
And  in  practical  work  we  have  to  pick  out  one  particular  integral, 
to  satisfy  the  conditions  under  which  any  given  experiment  is  per- 
formed.    Suppose  that  a  value  of  V  is  required  in  the  equation 

'  7)3?  +  ^p"=  °'  '•  '  *  *  t1) 
such  that  when  y  =  oo,  V  =  0 ;  and  when  y  =  0,  V  =  f(x).  As 
on  page  484,  first  assume  that 

V  =  aw*?*, 
is  a  solution,  when  a  and  /?  are  constant.     Substitute  in  (1)  and 
divide  by  eay  +  Px,  and 

a2  +  (P  =  0 

remains.  If  this  condition  holds,  the  assumed  value  of  V  is  a 
solution  of  (1).  Hence  V  =  eav±-tax,  are  also  solutions  of  (1), 
therefore  also  eayeiax  and  e^e  ~ iax  are  solutions.  Add  and  divide 
by2i,  or  subtract  and  divide  by  2;  from  (13)  and  (15),  page  286, 
it  follows  that 

V  =  eay  cos  ax ;  and  V  =  e«y  sin  cur,  .  .  (2) 
must  also  be  solutions  of  (1).  Multiply  the  first  of  equations  (2) 
by  cos  a\,  and  the  second  by  sin  aA.  The  results  still  satisfy  (1). 
Add,  and  from  (24),  page  612, 

e~°-y  cos  a(A.  -  x) 
must  also  satisfy  (1).     Multiply  by  f(\)d\,  and  the  result  is  still  a 
solution  of  (1) 

e~ayf(k) .  cos  a(A  -  x)dk. 

Multiply  by  \\ir  and  find  the  limits  when  a  has  different  values 
between  0  and  oo  .  Hence,  from  (30)  and  (32),  page  479,  we  have 
the  particular  solution  satisfying  the  required  conditions 


-     da       e*aVWcosa(A  -  x)d\. 

•J  0  J  -  00 


(3) 


Examples. — (1)  A  large  iron  plate  tr  cm.  thick  and  at  a  uniform  tempera- 
ture of  100°  is  suddenly  placed  in  a  bath  at  zero  temperature  for  10  seconds. 
Required  the  temperature  of  the  middle  of  the  plate  at  the  end  of  10  seconds, 
(supposing  that  the  difiusivity  k  of  the  plate  is  0-2  C.O.S.  units,  and  that  the 


494  HIGHER  MATHEMATICS.  §  154. 

surfaces  of  the  plate  are  kept  at  zero  temperature  the  whole  time.  If  heat 
flows  perpendicularly  to  the  two  faces  of  the  plate,  any  plane  parallel  to  these 
faces  will  have  the  same  temperature.     Thus  the  temperature  depends  on 

dt  ~  Kdx» w 

as  in  (1)  page  483.  The  conditions  to  be  satisfied  by  the  solution  are  that 
6  =  100°,  when  t  =  0 ;  0  =  0,  when  x  =  0  ;  0  =  0,  when  x  =  7r.  First,  to  get 
particular  solutions.  Assume  6  =  e<vc+pt  is  a  solution,  as  on  page  484.  Hence 
show  that 

e  =  e~  «& cos /*», (5) 

d  =  e  -  *m2*  sin  fix, (6) 

are  solutions  of  (4).  By  assigning  particular  values  to  jx,  we  shall  get  par- 
ticular solutions  of  (4).  Second,  to  combine  these  particular  solutions  so  as  to 
get  a  solution  of  (4)  to  satisfy  the  three  conditions,  we  must  observe  that  (6) 
is  zero  when  x  =  0,  for  all  values  of  /*,  and  that  (6)  is  also  zero  when  x  =  ir  if 
i  is  an  integral  number.  If,  therefore,  we  put  6  equal  to  a  sum  of  terms  of 
the  form  Ae  -  *i&t  sin  nx,  say, 

6  =  a^~  Kt  sinaj  +  a%e  ~  4l<t  sin  2x  +  a3e  "  9lct  sin  3a;  4- (7) 

to  n  terms,  this  solution  will  satisfy  the  second  and  third  of  the  above  con- 
ditions, because  sin  ir  =  0  =  sin  0.     When  t  =  0,  (7)  reduces  to 

d  =  ax  sin  x  +  a2  sin  2x  +  a3  sin  Sx  +  .  .  .  .        .        (8) 

But  for  all  values  of  x  between  0  and  ?r,  we  see  from  Ex.  (1),  page  475  that  if 
6  =  100,  then,  from  43,  an  =  0  if  n  is  even,  and  400/w.7r,  if  n  is  odd. 

400/  .  1-    ■  1   .  •  \ 

6  =  —   sin  x  +  -q  sm  Sx  +  -=  sin  5x  +  . . .  ).      .        .        (9) 

We  must  substitute  the  coefficients  of  this  series  for  av  a2,  a3,  .  .  .  in  (7),  to  get 
a  solution  satisfying  all  the  required  conditions.  Note  a2,  a4,  .  .  .  in  (9)  are 
zero.     We  thus  obtain  the  required  solution 

400/  1  \ 

0  =  —  (  e  ~  Kt  sin  x  +  ~^e  ~  9>ct  sin  3s  +  . . .  J.      .        .        (10) 

To  introduce  the  numerical  data.  When  x  =  £tt,  t  =  10,  k  =  0-2.  Hence 
use  a  table  of  logarithms.  The  result  is  accurate  to  the  tenth  of  a  degree  if 
all  terms  of  the  series  other  than  the  first  be  suppressed.    Hence  use 

400     „       7T      400      0 
6=  — <?-2sin-r= — e~2 

IT  A  IT 

for  the  numerical  calculation.  Note  sin  %tt  =  1.  Ansr.  17  "2°  0.  In  the 
preceding  experiment,  if  the  plate  is  c  centimetres  instead  of  v  centimetres 
thick,  use  the  development 

400/  .    irx       i    .     3tto;  \ 

*=—  (slnT  +  3sm-B-  +  ---> 
from  x  =  0  to  x  =  c,  in  place  of  (9). 

(2)  An  infinitely  large  solid  with  one  plane  face  has  a  uniform  tempera- 
ture f(x).  If  the  plane  face  is  kept  at  zero  temperature,  what  is  the  tempera- 
ture of  a  point  in  the  solid  x  feet  from  the  plane  face  at  the  end  of  t  years  ? 
Let  the  origin  of  the  coordinate  axis  be  in  the  plane  face.     We  have  to 


§154.  FOUKIER'S  THEOREM  495 

solve  equation  (4)  subject  to  the  conditions  0  =  0,  when  x  =  0;  6  =  f(x),  when 
t  =  0.  Proceed  according  to  the  above  methods  for  (5),  (6),  and  (3).  We 
thus  obtain 

e  =  -  [  da[+<*e-  Ka2tf{\)  cos  o(\  -x).d\;        .        .        (11) 
""Jo      J  -  » 
since  positive  values  of  x  are  wanted  we  can  write,  from  (33),  page  480, 

6  =  -  I    da  I    e-  Ka2tf{K)  sin  ax .  sin  a\  .d\.         .        .        (12) 

Hence  from  (28),  page  612,  the  required  solution  is 

0  =  -  /     /(A)dA  /     e  -  Ka2'{cos  a(\  -  X)  -  COS  a(A  +  x))da. 

■•■°=^tUm\e~  M  -*~  4"  }*•  •   (13) 

This  last  integration  needs  amplification.    To  illustrate  the  method,  let 

/•OO 

m  =  /    0  -  a2*2cos  bx .  efo. 
Jo 

Laplace  (1810)  first  evaluated  the  integral  on  the  right  by  the  following 
*.  ingenious  device  which  has  been  termed  integration  by  differentiation.     Dif- 
ferentiate the  given  equation  and 

gr—~\    xe~ a2*2 sin  &e  •  dx, 

provided  6  is  independent  of  x.  Now  integrate  the  right  member  by  parts  in 
the  usual  way,  page  205, 

du  6  du  b   ,, 

db  2a2  u  2a? 


Integrate,  and 

To  evaluate  G,  put  6  =  0,  whence 


6s  _  OL 

log  u  =  -  j-2  +  G  ;  or  u  =  Ce     4a2. 


as  in  (10),  page  344.     Therefore 

e  ~  a2x2  cos  bx.dx  =  ^e     *<#• 

Returning,  after  this  digression,  to  the  original  problem.     Let  us  change 
the  variables  by  substituting  in  (13), 

0  =  2-7-  ;  .".  \=x  +  2pKs/F;  .-.  d\=2s/Kt.d0.       .        .       (14) 

What  will  be  the  effect  of  substituting  these  values  of  A,  and  d\  upon  the 

limits  of  integration  ?    Hitherto  this  has  not  been  taken  into  consideration 

because  we  have  dealt  either  with  indefinite  integrals,  page  201,  or  with 

definite  integrals  with  constant  limits,  page  240.     Here  we  see  at  once  that  if 

x 
A=0,  0=  -— j=;  and  if  A=  oo,  $=  +  oo.  •        •      (15) 

Consequently,  expression  (12)  assumes  the  form 

e  =  -/={  /I     *  A  ~  0V(2jS  sJ7t  +  x)dfi  -  r    x  e  -  02/(20  s!7t  -  x)de\ .  (16) 


496  HIGHER  MATHEMATICS.  §  154. 

If  the  initial  temperature  be  constant,  say  f(x)  —  6Q,  then,  from  (4),  page  241, 

the  required  solution  assumes  the  form 

x 

6  =  h,{  r     x   e  -  Pdfl  -  P     x   e  ~  WJ  =3  f V«7e  -  Pdfi.  (17) 

For  numerical  computation  it  is  necessary  to  expand  the  last  integral  in  series 
as  described  on  page  341.      Therefore 


20o  (    x  x* 


}.     .        .        .       (18) 


n/;IW«7  3.(2^)3 
If  100  million  years  ago  the  earth  was  a  molten  mass  at  7,000°  F.,  and,  ever 
since,  the  surface  had  been  kept  at  a  constant  temperature  0°  F.,  what  would 
be  the  temperature  one  mile  below  the  surface  at  the  present  time,  taking 
Lord  Kelvin's  value  K  =  400  ?  Ansr.  104°  F.  (nearly).  Hints.  0O  =  7,000°; 
x  =  5,280  ft. ;  t  =  100,000,000  years. 

,      _  2  x  7,000/  5280  \  = 

s/^UlSK*  x  20  x  10,000;  U  ' 
Lord  Kelvin,  "  On  the  Secular  Cooling  of  the  Earth,"  (W.  Thomson  and  P.  G. 
Tait's  Treatise  on  Natural  Philosophy,  1, 711, 1867),  has  compared  the  observed 
values  of  the  underground  temperature  increments,  ddjdx,  with  those  deduced 
by  assigning  the  most  probable  values  to  the  terms  in  the  above  expressions. 
The  close  agreement— Calculated :  1°  increment  per  A  ft.  descent.  Observed : 
1°  increment  per  -^  ft.  descent — led  him  to  the  belief^  that  the  data  are  nearly 
correct.  He  extended  the  calculation  in  an  obvious  way  and  concluded :  "  I 
think  we  may  with  much  probability  say  that  the  consolidation  cannot  have 
taken  place  less  than  20,000,000  years  ago,  or  we  should  have  more  under- 
ground heat  than  we  really  have,  nor  more  than  400,000,000  years  ago,  or  we 
should  not  have  so  much  as  the  least  observed  underground  increment  of 
temperature".  Vide  O.  Heaviside's  Electromagnetic  TJieory,  2,  12,  London, 
1899.  The  phenomena  associated  with  "  radio-activity  "  have  led  us  to  modify 
the  Original  assumption  as  to  the  nature  of  the  cooling  process.  See  Ex.  (2), 
page  492. 

(3)  Solve  (4)  for  two  very  long  bars  placed  end  to  end  in  perfect  contact, 
one  bar  at  1°C.  and  the  other  at  0°C.  under  the  (imaginary)  condition  that  the 
two  bars  neither  give  nor  receive  heat  from  the  surrounding  air.  Let  the 
origin  of  the  axes  be  at  the  junction  of  the  two  bars,  and  let  the  bars  lie  along 
the  #-axis.  The  limiting  conditions  are  :  When  t  =  0,  0  =  0,  when  x  is  less 
than  zero,  and  0  =  f(x)  =  1,  when  x  is  greater  than  zero.  It  is  required  to 
find  the  relation  between  0,  x,  and  t.  Start  from  Fourier's  integral  (32),  page 
479,  and  proceed  to  find  the  condition  that  u  may  be  f(x)  when  t  =  0  by  the 
method  employed  for  (3)  above.  Change  the  order  of  integration,  and  we 
obtain 

0  =  ^[_™f(\.)d\f  6  -  Kait  cos  o(\  -  x)da.  .        .         (19) 

Integrate  by  Laplace's  method  of  differentation,  and 

1  [+°°  (  sj*       fr>  *P\         1      /•+«  fr  -  *y* 

0  =  *J-J^Hz37te~    iKt  J  =  *^tJ-~me~    ^  dA-  (20> 


§  154.  FOURIER'S  THEOREM.  497 

But  if  0  =  0,  when  t  =  0,  from  x  =  -  oo   to  a;  =  0 ;  0  =  1,  when  t  =  0,  from 
x  =  0,  to  x  =  +  co  ;  then,  since  /(a;)  =  1, 

1       r    _  (A  -  *)2  , 
6=WSJoe       ^"^         •        *        •         <21> 
Make  the  substitutions  (14)  and  (15)  above,  and 

X 

o  =  3;/ tlu  -  *#  =  ^(/P«  "  "*»  +  /,"'  -  "VA       (22) 

by  (3),  page  241.     Then,  on  integration,  pages  341  and  463,  the  required  solu- 
tion is 

°  =  2  +  7^2^  "  Q\273)    +  672l(2^)5  -•••}•    '         (23) 

The  process  of  diffusion  of  heat  here  exemplified  is  quite  analogous  to  the 
diffusion  of  a  salt  from  a  solution  placed  in  contact  with  the  pure  solvent. 
If  k  be  determined,  it  is  possible  by  means  of  (23),  to  compute  the  weight  of 
salt  (0)  in  unit  volume  of  solution  at  any  time  (t),  and  at  any  distance  (x)  from 
the  junction  of  the  two  fluids.  When  a  set  of  values  of  0,  xt  and  t  are  known, 
k  can  be  computed  from  (23).  See  J.  G.  Graham,  Zeit.  phys.  Chem.,  50,  257 
1904,  for  an  example. 


n 


CHAPTEE  IX. 
PROBABILITY  AND  THE  THEORY  OF  ERRORS. 

"Perfect  knowledge  alone  can  give  certainty,  and  in  Nature  perfect 
knowledge  would  be  infinite  knowledge,  which  is  clearly  beyond 
our  capacities.  We  have,  therefore,  to  content  ourselves  with 
partial  knowledge — knowledge  mingled  with  ignorance,  producing 
doubt." — W.  Stanley  Jevons. 

"  Lorsqu'il  n'est  pas  en  notre  pouvoir  de  discerner  les  plus  vraies 
opinions,  nous  devons  suivre  les  plus  probables."  1 — Rene 
Descartes. 

§  155.  Probability. 

Neaely  every  inference  we  make  with  respect  to  any  future  event 
is  more  or  less  doubtful.  If  the  circumstances  are  favourable,  a 
forecast  may  be  made  with  a  greater  degree  of  confidence  than 
if  the  conditions  are  not  so  disposed.  A  prediction  made  in  ignor- 
ance of  the  determining  conditions  is  obviously  less  trustworthy 
than  one  based  upon  a  more  extensive  knowledge.  If  a  sports- 
man missed  his  bird  more  frequently  than  he  hit.  we  could  safely 
infer  that  in  any  future  shot  he  would  be  more  likely  to  miss  than 
to  hit.  In  the  absence  of  any  conventional  standard  of  compari- 
son, we  could  convey  no  idea  of  the  degree  of  the  correctness  of 
our  judgment.  The  theory  of  probability  seeks  to  determine  the 
amount  of  reason  which  we  may  have  to  expect  any  event  when 
we  have  not  sufficient  data  to  determine  with  certainty  whether  it 
will  occur  or  not  and  when  the  data  will  admit  of  the  application 
of  mathematical  methods. 

A  great  many  practical  people  imagine  that  the  "doctrine  of 
probability  "  is  too  conjectural  and  indeterminate  to  be  worthy  of 
serious  study.     Liagre 2  very  rightly  believes  that  this  is  due  to 

1  Translated  :  "  When  it  is  not  in  our  power  to  determine  what  is  true,  we  ought 
to  act  according  to  what  is  most  probable  ". 

2  J.  B.  J.  Liagre's  Calcul  des  Probabilites,  Bruxelles,  1879. 

49S 


§  155.    PROBABILITY  AND  THE  THEORY  OP  ERRORS.      499 

the  connotation  of  the  word — probability.  The  term  is  so  vague 
that  it  has  undermined,  so  to  speak,  that  confidence  which  we 
usually  repose  in  the  deductions  of  mathematics.  So  great,  indeed, 
has  been  the  dominion  of  this  word  over  the  mind  that  all  applica- 
tions of  this  branch  of  mathematics  are  thought  to  be  affected  with 
the  unpardonable  sin — want  of  reality.  Change  the  title  and  the 
"theory"  would  not  take  long  to  cast  off  its  conjectural  character, 
and  to  take  rank  among  the  most  interesting  and  useful  applica- 
tions of  mathematics. 

Laplace  remarks  at  the  close  of  his  Essai  philosophique  sur  les 
Probabilitts,  Paris,  1812,  "the  theory  of  probabilities  is  nothing 
more  than  common-sense  reduced  to  calculation.  It  determines 
with  exactness  what  a  well-balanced  mind  perceives  by  a  kind  of 
instinct,  without  being  aware  of  the  process.  By  its  means  nothing 
is  left  to  chance  either  in  the  forming  of  an  opinion,  or  in  the  re- 
cognizing of  the  most  advantageous  view  to  select  when  the  occasion 
should  arise.  It  is,  therefore,  a  most  valuable  supplement  to  the 
ignorance  and  frailty  of  the  human  mind.  ..." 

I.  If  one  of  two  possible  events  occurs  in  such  a  way  that  one  of 
the  events  must  occur  in  a  ways,  the  other  in  b  ways,  the  probability 
that  the  first  will  happen  is  a/(a  +  b),  and  the  probability  that  the 
second  ivill  happen  is  b/(a  +  b).  If  a  rifleman  hits  the  centre  of 
a  target  about  once  every  twelve  shots  under  fixed  conditions  of 
light,  wind,  quality  of  powder,  etc.,  we  could  say  that  the  value 
of  his  chance  of  scoring  a  "  bullseye  "  in  any  future  shot  is  1  in  12, 
or  TT2,  and  of  missing,  11  in  12,  or  \\.  If  a  more  skilful  shooter 
hits  the  centre  about  five  times  every  twelve  shots,  his  chance  of 
success  in  any  future  shot  would  be  5  in  12,  or  -^,  and  of  missing 
T7.j.  Expressing  this  idea  in  more  general  language,  if  an  event 
can  happen  in  a  ways  and  fail  in  b  ways,  the  probability  of  the 
event 

Happening  =  ^^  J    Failing  =  j-j-j,      .  .  (l) 

provided  that  each  of  these  ways  is  just  as  likely  to  happen  as  to 
fail.     By  definition, 

Number  of  ways  the  event  occurs 

,a  • l  y  "~  Number  of  possible  ways  the  event  may  happen'  '  ' 

Example. — If  four  white,  and  six  black  balls  are  put  in  a  bag,  show  that 
the  probabiltty  that  a  white  ball  will  be  drawn  is  ^,  and  that  a  black  ball 
will  be  drawn  £.     In  betting  parlance,  the  odds  are  6  to  4  against  white. 


500  HIGHER  MATHEMATICS.  §  155. 

IL  If  p  denotes  the  probability  that  an  event  will  happen,  1-p 
denotes  the  probability  that  the  event  will  fail.  The  shooter  at  the 
target  is  certain  either  to  hit  or  to  miss.  In  mathematics,  unity  is 
supposed  to  represent  certainty,  therefore, 

Probability  of  hitting  +  Probability  of  missing  =  Certainty  =  1.         (3) 

If  the  event  is  certain  not  to  happen  the  probability  of  its  oc- 
currence is  zero.  Certainty  is  the  unit  of  probability.  Degrees 
of  probability  are  fractions  of  certainty. 

Of  course  the  above  terms  imply  no  quality  of  the  event  in 
itself,  but  simply  the  attitude  of  the  computer's  own  mind  with 
respect  to  the  occurrence  of  a  doubtful  event.  We  call  an  event 
impossible  when  we  cannot  think  of  a  single  cause  in  favour  of  its 
occurrence,  and  certain  when  we  cannot  think  of  a  single  cause 
antagonistic  to  its  occurrence.  All  the  different  "  shades  "  of 
probability — improbable,  doubtful,  probable — lie  between  these 
extreme  limits. 

Strictly  speaking,  there  is  no  such  thing  as  chance  in  Nature. 
The  irregular  path  described  by  a  mote  "  dancing  in  a  beam  of 
sunlight "  is  determined  as  certainly  as  the  orbit  of  a  planet  in  the 

heavens. 

All  nature  is  but  art,  unknown  to  thee ; 
All  chance,  direction  thou  can'st  not  see ; 
All  discord,  harmony  not  understood. 

The  terms  "  chance"  and  "probability"  are  nothing  but 
conventional  modes  of  expressing  our  ignorance  of  the  causes  of 
events  as  indicated  by  our  inability  to  predict  the  results.  "  Pour 
une  intelligence "  (omniscient),  says  Liagre,  "  tout  6venement  a 
venir  serait  certain  ou  impossible.'" 

III.  The  probability  that  both  of  two  independent  events  will 
happen  together  is  the  product  of  their  separate  probabilities.  Let 
p  denote  the  probability  that  one  event  will  happen,  q  the  probability 
that  another  event  will  happen,  the  probability,  P,  that  both  events 
will  happen  together  is 

P-M W 

This  may  be  illustrated  in  the  following  manner :  A  vessel  A 
contains  aY  white  balls,  b1  black  balls,  and  a  vessel  B  contains  a2 
white  balls  and  b2  black  balls,  the  probability  of  drawing  a  white 
ball  from  A  is  p1  =  a1/(a1  +  bj,  and  from  B,  p2  =  a2/(a2  +  b2).  The 
total  number  of  pairs  of  balls  that  can  be  formed  from  the  total 
number  of  balls  is  (ax  +  bj)  (a2  +  b2). 


§  155.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.     501 

Example. — In  any  simultaneous  drawing  from  each  vessel,  the  probability 
that 

Two  white  balls  will  occur  is  :  ala2j{a1  +  bx)  (a,£  +  62):;     .        .  (5) 

Two  black  balls  will  occur  is :  616a/(a1  +  bx)  (a2  +  62) ;  .         .       (6) 

White  ball  drawn  from  the  first,  black  ball  from  the  next,  is : 

fliVK  +  &i)  K  +  h) ;       •      ■      •      •     (?) 

Black  ball  drawn  from  the  first,  white  ball  from  the  next,  is : 

o2&i/K  +  &i)  K  +  h) ;      .      .      .      .     (R) 

Black  and  white  ball  occur  together,  is :  (axb2  +  bxa^l{al  +  bx)  (a2  +  &2)  ■  (9) 
The  sum  of  (5),  (6),  (9)  is  unity.  This  condition  is  required  by  the  above 
definition. 

An  event  of  this  kind,  produced  by  the  composition  of  several 
events,  is  said  to  be  a  compound  event.  To  throw  three  aces  with 
three  dice  at  one  trial  is  a  compound  event  dependent  on  the  con- 
currence of  three  simple  events.  Errors  of  observation  are  com- 
pound events  produced  by  the  concurrence  of  several  independent 
errors. 

Examples. — (1)  If  the  respective  probabilities  of  the  occurrence  of  each 
of  n  independent  errors  is  p,  q,  r . . . ,  the  probability  P  of  the  occurrence  of 
all  together  is  P=pxqxrx... 

(2)  If,  out  of  every  100  births,  49  are  male  and  51  female,  what  is  the 
probability  that  the  next  two  births  shall  be  both  boys ;  both  girls ;  and  a 
boy  first,  and  a  girl  next  ?  Ansr.  0-2401 ;  0-2601 ;  0-2499.  Hint,  ^fr  *  t4A  J 
roV  x  AV  >  tW  x  ^h' 

IV.  The  probability  of  the  occurrence  of  several  events  which 
cannot  occur  together  is  the  sum  of  the  probabilities  of  their 
separate  occurrences.  If  p,  q, . .  .  denote  the  separate  probabilities 
of  different  events,  the  probability,  P,  that  one  of  the  events  will 
happen  is, 

P  =  p  +  q+ (10) 

Examples. — (1)  A  bag  contains  12  balls  two  of  which  are  white,  four 
black,  six  red,  what  is  the  probability  that  the  first  ball  drawn  will  be  a 
white,  black,  or  a  red  one  ?  The  probability  that  the  ball  will  be  a  white  is 
I,  a  black  £,  etc.  The  probability  that  the  first  ball  drawn  shall  be  a  black  or 
a  white  ball  is  £. 

(2)  In  continuation  of  Ex.  (2),  preceding  set,  show  that  the  probability 
that  one  shall  be  a  boy  and  the  other  a  girl  is  0-4998  ;  and  that  both  shall  be 
of  the  same  sex,  0-5002.     Hint.  0-2499  +  0-2499  ;  0-2401  +  0-2601. 

V.  If  p  denotes  the  probability  that  an  event  will  happen  on  a 
single  trial,  the  probability,  P,  that  it  will  happen  r  times  in  n 
trials  is 

P.*-9--;(*-'  +  V-ri".      •     (ID- 

The  probability  that  the  event  will  fail  on  any  single  trial  is  1  -  p  ; 
the  probability  that  it  will  fail  every  time  is  (1  -  p)n.     The  proba- 


502  HIGHER  MATHEMATICS.  §  155. 

bility  that  it  will  happen  on  the  first  trial  and  fail  on  the  succeeding 
n  —  1  trials  is  p(l  -p)n~l,  from  (4).  But  the  event  is  just  as  likely 
to  happen  on  the  2nd,  3rd .  . .  trials  as  on  the  first.  Hence  the 
probability  that  the  event  will  happen  just  once  in  the  n  trials  is, 
from  (4),  and  (10),1 

(p  +  p  +  .  .  .  +  n  times)  x  (1  -  p)n~l ;  or,  np(l  -  p)n~l.  (12) 
The  probability  that  the  event  will  occur  on  the  first  two  trials 
and  fail  on  the  succeeding  n-2  trials  is  p2(l-p)n~2.  But  the 
event  is  as  likely  to  occur  during  the  1st  and  3rd,  2nd  and  4th, . .  . 
trials.  Hence  the  probability  that  it  will  occur  just  twice  during 
the  n  trials  is 

i»(»- 1)^(1 -;p)-2.         .  .  .         (13) 

The  probability  that  it  will  occur  r  times  in  n  trials  is,  therefore, 
represented  by  formula  (11). 

Examples. — (1)  What  is  the  probability  of  throwing  an  ace  exactly  three 
times  in  four  trials  with  a  single  die  ?  Ansr.  ^f T.  Hint,  n  =  4 ;  r  =  3 ; 
there  is  one  chance  in  six  of  throwing  an  ace  on  a  single  trial,  hence  p  =  %  ; 
n  -  r  =  1 ;  jr  =  (|):!;  1  - p  =  *.    Hence,  *i»2J  x  ^ 

(2)  What  is  the  probability  of  throwing  a  deuce  exactly  three  times  in 
three  trials  ?    Ansr.  ^.    n  =  3  ;  r  =  3 ;  (1  -p)n  ~r  =  5°  =  1;  p  =  £,  etc. 

VI.  If  p  denotes  the  very  small  probability  that  an  event  will 
happen  on  a  single  trial,  the  probability ■,  P,  that  it  will  happen  r 
times  in  a  very  great  number,  n,  trials  is 

(np)' 


r! 


(14) 


From  formula  (11),  however  small  p  may  be,  by  increasing  tne 
number  of  trials,  we  can  make  the  probability  that  the  event  will 
happen  at  least  once  in  n  trials  as  great  as  we  please.  The  proba- 
bility that  the  event  will  fail  every  time  in  n  trials  is  (1  -  p)n,  and  if 
p  be  made  small  enough  and  n  great  enough,  we  can  make  (1  -  p)n 
as  small  as  we  pease.2  If  n  is  infinitely  great  and  p  infinitely 
small,  we  can  write  n  =  n-l  =  n-2  =  ... 

w     -.  n(n-l)  _  „  (np)2 

.'.  (1  ~p)n=  l-np+     v2,     y~.  .  .  =1  -np  +  ^jf-  -...  (approx.)  ; 
(l-p)n=  e-*  (appro*.).  •  •  •         (15) 

1  The  student  may  here  find  it  necessary  to  read  over  §  190,  page  602. 

2  The  reader  should  test  this  by  substituting  small  numbers  in  place  of  p,  and 
large  ones  for  n.  Use  the  binomial  formula  of  §  97,  page  282.  See  the  remarks  on 
page  24,  §  11. 


§  155.    PKOBABILITY  AND  THE  THEORY  OF  ERRORS.     503 

(14)  follows  immediately  from  (11)  and  (15).     This  result  is  very 
important. 

Example. — If  n  grains  of  wheat  are  scattered  haphazard  over  a  sur- 
face s  units  of  area,  show  that  the  probability  that  a  units  of  -area  will 
contain  r  grains  of  wheat  is 


(anV 


r! 
Thus,  n  .  ds/s  represents  the  infinitely  small  probability  that  the  small  space 
ds  contains  a  grain  of  wheat.     If  the  selected  space  be  a  units  of  area,  we 
may  suppose  each  ds  to  be  a  trial,  the  number  of  trials  will,  therefore,  be 
aids.    Hence  we  must  substitute  anjs  for  np  in  (14)  for  the  desired  result. 

VII.  The  probability,  P,  that  an  event  will  occur  at  least  r 
times  in  n  trials  is 

P=2in  +  nVn-\l-v)+  n(n2~ 1 V"2  (*  ~Vf  +  •  •  •  to  (*  -  r)  terms  (16) 

For  if  it  cccur  every  time,  or  fail  only  once,  twice,  .  .  .  ,  ovn  -  r 
times,  it  occurs  r  times.  The  whole  probability  ol  its  occurring  at 
least  r  times  is  therefore  the  sum  of  its  occurring  every  time,  of 
failing  only  once,  twice,  .  .  :  ,  n  -  r  times,  etc. 

Example. — What  is  the  probability  of  throwing  a  deuce  three  times  at 
least  in  four  trials?  Ansr.  x|,.  Here^«  =  (£)4;  and  the  next  term  of  (16)  is 
4  x  5  x  (*)*. 

Sometimes  a  natural  process  proves  far  too  complicated  to  admit 
of  any  simplification  by  means  of  "  working  hypotheses  ".  The 
question  naturally  arises,  can  the  observed  sequence  of  events  be 
reasonably  attributed  to  the  operation  of  a  law  of  Nature  or  to  chance? 
For  example,  it  is  observed  that  the  average  of  a  large  number  of 
readings  of  the  barometer  is  greater  at  nine  in  the  morning  than 
at  four  in  the  afternoon;  Laplace  {Theorie  analytique  da  Proba- 
bility, Paris,  49, 1820)  asked  whether  this  was  to  be  ascribed  to  the 
operation  of  an  unknown  law  of  Nature  or  to  chance  ?  Again,  G. 
Kirchhoff  (Monatsberichte  der  Berliner  Ahademie,  Oct.,  1859)  inquired 
if  the  coincidence  between  70  spectral  lines  in  iron  vapour  and  in 
sunlight  could  reasonably  be  attributed  to  chance.  He  found  that 
the  probability  of  a  fortuitous  coincidence  was  approximately  as 
1  :  1,000000,000000.  Hence,  he  argued  that  there  can  be  no 
reasonable  doubt  of  the  existence  of  iron  in  the  sun.  Mitchell 
{Phil.  Trans.,  57,  243,  1767;  see  also  Kleiber,  Phil.  Mag.,  [5],  2*, 
439,  1887)  has  endeavoured  to  calculate  if  the  number  of  star 
clusters  is  greater  than  what  would  be  expected  if  the  stars  had  been 
distributed  haphazard  over  the  heavens.     A.  Schuster  (Proc.  Boy. 


504  HIGHER  MATHEMATICS.  §  156. 

Soc,  31,  337,  1881)  has  tried  to  answer  the  question,  Is  the  number 
of  harmonic  relations  in  the  spectral  lines  of  iron  greater  than  what 
a  chance  distribution  would  give  ?  Mallet  (Phil  Trans.,  171, 1003, 
1880)  and  R.  J.  Strutt  (Phil.  Mag.,  [6],  1,  311,  1901)  have  asked,  Do 
the  atomic  weights  of  the  elements  approximate  as  closely  to  whole 
numbers  as  can  reasonably  be  accounted  for  by  an  accidental  co- 
incidence ?  In  other  words  :  Are  there  common-sense  grounds  for 
believing  the  truth  of  Prout's  law,  that  "  the  atomic  weights  of  the 
other  elements  are  exact  multiples  of  that  of  hydrogen  "  ? 

The  theory  of  probability  does  not  pretend  to  furnish  an  in- 
fallible criterion  for  the  discrimination  of  an  accidental  coincidence 
from  the  result  of  a  determining  cause.  Certain  conditions  must 
be  satisfied  before  any  reliance  can  be  placed  upon  its  dictum. 
For  example,  a  sufficiently  large  number  of  cases  must  be  avail- 
able. Moreover,  the  theory  is  applied  irrespective  of  any  know- 
ledge to  be  derived  from  other  sources  which  may  or  may  not 
furnish  corroborative  evidence.  Thus  KirchhofF s  conclusion  as  to 
the  probable  existence  of  iron  in  the  sun  was  considerably 
strengthened  by  the  apparent  relation  between  the  brightness  of 
the  coincident  lines  in  the  two  spectra. 

For  details  of  the  calculations,  the  reader  must  consult  the 
original  memoirs.  Most  of  the  calculations  are  based  upon  the 
analysis  in  Laplace's  old  but  standard  Theorie  (I.e.).  An  excellent 
resume  of  this  latter  work  will  be  found  in  the  Encyclopedia  Metro- 
politan. The  more  fruitful  applications  of  the  theory  of  prob- 
ability to  natural  processes  have  been  in  connection  with  the  kinetic 
theory  of  gases  and  the  "law  "  relating  to  errors  of  observation. 

§  156.  Application  to  the  Kinetic  Theory  of  Gases. 

The  purpose  of  the  kinetic  theory  of  gases  is  to  explain  the 
physical  properties  of  gases  from  the  hypothesis  that  a  gas  consists 
of  a  great  number  of  molecules  in  rapid  motion.  The  following 
illustrations  are  based,  in  the  first  instance,  on  a  memoir  by 
R.  Clausius  (Phil.  Mag.,  [4],  17,  81,  1859).  For  further  develop- 
ments, O.  E.  Meyer's  The  Kinetic  Theory  of  Gases,  London,  1899, 
may  be  consulted. 

I.  To  shoio  that  the  probability  that  a  single  molecule,  moving 
in  a  swarm  of  molecules  at  rest,  xuill  traverse  a  distance  x  without 
collision  is 


§  156.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.       505 

P-V"1,  .  .  .  .  (17) 
where  I  denotes  the  probable  value  of  the  free  path  the  molecule 
can  travel  without  collision,  and  x/l  denotes  the  ratio  of  the  path 
actually  traversed  to  the  mean  length  of  the  free  path.  "  Free 
path"  is  denned  as  the  distance  traversed  by  a  molecule  between 
two  successive  collisions.  The  '*  mean  free  path  "  is  the  average 
of  a  great  number  of  free  paths  of  a  molecule.  Consider  any 
molecule  moving  under  these  conditions  in  a  given  direction.  Let 
a  denote  the  probability  that  the  molecule  will  travel  a  path  one 
unit  long  without  collision,  the  probability  that  the  molecule  will 
travel  a  path  two  units  long  is  a .  a,  or  a2,  and  the  probability  that 
the  molecule  will  travel  a  path  x  units  long  without  collision  is, 
from  (4), 

P  =  a',  .         .         .         .       (18) 

where  a  is  a  proper  fraction.  Its  logarithm  is  therefore  negative. 
(Why  ?)  If  the  molecules  of  the  gas  are  stationary,  the  value  of 
a  is  the  same  whatever  the  direction  of  motion  of  the  single  mole- 
cule.    From  (15),  therefore, 

X 

p  =  0~I 

where  I  =~l/log  a.  We  can  get  a  clear  idea  of  the  meaning  of  this 
formula  by  comparing  it  with  (15).  Supposing  the  traversing  of 
unit  path  is  reckoned  a  "  trial,"  x  in  (17)  then  corresponds  with  n 
in  (15).  l/l  in  (17)  replaces  p  in  (15).  l/l,  therefore,  represents 
the  probability  that  an  event  (collision)  will  happen  during  one 
trial.  If  I  trials  are  made,  a  collision  is  certain  to  occur.  This  is 
virtually  the  definition  of  mean  free  path. 

II.  To  show  that  the  length  of  the  path  which  a  molecule,  moving 
amid  a  swarm  of  molecules  at  rest  can  traverse  without  collision  is 
probably 

Xs 
Z  =  4- (19) 

where  X  denotes  the  mean  distance  between  any  two  neighbouring 
molecules,  p  the  radius  of  the  sphere  of  action  corresponding  to  the 
distance  apart  of  the  molecules  during  a  collision,  -n-  is  a  constant 
with  its  usual  signification.  Let  unit  volume  of  the  gas  contain  N 
molecules.  Let  this  volume  be  divided  into  N  small  cubes,  each  of 
which,  on  the  average,  contains  only  one  molecule.  Let  X  denote 
the  length  of  the  edge  of  one  of  these  little  cubes.  Only  one  mole- 
cule is  contained  in  a  cube  of  capacity  A.3.     The  area  of  a  cross 


506  HIGHEE  MATHEMATICS.  §  156. 

section  through  the  centre  of  a  sphere  of  radius  p,  is  trp2,  (13), 
page  604.  If  the  moving  molecule  travels  a  distance  A,  the  hemi- 
spherical anterior  surface  of  the  molecule  passes  through  a 
cylindrical  space  of  volume  7rp2A  (26),  page  605.  Therefore,  the 
probability  that  there  is  a  molecule  in  the  cylinder  7rp2A  is  to  1  as 
Trp2k  is  to  A3,  that  is  to  say,  the  probability  that  the  molecule  under 
consideration  will  collide  with  another  as  it  passes  over  a  path  of 
length  A,  is  7rp2A  :  A3.    The  probability  that  there  will  be  no  collision 

is  1  -  \C-     From  (17), 

-K  P27T 

P  =  e     i  -  1  -  ^-.       .        .        .       (20) 

According  to  the  kinetic  theory,  one  fundamental  property  of 
gases  is  that  the  intermolecular  spaces  are  very  great  in  compari- 
son with  the  dimensions  of  the  molecules,  and,  therefore,  /r7r/A2  is 
very  small  in  comparison  with  unity.  Hence  also  Xjl  is  a  small 
magnitude  in  comparison  with  unity.  Expand  e~kl1  according  to 
the  exponential  theorem  (page  285),  neglect  terms  involving  the 
higher  powers  of  A,  and 

-K              A 
e     <  =  1  -  i (21) 

From  (20)  and  (21), 

A  3  ffllTX. 

l  =  ±~;or,P=e--W.      .         .         .       (22) 


p- 


7T 


Example. — The  behaviour  of  gases  under  pressure  indicates  that  p  is  very 
much  smaller  than  A.  Hence  show  that  "  a  molecule  passes  by  many  other 
molecules  like  itself  before  it  collides  with  another".  Hint.  From  the  first 
of  equations  (22),  I :  A  =  A2  :  p2ir.     Interpret  the  symbols. 

III.  To  show  that  the  mean  value  of  the  free  path  of  n  molecules 
moving  under  the  same  conditions  as  the  solitary  molecule  just  con- 
sidered, is 

•*-£.  .        .    •     .        .      (23) 

Out  of  n  molecules  which  travel  with  the  same  velocity  in  the 
same  direction  as  the  given  molecule,  ne~xl1  will  travel  the  distance 
x  without  collision,  and  ne-(x  +  dx)l1  will  travel  the  distance  x  +  dx 
without  collision.     Of  the  molecules  which  traverse  the  path  x, 

n(el-e~     ■    \  =  ne   l(l-e    l)  =  ^e ' ldx, 

of  them  will  undergo  collision  in  passing  over  the  distance  dx. 
The  last  transformation  follows  directly  from  (21).     The  sum  of 


§  15G.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      507 

all  the  paths  traversed  by  the  molecules  passing  x  and  x  +  dx  is 

x     —  - 
jne    ldx. 

Since  each  molecule  must  collide  somewhere  in  passing  between 
the  limits  x  =  0  and  x  =  go,  the  sum  of  all  the  possible  paths 
traversed  by  the  n  molecules  before  collision  is 


r  x 


ldx, 


and  the  mean  value  of  these  n  free  paths  is 

Integrate  the  indefinite  integral  as  indicated  on  page  205.  From 
(4)  we  get  (23).  This  represents  the  mean  free  path  of  these 
molecules  moving  with  a  uniform  velocity. 

Examples. — (1)  A  molecule  moving  with  a  velocity  V  enters  a  space 
filled  with  n  stationary  molecules  of  a  gas  per  unit  volume,  what  is  the  prob- 
ability that  this  molecule  will  collide  with  one  of  those  at  rest  in  unit  time  ? 
Use  the  above  notation.  The  molecule  travels  the  space  Fin  unit  time.  In 
doing  this,  it  meets  with  imp2 V  molecules  at  rest.  The  probable  number  of 
collisions  in  unit  time  is,  therefore,  irnpW,  which  represents  the  probability 
of  a  collision  in  unit  time. 

(2)  Show  that  the  probable  number  of  collisions  made  in  unit  time  by  a 
molecule  travelling  with  a  uniform  velocity  F,  in  a  swarm  of  N  molecules  at 
rest,  is 

f-^ <-> 

What  is  the  relation  between  this  and  the  preceding  result?  Note  the 
number  of  collisions  =  V/l ;  and  N\"  =  1. 

IV.  -The  number  of  collisions  made  in  unit  time  by  a  molecule 
moving  with  uniform  velocity  in  a  direction  which  makes  an  angle  0 
with  the  direction  of  motion  of  a  swarm  of  molecules  also  moving 
with  the  same  uniform  velocity  is  'probably 

9-^2v  sin  10.  ...       (25) 

Let  v  be  the  resultant  velocity  of  one  molecule,  and  xv  yly  zl 
the  three  component  velocities,  then,  from  the  parallelopiped  of 
velocities,  page  125, 

v*  =  x*  +  y*  +  V  | 

xY  =  v  cos  #! ;  y1  =  v  sin  01 .  cos  fa ;  zY  =  v  sin  01 .  sin  fa)'  ^  ' 
If  one  set  of  molecules  moves  with  a  uniform  velocity,  v,  whose 
components  are  x,  y,  z,  relative  to  the  given  molecule  moving  with 


508  HIGHER  MATHEMATICS.  §  156. 

the  same  uniform  velocity,  v,  whose  components  are  xv  yv  zv  then, 

v*  =  X2  +  y2  +  Z2  .  <         t         ,       (27) 

#  =  v  cos  0 ;  ?/  =  v  sin  # .  cos  0 ;  z  =  sin  0 .  cos  <£,  .  (28) 
and  the  relative  resultant  velocity,  v,  of  one  molecule  with  respect 
to  the  other  considered  at  rest,  is 

V=  J  (xx  -  x)2  +  (Vl  -  yf  +  (z,  -  z)\         .       (29) 
If  we  choose  the  three  coordinate  axes  so  that  the  ic-axis  coincides 
with  the  direction  of  motion  of  the  given  molecule,  we  may  sub- 
ititute  these  values  in  (26),  remembering  that  cos  0°  =  1,  sin  0°  =  0, 
yi  =  0;  z^O;  r.x^v.  .         .       (30) 

Substitute  (30)  and  (26)  in  (29),  we  get 

V=  J  (v  -  v  cos  0)2  +  v2sin20 .  cos2<£  +  *;2sin2<9 .  sin2</> ; 
.-.  V=  V  v2  -  2v2cos  6  +  v2cos20  +  ?;2sin20, 
since  sin2<£  +  cos2<£  =  1.      Similarly,  cos20  +  sin20  =  1,  and  con- 
sequently 

V  =  v  J  2  -  2  cos  0  =  v  J  2(1  -  cos0). 
But  we  know,  page  612,  that  1  -  cos  x  =  2(sin  \x)2,  hence, 

V=2v  sin  ^0.  .         .         .       (31) 

Having  found  the  relative  velocity  of  the  molecules,  it  follows 
directly  from  (24)  and  (31),  that 

Number  of  collisions  =      .  3  -  =  ~rj  AV  Sin  \v. 

V.  The  number  of  collisions  encountered  in  unit  time  by  a  mole- 
cule moving  in  a  swarm  of  molecules  in  all  directions,  is 

i:S?E  f32i 

Let  V  denote  the  velocity  of  the  molecules,  then  the  different 
motions  can  be  resolved  into  three  groups  of  motions  according  to 
the  parallelopiped  of  velocities.  Proceed  as  in  the  last  illustration. 
The  number  of  molecules,  n,  moving  in  a  direction  between  0  and 
0  +  dO  is  to  the  total  number  of  molecules,  N,  in  unit  volume  as 

n :  N  =  2tt  sin  OdO  :  4tt  ;  or,  n  =  $N  sin  OdO.  .  (33) 
Since  the  angle  0  can  increase  from  0°  to  180°,  the  total  number  of 
collisions  is        » 

Vp27r    n      Vp2*  .    6    1   .     ... 
~~XF'N  =  T3-sm  2  '  2  Sm  m- 

To  get  the  total  number  of  collisions,  it  only  remains  to  integrate 
for  all  directions  of  motion  between  0°  and  180°.  Thus  if  A  denotes 
the  number  of  collisions. 


§  156.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.       509 

A  =  *—§- 1  sing,  sin  Odd  =     ,3     I  sin2n  .  cos  -zdO  =  k  .    U- , 

Dy  the  method  of  integration  on  page  212. 

Example. — Find  the  length  of  the  free  path  of  a  molecule  moving  in  a 
swarm  of  molecules  moving  in  all  directions,  with  a  velocity  V.     Ansr. 

Length  of  free  path  =  VJA  =  fxVp8*.     .        .        .       (34) 
For  the  hypothesis  of  uniform  velocity  see  §  164,  page  534. 

VI.  Assuming  that  two  unlike  molecules  combine  during  a  colli- 
sion, the  velocity  of  chemical  reaction  between  two  gases  is 

d  t 
S  =  OT, (35) 

where  N  and  N'  are  the  number  of  molecules  of  each  of  the  two 
gases  respectively  contained  in  unit  volume  of  the  mixed  gases, 
dx  denotes  ihe  number  of  molecules  of  one  gas  in  unit  volume 
which  combines  with  the  other  in  the  time  dt ;  k  is  a  constant. 
Let  the  two  gases  be  A  and  B.  Let  A.  and  A'  respectively  denote 
the  distances  between  two  neighbouring  molecules  of  the  same 
kind,  then,  as  above, 

JVX8  =  NX!3  =  1.  ...       (36) 

Let  p  be  the  radius  of  the  sphere  of  action,  and  suppose  the  mole- 
cules combine  when  the  sphere  of  action  of  the  two  kinds  of 
molecules  approaches  within  2p,  it  is  required  to  find  the  rate  of 
combination  of  the  two  gases.  The  probability  that  a  B  molecule 
will  come  within  the  sphere  of  action  of  an  A  molecule  in  unit  time 
is  Virpt/X?,  by  (24).     Among  the  N  molecules  of  B, 

2 

N^-Vdt;  or,  NN'irpWdt,  .         .       (37) 

by  (36),  combine  in  the  time  dt.     But  the  number  of  molecules 

which  combine  in  the  time  dt  is  -  dN  =  -  dN',  or,  from  (37), 

dN  =  dN  =  -NNirpWdt. 

If  dx  represents  the  number  of  molecules  in  unit  volume  which 

combines  in  the  time  dt, 

die 
dx  =  dN  =  dJSr  =  Trp*VNNdt.     .-.  -^  =  kNW, 

by  collecting  together  all  the  constants  under  the  symbol  k.  This 
will  be  at  once  recognized  as  the  law  of  mass  action  applied  to 
bimolecular  reactions.  J.  J.  Thomson's  memoir,  "  The  Chemical 
Combination  of  Gases,"  Phil.  Mag.,  [5],  18,  233,  1884,  might  now 
be  read  by  the  chemical  student  with  profit. 


510  HIGHER  MATHEMATICS.  §  157. 

§  157.    Errors  of  Observation. 

If  a  number  of  experienced  observers  agreed  to  test,  indepen- 
dently, the  accuracy  of  the  mark  etched  round  the  neck  of  a  litre 
flask  with  the  greatest  precision  possible,  the  inevitable  result 
would  be  that  every  measurement  would  be  different.  Thus,  we 
might  expect 

1-0003;  0-9991;  1-0007;  1-0002;  1-0001;  0-9998;... 
Exactly  the  same  thing  would  occur  if  one  observer,  taking  every 
known  precaution  to  eliminate  error,  repeats  a  measurement  a 
great  number  of  times.     The  discrepancies  doubtless  arise  from 
various  unknown  and  therefore  uncontrolled  sources  of  error. 

We  are  told  that  sodium  chloride  crystallizes  in  the  form  of 
cubes,  and  that  the  angle  between  two  adjoining  faces  of  a  crystal 
is,  in  consequence,  90°.  As  a  matter  of  fact  the  angle,  as  measured, 
varies  within  0'5°  either  way.  No  one  has  yet  exactly  verified 
the  Gay  Lussac-Humboldt  law  of  the  combination  of  gases ;  nor 
has  any  one  yet  separated  hydrogen  and  oxygen  from  water,  by 
electrolysis,  in  the  proportions  required  by  the  ratio,  2H2  :  02. 

The  irregular  deviations  of  the  measurements  from,  say,  the 
arithmetical  mean  of  all  are  called  accidental  errors.  In  the 
following  discussion  we  shall  call  them  "  errors  of  observa- 
tion "  unless  otherwise  stated.  These  deviations  become  more 
pronounced  the  nearer  the  approach  to  the  limits  of  accurate 
measurement.  Or,  as  Lamb l  puts  it,  "  the  more  refined  the 
methods  employed  the  more  vague  and  elusive  does  the  supposed 
magnitude  become  ;  the  judgment  flickers  and  wavers,  until  at 
last  in  a  sort  of  despair  some  result  is  put  down,  not  in  the  belief 
that  it  is  exact,  but  with  the  feeling  that  it  is  the  best  we  can 
make  of  the  matter".  It  is  the  object  of  the  remainder  of  this 
chapter  to  find  what  is  the  best  we  can  make  of  a  set  of  discordant 
measurements. 

The  simplest  as  well  as  the  most  complex  measurements  are 
invariably  accompanied  by  these  fortuitous  errors.  Absolute 
agreement  is  itself  an  accidental  coincidence.  Stanley  Jevons 
says,  "it  is  one  of  the  most  embarrassing  things  we  can  meet 
when  experimental  results  agree  too  closely".  Such  agreement 
should  at  onGe  excite  a  feeling  of  distrust. 

!H.  Lamb,  Presidential  Address,  B.A.  meeting,  1904  ;  Nature,  70,  372,  1904. 


§  158.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      511 


Fig.  165. 


The  observed  relations  between  two  variables,  therefore,  should 
not  be  represented  by  a  point  in  space,  rather  by  a  circle  around 
whose  centre  the  different  observations  will  be  grouped  (Fig.  165). 
Any  particular  observation  will  find  a  place 
somewhere  within  the  circumference  of  the 
circle.  The  diagram  (Fig.  165)  suggests  our  old 
illustration,  a  rifleman  aiming  at  the  centre  of  a 
target.  The  rifleman  may  be  likened  to  an  ob- 
server ;  the  place  where  the  bullet  hits,  to  an 
observation ;  the  distance  between  the  centre 
and  the  place  where  the  bullet  hits  the  target 
resembles  an  error  of  observation.  A  shot  at  the  centre  of  the 
target  is  thus  an  attempt  to  hit  the  centre,  a  scientific  measure- 
ment is  an  attempt  to  hit  the  true  value  of  the  magnitude 
measured  (Maxwell). 

The  greater  the  radius  of  the  circle  (Fig.  165),  the  cruder  and 
less  accurate  the  measurements ;  and,  vice  versd,  the  less  the 
measurements  are  affected  by  errors  of  observation,  the  smaller 
will  be  the  radius  of  the  circle.  In  other  words,  the  less  the  skill 
of  the  shooter,  the  larger  will  be  the  target  required  to  record  his 
attempts  to  hit  the  centre. 


§  158.  The  "Law"  of  Errors. 

These  errors  may  be  represented  pictorially  another  way. 
Draw  a  vertical  line  through  the  centre  of  the  target  (Fig.  165) 
and  let  the  hits  to  the  right  of  this  line  represent  positive  errors 
and  those  to  the  left  negative  errors.  Suppose  that  500  shots  are 
fired  in  a  competition  j  of  these,  ten  on  the  right  side  were  be- 
tween 0*4  and  0*5  feet  from  the  centre  of  the  target ;  twenty  shots 
between  0*3  and  0'4  feet  away ;  and  so  on,  as  indicated  in  the 
following  table. 


Positive 
Deviations  from 
Mean  between 

Number 
of  Error*. 

Percentage 
Number 
of  Errors. 

Negative 

Deviations  from 

Mean  between 

Number 
of  Errors. 

Percentage 
Number 
of  Errors. 

0-4  and  0  5 
0-3  and  0-4 
0-2  and  0-3 
0-1  and  0-2 
0-0  and  0-1 

10 
20 
40 
80 
100 

2 

4 

8 

16 

20 

0-4  and  0-5 
03  and  0-4 
0-2  and  0-3 
0-1  and  0-2 
0-0  and  0*1 

10 
20 
40 
80 
100 

2 

4 

8 

16 

20 

512 


HIGHER  MATHEMATICS. 


8  158. 


y\ 

r 

+x 


Fig.  166.— Probability 
Curve. 


Plot,  as  ordinates,  the  numbers  in  the  third  column  with  the 
means  of  the  two  corresponding  limits  in 
the  first  column  as  abscissae.  The  curve 
shown  in  Pig.  166  will  be  the  result. 

By  a  study  of  the  last  two  diagrams, 
we  shall  find  that  there  is  a  regularity  in 
the  grouping  of  these  irregular  errors  which, 
as  a  matter  of  fact,  becomes  more  pro- 
nounced the  greater  the  number  of  trials 
we  take  into  consideration.     Thus,  it  is  found  that — 

1.  Small  errors  are  more  frequent  than  large  ones. 

2.  Positive  errors  are  as  frequent  as  negative  errors. 

3.  Very  large  positive  or  negative  errors  do  not  occur. 

Any  mathematical  relation  between  an  error,  x,  and  the  frequency, 
or  rather  the  probability,  of  its  occurrence,  y,  must  satisfy  these 
characteristics.     When  such  a  function, 

2/ =/(*)> 
is  plotted,  it  must  have  a  maximum  ordinate  corresponding  with  no 
error ;  it  must  be  symmetrical  with  respect  to  the  ?/-axis,  in  order 
to  satisfy  the  second  condition  ;  and  as  x  increases  numerically, 
y  must  decrease  until,  when  x  becomes  very  large,  y  must  become 
vanishingly  small.     Such  is  the  curve  represented  by  the  equation, 

y  =  ke-h2*\      .         .         .         .         (1) 
where  h  and  k  are  constants.1     The  graph  of  this  equation,  called 
the  probability  curYe,  or  curve  of  frequency,  or  curve  of  errors, 
is   obtained   by  assigning  arbitrary  constant 
values  to  h  and  k  and  plotting  a  set  of  corre- 
sponding values  of  x  and  y  in  the  usual  way.2 
I.  To  find  a  meaning  for  the  constant  k. 
Put  x  =  0,  then  y  =  k}  that  is  the  maximum 
ordinate  of  the  curve.     Now  put  h  =  1,  and 
make  k  successively  |,  1,  2,  3,  4.     Plot  cor- 
responding values  of  x  and  y,  as   shown  in 
Fig.  167.     Another  plan  is  "to  bend  a  loop  of 
wire  into  the  form  of  one  of  the  curves,  and 

Fig.  167.— Probability  to   place   a  lamp   behind   it  so   as  to   throw 

Curves  (h  constant,  k  r  ,  t  m,       ,  , 

variable).  the  shadow  upon  the  screen.     Ine  loop  and 

lamp  might  be  easily  made  to  move  in  such 


-   -/V 

o 

f     \ 

HX£- 

tt^& 

UZ^r- 

-$-^x& 

JM/  ^% 

2         10          1? 

1  Use  Table  XVII.,  page  626. 

2  E.  B.  Sargant,  "The  Education  of  Examiners, 


Mature,  70,  63,  1904. 


§  158.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      513 

a  manner  that  the  shadows  in  the  successive  positions  gave 
the  whole  series  of  curves."  If  we  agree  to  define  an  error  as 
the  deviation  of  each  measurement  from  the  arithmetical  mean,  k 
corresponds  with  those  measurements  which  coincide  with  the 
mean  itself,  or  are  affected  by  no  error  at  all.  The  height  at 
which  the  curve  outs  the  y-&xis  represents  the  frequency  of  oc- 
currence of  the  arithmetical  mean  •  k  has  nothing  to  do  with  the 
actual  shape  of  the  curve  beyond  increasing  the  length  of  the 
maximum  ordinate  as  the  accuracy  of  the  observations  increases. 

II.  To  find  a  meaning  for  the  constant  h.  Put  k  =  1,  and  plot 
corresponding  values  of  x  and  y  for  x  =  ±  0*3,  ±  0*4,  +  0*5,  +  0-6, 
. . .  when  h  =  £,  \,  1,  2,  3, . . .,  as  shown  in  Fig.  168.  In  this  way, 
it  will  be  observed  that  although  all  the  curves  cut  the  y-axis  at 


Fig.  168. — Probability  Curves  (k  constant,  h  variable). 

the  same  point,  the  greater  the  value  of  h,  the  steeper  will  be  the 
curve  in  the  neighbourhood  of  the  central  ordinate  Oy.  The 
physical  signification  of  this  is  that  the  greater  the  magnitude  of 
h,  the  more  accurate  the  re- 
sults and  the  less  will  be  the 
magnitude  of  the  deviation 
of  individual  measurements 
from  the  arithmetical  mean 
of  the  whole  set.  Hence 
Gauss  calls  h  the  absolute 
measure  or    modulus    of 


precision. 

III.  When  h  and  k  both    ^x 
vary,    we    get    the    set    of 
curves  shown  on  Fig.  169. 


V  _ 


+  JC 


Fig.  169.— Probability  Curves  (h  and  k 
both  variable). 


KK 


514 


HIGHER  MATHEMATICS. 


§158. 


A  good  shot  will  get  a  curve  enclosing  a  very  much  smaller  area 
than  one  whose  shooting  is  wild. 

We  must  now  submit  our  empirical  "law"  to  the  test  of 
experiment.  Bessel  has  compared  the  errors  of  observation  in 
470  astronomical  measurements  made  by  Bradley  with  those 
which  should  occur  according  to  the  law  of  errors.  The  results 
of  this  comparison  are  shown  in  the  following  table : — x 


Number  of  Errors  of  each 

Magnitude  of  Error  in 

Magnitude. 

Parts  of  a  Second  of 
Arc,  between : 

Observed. 

Theory. 

0     and  0-1 

94 

95 

04  and  0-2 

88 

89 

0*2  and  0-3 

78 

78 

0-3  and  0-4 

68 

64 

0-4  and  05 

51 

50 

0-5  and  06 

36 

36 

0-6  and  07 

26 

24 

0*7  and  0*8 

14 

15 

0-8  and  0-9 

10 

9 

0-9  and  1-0 

7 

5 

above  1*0 

8 

5 

This  is  a  remarkable  verification  of  the  above  formula.  The  theory, 
be  it  observed,  provides  for  errors  of  any  magnitude,  however 
large  j  in  practice,  there  is  a  limit  above  which  no  error  will  be 
found  to  occur.  The  dots  in  Fig.  170  represent  the  "observed 
errors"  in  some  determinations  of  the  velocity  of  light.  The 
graph,  plotted  from  the  error  curve 

2/  =  8-9e-°025a:2, 

as  you  can  see,  is  almost  a  faithful  representation  of  the  actual 

errors.     Airy  and  Newcomb  have  also 

shown  that  the  number  and  magnitude 

of  the  errors  affecting  extended  series 

of  observations  are  in  fair  accord  with 

theory.     But  in  every  case,  the  number 
Fig  170 

of  large  errors  actually  found  is  in  excess 

of  theory.     To  quote  an  instance.     S.  Newcomb  examined  684  ob- 


1  Taken  from  Encke's  paper  in  the  Berliner  Astronomisches  Jahrbuch,  249,  1834; 
or,  Taylor's  Scientific  Memoirs,  2,  317,  1841. 


§  158.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      515 

servations  of  the  transit  of  Mercury.  According  to  the  "law"  of 
errors,  there  should  be  5  errors  numerically  greater  than  ±  27". 
In  reality  49  surpassed  these  limits.  You  can  also  notice  how  the 
"big"  errors  accumulate  at  the  ends  of  the  frequency  curve  in 
Fig.  170. 

The  theory  assumes  that  the  observations  are  all  liable  to  the 
same  errors,  but  differ  in  the  accidental  circumstances  which  give 
rise  to  the  errors.1  Equation  (1)  is  by  no  means  a  perfect  repre- 
sentation of  the  law  of  errors.  The  truth  is  more  complex.  The 
magnitude  of  the  errors  seems  to  depend,  in  some  curious  way, 
upon  the  number  of  observations.  In  an  extended  series  of 
observations  the  errors  may  be  arranged  in  groups.  Each  group 
has  a  different  modulus  of  precision.  This  means  that  the  mod- 
ulus of  precision  is  not  constant  throughout  an  extended  series  of 
observations.  See  Encyc.  Brit.,  F.  Y.  Edgeworth's  art.  "  Law  of 
Error,"  28,  280,  1902.  But  the  probability  curve  represented  by 
the  formula 

y  =  ke-#*\ 

may  be  considered  a  very  fair  graphic  representation  of  the  law 
connecting  the  probability  of  the  occurrence  of  an  error  with  its 
magnitude.  All  our  subsequent  work  is  based  upon  this  empirical 
law  !  J.  Venn  in  his  Logio  of  Chance,  1896,  calls  the  "  exponential 
law  of  errors,"  a  law,  because  it  expresses  a  physical  fact  relating  to 
the  frequency  with  which  errors  are  found  to  present  themselves  in 
practice;  while  the  "method  of  least  squares  is  a  rule  showing 
how  the  best  representative  value  may  be  extracted  from  a  set  of 
experimental  results.  H.  Poincare,  in  the  preface  to  his  Thermo- 
dynamique,  Paris,  1892,  quotes  the  laconic  remark,  "everybody 
firmly  believes  in  it  (the  law  of  errors),  because  mathematicians 
imagine  that  it  is  a  fact  of  observation,  and  observers  that  it  is  a 
theorem  of  mathematics  ". 

Adrian  (1808)  appears  to  have  been  the  first  to  try  to  deduce 
the  above  formula  on  theoretical  grounds.  Several  attempts  have 
since  been  made,  notably  by  Gauss,  Hagen,  Herschel,  Laplace, 
etc.,  but  I  believe  without  success. 


1  Some  observers'  results  seem  more  liable  to  these  large  errors  than  others,  due, 
perhaps,  to  carelessness,  or  lapses  of  attention.  Thomson  and  Tait,  I  presume,  would 
call  the  abnormally  large  errors  "avoidable  mistakes  ". 

KK* 


1 


516  HIGHER  MATHEMATICS.  §  159. 

§  159.    The  Probability  Integral. 

Let  rc0,  xv  x2, . .  .x  be  a  series  of  errors  in  ascending  order  of 
magnitude  from  xQ  to  x.  Let  the  differences  between  the  succes- 
sive values  of  x  be  equal.  If  x  is  an  error,  the  probability  of 
committing  an  error  between  x0  and  x  is  the  sum  of  the  separate 
probabilities  ke~h2z*,  ke~^\2 . . .,  ,(10),  page  501,  or 

P  =  k(e-»W  +  e-»27  4y.i-h %<?-**.  •  (1) 
If  the  summation  sign  is  replaced  by  that  of  integration,  we  must 
let  dx  denote  the  successive  intervals  between  any  two  limits  x0 
and  x,  thus 

Now  it  is  certain  that  all  the  errors  are  included  between  the  limits 
+  oo,  and,  since  certainty  is  represented  by  unity,  we  have 

?jB]jJ-^-*-JT'      '      *      (2) 

from  page  345.     Or, 

k  =  -j*.dx (3) 

Substituting  this  value  of  k  in  the  probability  equation  (1),  pre- 
ceding section,  we  get  the  same  relation  expressed  in  another 
form,  namely, 

—re-,Mdat     ....        (4) 

V7T 

a  result  which  represents  the  probability  of  errors  of  observation 
between  the  magnitudes  x  and  x  +  dx.     By  this  is  meant  the  ratio : 

Number  of  errors  between  x  and  x  +  dx 
Total  number  of  errors 

The  symbols  y  and  P  are  convenient  abbreviations  for  this  cumbrous 
phrase.  For  a  large  number  of  observations  affected  with  accidental 
errors,  the  probability  of  an  error  of  observation  having  a  magnitude 
a;,  is, 

V  =  Ke-**dx,  (5) 

which  is  known  as  Gauss'  law  of  errors.  This  result  has  the 
same  meaning  as  y  —  ke-tf*2  of  the  preceding  section.  In  (4)  dx  re- 
presents the  interval,  for  any  special  case,  between  the  successive 
values  of  x.  For  example,  if  a  substance  is  weighed  to  the 
thousandth  of  a  gram,  dx  =  0*001 ;  if  in  hundredths,  dx  =  0*01, 


§  159.  PROBABILITY  AND  THE  THEORY  OF  ERRORS.  51? 
etc.     The  probability  that  there  will  be  no  error  is 

-~T  >  .  .         .  .  (b) 

the  probability  that  there  will  be  no  errors  of  the  magnitude  of  a 
milligram  is 

ooou 

T-     v    •  (7) 

The  probability  that  an  error  will  lie  between  any  two  limits  x0 
and  x  is 

P  =  A[V^2^.         ...        (8) 

The  probability  that  an  error  will  lie  between  the  limits  0  and  x  is 

p  =-r\    e-iWdx,        ...        (9) 

V7T  JO 

which  expresses  the  probability  that  an  error  will  be  numerically 
less  than  x.     We  may  also  put 

p--T-r«~  ****<*(**).      *         ■         •      (10) 

VttJo 
which  is  another  way  of  writing  the  probability  integral  (8).  In 
(8),  the  limits  x0  and  x ;  and  in  (9)  and  (10),  +  x.  By  differentia- 
tion and  the  usual  method  for  obtaining  a  minimum  value  of  any 
function,  we  find,  from  (1),  that  y,  in  y  m  Jse-*2*2,  is  a  minimum 
when 

But  we  have  seen  that  the  more  accurate  the  observations  the 
greater  the  value  of  h.  The  greater  the  value  of  h,  the  smaller 
the  value  of  S(a:2) ;  3(#2)  is  a  minimum  when  h  is  a  maximum. 
This  is  nothing  but  Legend  re's  principle  of  least  squares : 
The  most  probable  value  for  the  observed  quantities  is  that  for  which 
the  sum  of  the  squares  of  the  individual  errors  is  a  minimum. 
That  is  to  say,  when 

XQ*  +  Xl     +  X22  +   •  •  •    +  Xn2  =  A  MINIMUM,  .  (11) 

where  xv  x2, . . .,  xn1  represents  the  errors  respectively  affeoting  the 
first,  second,  and  the  nth  observations. 

To  illustrate  the  reasonableness  of  the  principle  of  least  squares 
we  may  revert  to  an  old  regulation  of  the  Belgian  army  in  which 
the  individual  scores  of  the  riflemen  were  formed  by  adding  up  the 
distances  of  each  man's  shots  from  the  centre  of  the  target.     The 


518  HIGHER  MATHEMATICS.  §  160. 

smallest  sum  won  "le  grand  prix"  of  the  regiment.  It  is  not 
difficult  to  see  that  this  rule  is  faulty.  Suppose  that  one  shooter 
scored  a  1  and  a  3  ;  another  shooter  two  2's.  It  is  obvious  that 
the  latter  score  shows  better  shooting  than  the  former.  The  shots 
may  deviate  in  any  direction  without  affeoting  the  score.  Conse- 
quently, the  magnitude  of  each  deviation  is  proportional,  not  to 
the  magnitude  of  the  straight  line  drawn  from  the  place  where 
the  bullet  hits  to  the  centre  of  the  target,  but  to  the  area  of  the 
oircle  described  about  the  centre  of  the  target  with  that  line  as 
radius.  This  means  that  it  would  be  better  to  give  the  grand 
prize  to  the  score  which  had  a  minimum  sum  of  the  squares  of 
the  distances  of  the  shots  from  the  centre  of  the  target.1  This 
is  nothing  but  a  graphic  representation  of  the  principle  of  least 
squares,  formula  (11).  In  this  way,  the  two  shooters  quoted  above 
would  respectively  score  a  10  and  an  8. 

§  160.  The  Best  Representative  Value  for  a  Set  of 
Observations. 

It  is  practically  useless  to  define  an  error  as  the  deviation  of 
any  measurement  from  the  true  result,  because  that  definition 
would  imply  a  knowledge  which  is  the  object  of  investigation. 
What  then  is  an  error  ?  Before  we  can  answer  this  question,  we 
must  determine  the  most  probable  value  of  the  quantity  measured. 
The  only  available  data,  as  we  have  just  seen,  are  always  as- 
sociated with  the  inevitable  errors  of  observation.  The  measure- 
ments, in  consequence,  all  disagree  among  themselves  within 
certain  limits.  In  spite  of  this  fact,  the  investigator  is  called 
upon  to  state  definitely  what  he  considers  to  be  the  most  probable 
value  of  the  magnitude  under  investigation.  Indeed,  every  chemical 
or  physical  constant  in  our  text-boohs  is  the  best  representative  value 
of  a  more  or  less  extended  series  of  discordant  observations.  For  in- 
stance, giant  attempts  have  been  made  to  find  the  exact  length  of 
a  column  of  pure  mercury  of  one  square  millimetre  cross-sectional 
area  which  has  a  resistance  of  one  ohm  at  0°  C.  The  following 
numbers  have  been  obtained  : 


106-33 ; 

106-31 ; 

106-24 ; 

106-32  ; 

106-29 ; 

106-21 ; 

106-32  ; 

106-27 ; 

106-19, 

1  See  properties  of  similar  figures,  page 


§  160.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      519 

centimetres  (J.  D.  Everett's  Illustrations  of  the  C.G.S.  System  of 
Units,  London,  176,  1891).  There  is  no  doubt  that  the  true  value 
of  the  required  constant  lies  somewhere  between  106-19  and 
106*33  ;  but  no  reason  is  apparent  why  one  particular  value  should 
be  chosen  in  preference  to  another.  The  physicist,  however,  must 
select  one  number  from  the  infinite  number  of  possible  values  be- 
tween the  limits  106-19  and  106*33  cm. 

I.  What  is  the  best  representative  value  of  a  set  of  discordant 
results  ?  The  arithmetical  mean  naturally  suggests  itself,  and 
some  mathematicians  start  from  the  axiom  :  "  the  arithmetical 
mean  is  the  best  representative  value  of  a  series  of  discrepant 
observations  ".  Various  attempts,  based  upon  the  law  of  errors, 
have  been  made  to  show  that  the  arithmetical  mean  is  the  best 
representative  value  of  a  number  of  observations  made  under  the 
same  conditions  and  all  equally  trustworthy.  The  f)roof  rests 
upon  the  fact  that  the  positive  and  negative  deviations,  being 
equally  probable,  will  ultimately  balance  each  other  as  shown  in 
Ex.  (1),  below.1 

Examples. — (1)  If  ax,  a^, .  ,.aH  are  a  series  of  observations,   a  their 

arithmetical  mean,  show  that  the  algebraic  sum  of  the  residual  errors  is 

(«!  -  a)  +  (Oj  -  a)  + . . .  +  (a„  -  a)  -  0.  .        .      (1) 

Hint.  By  definition  of  arithmetical  mean, 

a^  +  a?  +  . .  .+  On 

a  = ;  or,  na  =  a,  +  a,+ . . .  +  an. 

n 

Distribute  the  n  a's  on  the  right-hand  side  so  as  to  get  (1),  etc. 

(2)  Prove  that  the  arithmetical  mean  makes  the  sum  of  the  squares  of 
the  errors  a  minimum.     Hint.  See  page  550. 

En  passant,  notice  that  in  calculating  the  mean  of  a  number 
of  observations  which  agree  to  a  certain  number  of  digits,  it  is  not 
necessary  to  perform  the  whole  of  the  addition.     For  example,  the 
mean  of  the  above  nine  measurements  is  written 
106  +  £(*33  +  -32  +  -32  +  -31  +  *29  +  *27  +  *24  +  -21  +  49)  - 106*276. 

II.  The  best  representative  value  of  a  constant  interval.     When 

1  G.  Hinrichs,  in  his  The  Absolute  Atomic  Weights  of  the  Chemical  Elements, 
criticizes  the  selection  (and  the  selectors)  of  the  arithmetical  mean  as  the  best  re- 
presentative value  of  a  set  of  discordant  observations.  He  asks  :  "  If  we  cannot  use 
the  arithmetical  mean  of  a  large  number  of  simple  weighings  of  actual  shillings  as  the 
true  value  of  a  (new)  shilling,  how  dare  we  assume  that  the  mean  value  of  a  few 
determinations  of  the  atomic  weight  of  a  chemical  element  will  give  us  its  true  value  ? " 
But  there  seems  to  be  a  misunderstanding  somewhere.  F.  Y.  Edgeworth  has  "The 
Choice  of  Means,"  Phil.  Mag.,  [5],  24,  268,  1887,  and  several  articles  on  related 
subjects  in  the  same  journal  between  1883  and  1889. 


520  HIGHER  MATHEMATICS.  §  160. 

the  best  representative  value  of  a  constant  interval  x  in  the  ex- 
pression xn=  x0  +  nx  (where  n  is  a  positive  integer  1,  2 .  .  .),  is  to 
be  determined  from  a  series  of  measurements  x0,  xv  x2, . .  .,  such 
that 

xx  m  x0  +  x;  x2  -  x0  +  2x ;  . . .  xn  -  xQ  +  nx, 
where  xQ  denotes  the  first  observation,  xx  the  second  reading  when 
n  —  1 ;  x2,  the  third  reading  when  n  =  2 ; . . .  The  best  value  for 
the  constant  interval  x  has  to  be  computed.     Obviously, 

x  -  xx  -  x0 ;  x  =  x2  -  xx ;  . .  .  x  =  xn  -  xn„  v 
The  arithmetical  mean  cannot  be  employed  because  it  reduces  to 

n 
the  same  as  if  the  first  and  last  term  had  alone  been  measured. 
In  such  cases  it  is  usual  to  refer  the  results  to  the  expression 

x  =  «(*  ~  ^fo  "  xi)  +  (n  ~  3)(g»-i  ~  x*)  +  '  •  •         /en 

^(7l2  -  1)  »  ^ 

which  has  been  obtained  from  the  last  of  equations  (4),  page  327, 
by  putting 

2(#)  =  :%)  =  l  +  2+  ...  +w  =  ^w(w  +  l); 
2(#2)  =  2(rc2)  =  l2  +  22+  ...  +n2  =  ^n(n  +  l)  (2n  +  l); 
%)  =  5(a?n)=a;1+ic2+  .,.*„;  5(^)-S(wa?B)»a?1  +  2aj2+...  +najn. 
If  w  is  odd,  the  middle  measurement  does  not  come  in  at  all. 
It  is  therefore  advisable  to  make  an  even  number  of  observations. 
Such  measurements  might  occur  in  finding  the  length  of  a  rod  at 
different  temperatures  ;  the  oscillations  of  a  galvanometer  needle  ; 
the  interval  between  the  dust  figures  in  Kundt's  method  for  the 
velocity  of  sound  in  gases  ;  the  influence  of  0H2  on  the  physical  and 
chemical  properties  of  homologous  series  of  organic  chemistry,  etc. 

Examples. — .(1)  In  a  Kundt's  experiment  for  the  ratio  of  the  specific 
heats  of  a  gas,  the  dust  figures  were  recorded  in  the  laboratory  notebook  at 
30-7,  43-1,  55-6,  67*9,  80-1,  92-3,  104*6,  116*9,  129-2,  141*7,  154*0,  166-1  centi- 
metres. What  is  the  best  representative  value  for  the  distance  between  the 
nodes  ?    Ansr.  12*3  cm. 

(2)  The  following  numbers  were  obtained  for  the  time  of  vibration,  in 
seconds,  of  the  "  magnet  bar  "  in  Gauss  and  Weber's  magnetometer  in  some 
experiments  on  terrestrial  magnetism  :  3*25  ;  9*90 ;  16*65 ;  23*35 ;  30*00 ;  36*65  ; 
43*30 ;  50-00 ;  56*70  ;  63*30  ;  69-80  ;  76-55 ;  83-30  ;  89*90  ;  96*65 ;  103-15  ;  109-80 ; 
116-65  ;  123-25  ;  129*95  ;  136-70 ;  143-35.  Show  that  the  period  of  vibration  is 
6-707  seconds. 

(3)  An  alternative  method  not  dependent  upon  "  least  squares  "  is  shown 
in  the  following  example  :  a  swinging  galvanometer  needle  "  reversed  "  at  (a) 


§  16,1.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      521 

10  min.  9-90  seo. ;  (6)  10  min.  23-20  sec. ;  (c)  10  min.  36-45  sec.  ;  (d)  10  min. 
49-80  sec. ;  (e)  11  min.  3*25  sec. ;  (/)  11  min.  16*60  sec,  required  the  period  of 
oscillation.  Subtract  (a)  from  (d)  and  divide  the  result  by  3.  We  get  13*300  ; 
subtract  (b)  from  (e).  We  get  13*350 ;  similarly,  from  (c)  and  (/)  we  get  13*383. 
Mean  =  18*344  =  period  of  oscillation. 


§  161.  The  Probable  Error. 

Some  observations  deviate  so  little  from  the  mean  that  we  may 
consider  that  value  to  be  a  very  close  approximation  to  the  truth,  in 
other  cases  the  arithmetical  mean  is  worth  very  little.  The  question, 
therefore,  to  be  settled  is,  What  degree  of  confidence  may  we  have  in 
selecting  this  mean  as  the  best  representative  value  of  a  series  of  ob- 
servations ?    In  other  words,  How  good  or  how  bad  are  the  results  ? 

We  could  employ  Gauss'  absolute  measure  of  precision  to  answer 
this  question.  It  is  easy  to  show  that  the  measure  of  precision  of 
two  series  of  observations  is  inversely  as  their  accuracy.  If  the 
probabilities  of  an  error  xv  lying  between  0  and  lv  and  of  an  error 
x2t  between  0  and  lt,  are  respectively 

Pl  *  ^\he~h^d^^)  i  p*  -  jfcl  «-*tVi(V^. 

it  is  evident  that  when  the  observations  are  worth  an  equal  degree 
of  confidence,  Px  —  l\. 

.*.  l^h\  **   ^2^2  »  or»  h  '  *^2  **  ni''n\i 

or  the  measure  of  precision  of  two  series  of  observations  is  in- 
versely as  their  accuracy.  An  error  lx  will  have  the  same  degree 
of  probability  as  an  error  l2  when  the  measure  of  precision  of  the 
two  series  of  observations  is  the  same.  For  instance  if  hx  =  4ch2, 
Px  m  P2  when  l2  =  4^,  or  four  times  the  error  will  be  committed  in 
the  second  series  with  the  same  degree  of  probability  as  the  single 
error  in  the  first  set.  In  other  words,  the  first  series  of  obser- 
vations will  be  four  times  as  accurate  as  the  second.  On  account 
of  certain  difficulties -in  the  application  of  this  criterion,  its  use  is 
mainly  confined  to  theoretical  discussions. 

One  way  of  showing  how  nearly  the  arithmetical  mean  repre- 
sents all  the  observations,  is  to  suppose  all  the  errors  arranged 
in  their  order  of  magnitude,  irrespective  of  sign,  and  to  select  a 
quantity  whioh  will  occupy  a  place  midway  between  the  extreme 
limits,  so  that  the  number  of  errors  less  than  the  assumed  error  is 
the  same  as  those  which  exceed  it.     This  is  called  the  probable 


522 


HIGHER  MATHEMATICS. 


§161. 


error,  not  "the  most  probable   error,"  nor  "the  most  probable 
yalue  of  the  actual  error  ". 

The  probable  error  determines  the  degree  of  confidence  we  may 
have  in  using  the  mean  as  the  best  representative  value  of  a  series 
of  observations.  For  instance,  the  atomic  weight  of  oxygen  (H  =  l) 
is  said  to  be  15*879  with  a  probable  error  +  0"0003.  This  means 
that  the  arithmetical  mean  of  a  series  of  observations  is  15*879, 
and  the  probability  is  | — that  is,  the  odds  are  even,  or  you  may  bet 
£1  against  £1 — that  the  true  atomic  weight  of  oxygen  lies  between 
15*8793  and  15-8787. 

Eeferring  to  Fig.  171,  let  MP  and  M'P'  be  drawn  at  equal  dis- 
tances from  Oy  in  such  a  way  that  the  area  bounded  by  these 
lines,  the  curve,  and  the  #-axis  (shaded  part  in  the  figure),  is  equal 
to  half  the  whole  area,  bounded  by  the  whole  curve  and  the  #-axis, 
then  it  will  be  obvious  that  half  the  total  observations  will  have 
errors  numerically  less  than  OM\  and  half,  numerically  greater 

than  OM,  that  is,  OM  re- 
presents the  magnitude  of 
the  probable  error,  MP  its 
probability. 

The  way  some  investi- 
gators refer  to  the  smallness 
of  the  probable  error  affect- 
ing their  results  conveys  the 
impression  that  this  canon 
has  been  employed  to  em- 
phasize the  accuracy  of  the  work.  As  a  matter  of  fact,  the  probable 
error  does  not  refer  to  the  accuracy  of  the  work  nor  to  the  mag- 
nitude of  the  errors,  but  only  to  the  proportion  in  which  the  errors 
of  different  magnitudes  occur.  A  series  of  measurements  affected 
with  a  large  probable  error  may  be  more  accurate  than  another 
series  with  a  small  probable  error,  because  the  second  set  may  be 
affected  with  a  large  constant  error  (q.v.). 

The  number  of  errors  greater  than  the  probable  error  is  equal 
to  the  number  of  errors  less  than  it.  Any  error  selected  at  ran- 
dom is  just  as  likely  to  be  greater  as  less  than  the  probable  error. 
Hence,  the  probable  error  is  the  value  of  x  in  the  integral 


r.^:*** 


a) 


page  517.     From  Table  X.,  page  621,  when  P  =  J,  hx  =  04769 ; 


y  =  -j=e-»2** (3) 


§  161.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      523 

or,  if  r  is  the  probable  error, 

hr  =  0-4769 (2) 

Now  it  has  already  been  shown  that 

hdx 

From  page  500,  therefore,  the  probability  of  the  occurrence  of  the 
independent  errors,  xv  x2,  .  .  .,  xn  is  the  product  of  their  separate 
probabilities,  or 

*_*»«-««*,        ...        (4) 

For  any  set  of  observations  in  which  the  measurements  have  been 
as  accurate  as  possible,  h  has  a  maximum'  value.  Differentiating 
the  last  equation  in  the  usual  way,  and  equating  dP/dh  to  zero, 


W^r  •     •     •         w 


Substitute  this  in  (2) 

r  «±  0-6745  ^p-\        ,        .  (6) 

But  2(rc2)  is  the  sum  of  the  squares  of  the  true  errors.  The  true 
errors  are  unknown.  By  the  principle  of  least  squares,  when 
measurements  have  an  equal  degree  of  confidence,  the  most  prob- 
able value  of  the  observed  quantities  are  those  which  render  the 
sum  of  the  squares  of  the  deviations  of  each  observation  from  the 
mean,  a  minimum.  Let  2(va)  denote  the  sum  of  the  squares  of 
the  deviations  of  each  observation  from  the  mean.  If  n  is  large, 
we  may  put  2(a?2)  =  2(t>2) ;  but  if  n  is  a  limited  number, 

%{v2)  <  2(*2) ;  /.  %{x2)  =  2(^2)  +  uK  .         (7) 

All  we  know  about  u2  is  that  its  value  decreases  as  n  increases, 
and  increases  when  %{x2)  increases.  It  is  generally  supposed  that 
the  best  approximation  is  obtained  by  writing 

n     '  ' '      n        n  -  Y 
Hence,  the  probable  error,  r,  of  a  single' observation  is 


\  n  -  V 


r  «  ±  0-6745  a/     _  v,  .        single  observation  (8) 

which  is  virtually  Bessel's  formula,  for  the  probable  error  of  a 
single  observation.  $(v2)  denotes  the  sum  of  the  squares  of  the 
numbers    formed    by    subtracting   each    measurement    from    the 


524  HIGHER  MATHEMATICS.  §  162. 

arithmetical  mean  of  the  whole  series,  n  denotes  the  number  of 
measurements  actually  taken.  The  probable  error,  R,  of  the  arith- 
metical mean  of  the  whole  series  of  observations  is 

,  1N.         .  .    ALL  OBSERVATIONS    (9) 

n\n  -  1)  v  ' 

The  derivation  of  this  formula  is  given  as  Ex.  (2),  page  531. 
The  last  two  results  show  that  the  probable  error  is  diminished 
by  increasing  the  number  of  observations.  (8)  and  (9)  are  only 
approximations.  They  have  no  signification  when  the  number  of 
observations  is  small.  Hence  we  may  write  §  instead  of  0*674:5. 
For  numerical  applications,  see  next  section. 

The  great  labour  involved  in  the  squaring  of  the  residual  errors 
of  a  large  number  of  observations  may  be  avoided  by  the  use  of 
Peter's  approximation  formula.  According  to  this,  the  prob- 
able error,  r,  of  a  single  observation  is 

r  =  +  0-8453    ,  ,        .:,    .       single  observation  (10) 
Jn{n  -  1)  v    ' 

where  25(  +  v)  denotes  the  sum  of  the  deviations  of  every  observa-. 
tion  from  the  mean,  their  sign  being  disregarded.  The  probable 
error,  R,  of  the  arithmetical  mean  of  the  whole  series  of  observa- 
tions is 

S(+  v) 

R  =  +  0-8453^     } ==-.      .  ALL  OBSERVATIONS    (11) 

njn  -  1  v    ' 

Tables  VI.  to  IX.,  pages  619  to  623,  will  reduce  the  labour  in 
numerical  calculations  with  Bessel's  and  with  Peter's  formulas. 


§  162.  Mean  and  Average  Errors. 

The  arbitrary  choice  of  the  probable  error  for  comparing  the 
errors  which  are  committed  with  equal  facility  in  different  sets  of 
observations,  appears  most  natural  because  the  probable  error 
occupies  the  middle  place  in  a  series  arranged  according  to  order 
of  magnitude  so  that  the  number  of  errors  less  than  the  fictitious 
probable  error,  is  the  same  as  those  which  exceed  it.  There  are 
other  standards  of  comparison.  In  Germany,  the  favourite  method 
is  to  employ  the  mean  error,  which  is  defined  as  the  error  whose 
square  is  the  mean  of  the  squares  of  all  the  errors,  or  the  u  error 
which,  if  it  alone  were  assumed  in  all  the  observations  indifferently, 
would  give  the  same  sum  of  the  squares  of  the  errors  as  that  which 


§  162.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      525 

actually  exists  ".     We  have  seen,  on  page  516,  (5),  that  the  ratio, 

Number  of  errors  between  x  and  x  +  dx  _    h     ,  ^2  , 
Total  number  of  errors  ~  'J^e  '  *  dx' 

Multiply  both  sides  by  #2and  we  obtain 

Sum  of  squares  of  errors  between  x  and  x  +  dx  _    h  2 

Total  number  of  errors  ~  J~x  e?l  *  ",x* 

By  integrating  between  the  limits  +  oo  and  -  oo  we  get 

Sum  of  squares  of  all  the  errors       5(#2)  _     h    f+GC  2   _s2l2. 
Total  number  of  errors  =  ~^~  ~  lJZ)-°?  & 

Let  m  denote  the  mean  error,  then,  by  integration  as  on  page  343, 

and  from  (2)  of  the  preceding  section,  we  get 

r  =  06745m.  ...         (2) 

From  (8)  and  (9),  preceding  section,  the   mean  error,  m,  which 
affects  each  single  observation  is  given  by  the  expression    . 

m  =±  A/w  _  i  ;        •  •  SINGLE  OBSERVATION   (3) 

and  the  mean  error,  M,  which  affects  the  whole  series  of  results, 

/      _  jy  .  .       ALL  OBSERVATIONS   (4) 

The  mean  error  must  not  be  confused  with  the  "  mean  of  the 
errors,"  or,  as  it  is  sometimes  called,  the  average  error,1  another 
standard  of  comparison  denned  as  the  mean  of  all  the  errors  re- 
gardless of  sign.  If  a  denotes  the  average  error,  we  get  from 
page  235, 

a  -  %+$  ^f"w-*«(to.si7= .  r  ,  0-8454a.        (5) 

The  average  error  measures  the  average  deviation  of  each 
observation  from  the  mean  of  the  whole  series.  It  is  a  more 
useful  standard  of  comparison  than  the  probable  error  when  the 
attention  is  directed  to  the  relative  accuracy  of  the  individual 
observations  in  different  series  of  observations.  The  average  error 
depends  not  only  upon  the  'proportion  in  which  the  errors  of  differ- 

1  Some  writers  call  our  "  average  error  "  the  "  mean  error,"  and  our  "  mean  error  " 
the  "  error  of  mean  square  ". 


526 


HIGHER  MATHEMATICS. 


§162. 


ent  magnitudes  occur,  but  also  on  the  magnitude  of  the  individual 
errors.  The  average  error  furnishes  useful  information  even  when 
the  presence  of  (unknown)  constant  errors  renders  a  further  appli- 
cation of  the  "theory  of  errors  "  of  questionable  utility,  because  it 
will  allow  us  to  compare  the  magnitude  of  the  constant  errors 
affecting  different  series  of  observations,  and  so  lead  to  their  dis- 
covery and  elimination. 

The  reader  will  be  able  to  show  presently  that  the  average  error, 
A,  affecting  the  mean  of  n  observations  is  given  by  the  expression 

a  —  +  /  — . 

rts/n 


(6) 


This  determines  the  effect  of  the  average  error  of  the  individual 
observations  upon  the  mean,  and  serves  as  a  standard  for  comparing 
the  relative  accuracy  of  the  means  of  different  series  of  experiments 
made  under  similar  conditions. 

Examples. — (1)  The  following  galvanometer  deflections  were  obtained  in 
some  observations  on  the  resistance  of  a  circuit :  37*0,  36-8,  36*8,  36*9,  37*1. 
Find  the  probable  and  mean  errors.  This  small  number  of  observations  is 
employed  simply  to  illustrate  the  method  of  using  the  above  formulae.  In 
practical  work,  mean  or  probable  errors  deduced  from  so  small  a  number  of 
observations  are  of  little  value.    Arrange  the  following  table  : — 


Number  of 

Deflection 

Departure  from 

A 

Observation. 

Observed. 

Mean. 

1 

37*0 

+  0-08 

0-0064 

2 

36-8 

-0-12 

0-0144 

3 

36-8 

-0-12 

0-0144 

4 

36*9 

-0-02 

0-0004 

5 

37-1 

+  0-18 

0-0324 

Mean  =  36-92 ;  2(v2)  m  0-0680. 

The  numbers  in  the  last  two  columns  have  been  oaloulated  from  those  in 
the  second.    Sinoe  n  =»  5,  and  writing  $  for  0-6745. 

Mean  error  of  a  single  result       =  +  v0-^  =»  +  0-13. 


Mean  error  of  the  mean 
Probable  error  of  a  single  result 
Probable  error  of  the  mean 


=  ±  \/°-?r  -  ±  0*058. 

—  6.4  — 


+  -I  n/°J£?  «  +  0-087. 


=  +  2  J^f  =  +  0-039. 


Average  error  of  a  single  result    =  +  ^  m  ±  0*104. 


Average  error  of  the  mean 


+  «"   .  +  0-0465. 

~  6  *Jo        — 


§162.    PKOBABILITY  AND  THE  THEORY  OF  ERRORS.      527 

The  mean  error  of  the  arithmetical  mean  of  the  whole  set  of  observations  is 
written,  36-92  +  0*058  ;  the  probable  error,  36*92  ±  0*039.  It  is  unnecessary 
to  include  more  than  two  significant  figures.  You  will  find  the  Tables  on 
pages  619  and  620  convenient  for  the  numerical  work. 

(2)  F.  Rudberg  (Pogg.  Ann.,  41, 271, 1837),  found  the  coefficient  of  expansion 
a  of  dry  air  by  different  methods  to  be  a  x  100  =  0*3643,  0*3654,  0*3644,  0*3650, 
0*3653,  0*3636,  0*3651,  0*3643,  0*3643,  0*3645,  0*3646,  0*3662,  0*3840,  0*3902, 
0*3652.  Required  the  probable  and  mean  errors  on  the  assumption  that  the 
results  are  worth  an  equal  degree  of  confidence. 

(3)  From  Ex.  (3),  page  161,  show  that  the  mean  error  is  the  abscissa  of 
the  point  of  inflexion  of  the  probability  curve.    For  simplicity,  put  h  =e  1. 

(4)  Cavendish  has  published  the  result  of  29  determinations  of  the  mean 
density  of  the  earth  (Phil.  Trans.,  88,  469,  1798)  in  which  the  first  significant 
figure  of  all  but  one  is  5 :— 4*88 ;  5-50 ;  5*61 ;  5*07 ;  5*26 ;  5*55 ;  5*36 ;  5*29  ; 
5*58  ;  5*65  ;  5*57  ;  5*53 ;  5*62  ;  5*29  ;  5-44 ;  5*34  ;  5*79  ;  510 ;  5-17  ;  5-39  ;  5-42 ; 
5*47  ;  5*63  ;  5*34  ;  5*46  ;  5*30  ;  5*75  ;  5*68  ;  5*85.  Verify  the  following 
results:  Mean=5*45;  2(  +  v)  =  5*04;  S(*>2) -1-367  ;  M=  ±0041;  m=  ±0*221; 
£«=  ±0*0277;  r  =  ±0*149;   a= 0*  18  ;  4=  ±0*038. 

The  relation  between  the  probable  error,  the  mean  error,  the 
average  error,  and  the  absolute  measure  of  an  error  can  be  ob- 
tained from  (2),  page  523 ;  (2),  page  516;  and  (5),  page  516.  We 
have,  in  fact,  if  modulus,  h  =  1*0000;  mean  error,  m  =»  0*7071 ; 
average  error,  a  —  0*5642 ;  probable  error,  r  =  0*4769. 

The  following  results  are  convenient  in  combining  measurements 
affected  with  different  mean  or  probable  errors  : 

I.  The  mean  error  of  the  sum  or  difference  of  a  number  of 
observations  is  equal  to  the  square  root  of  the  sum  of  the  squares  of 
the  mean  errors  of  each  of  the  observations.  Let  xv  x2,  represent 
two  independent  measurements  whose  sum  or  difference  combines 
to  make  a  final  result  X,  so  that 

X  =  x1  +  xv 

Let  the  mean  errors  of  xx  and  x2,  be  mx  and  m2  respectively.  If 
M  denotes  the  mean  error  in  X, 

X  ±  M  =  (x1  ±  mx)  +  (x2  ±  m2). 

.*.  ±  M  =  ±  mY  ±  m2. 

However  we  arrange  the  signs  of  M,  mv  m2>  in  the  last  equation, 
we  can  only  obtain,  by  squaring,  one  or  other  of  the  following  ex- 
pressions : — 

M?  =  mx2  +  2mlm2  +  m22 ;  or,  .M2  =  m^  -  2m1ra2  +  m22, 

it  makes  no  difference  which.     Hence  the  mean  error  is  to  be  found 


528  HIGHER  MATHEMATICS.  §  162. 

by  taking  the  mean  of  both  these  results.     That  is  to  say, 

M2  =  ?V  +  m22 ;  or,  M  =  +  Jm^  +  ra22, 
because  the  terms  containing  +  m^m2  and  -  m^m^  cancel  each  other. 
This  means  that  the  products  of  any  pair  of  residual  errors  (m1m2f 
m^mz, . .  .)  in  an  extended  series  of  observations  will  have  positive 
as  often  as  negative  signs.  Consequently,  the  influence  of  these 
terms  on  the  mean  value  will  be  negligibly  small  in  comparison  with 
the  terms  m^,  w22,  m32,  .  .  .,  which  are  always  positive.  Hence,  for 
any  number  of  observations, 

W  =  m*  +  m*  +  .  .  .  ;  or ,M  =  +  ,/W  +  m22  +  .  .  .).  (7) 
From  equation  (2),  page  525,  the  mean  error  is  proportional  to  the 
probable  error  B,  m1  to  rv  .  .  .,  hence, 

-B2  =  V  +  r22  +  .  .  . ;  or  ,B  =  ±  J  fa*  +  r*  +  .  .  .).  (8) 
In  other  words,  the  'probable  error  of  the  sum  or  difference  of  two 
quantities  A  and  B  respectively  affected  with  probable  errors  ±  a 

and  ±  b  is  

B  =  ±  J  a*  +  b\  .         .         .       (9) 

Examples. — (1)  The  moleoular  weight  of  titanium  chloride  (TiClJ  is  known 
to  be  188*545  with  a  probable  error  +  0*0092,  and  the  atomic  weight  of  chlorine 
35-179  +  0-0048,  what  is  the  atomic  weight  of  titanium?    Ansr.  47*829  ± 
0*0213.     Hints. 
188-545  -  4  x  35-179  =  47*829 ;  B  =  ^(0*0092)2  +  (4  x  0*0048)2  =  ±  -0213. 

It  will  be  obvious  that  we  shall  ignore  the  advice  given  in  §  94,  pages  273 
to  276,  if  we  are  not  very  careful  in  the  interpretation  of  the  probable  error  in 
these  illustrative  examples. 

(2)  The  mean  errors  affecting  6X  and  02  in  the  formula  B  =  fe(02  -  6})  are 
respectively  +  0*0003  and  +  0*0004,  what  is  the  mean  error  affecting  02  -  0, 
and  3(02  -  0J  ?    Ansr.  +  0*0005  and  ±  00015. 

II.  The  probable  error  of  the  product  of  two  quantities  A  and 
B  respectively  affected  with  the  probable  errors  ±  a  and  ±b  is 

B  =  ±  J{Aby  +  (Baf.  .        .       (10) 

If  a  third  mean,  G,  with  a  probable  error,  +  c,  is  included, 

B  =  +  J(BCay+  (ACb)*  +  {ABcf.  .        .      (11) 

Examples. — (1)  Thorpe  found  that  the  molecular  ratio 
4Ag :  TiCl<  m  100 :  44-017  ±  00031. 
Henoe  determine  the  molecular  weight  of  titanium  tetrachloride,  given  the 
atomic  weight  of  silver  »  107*108  ±  0*0031.    Ansr.  188*583  ±  0*0144.    Hint. 
&  -  ±  n/{(4  x  107-108  x  0-0031)8  +  (44-017  x  4  x  0*0031)*}. 
(2)  The  specific  heat  of  tin  is  0*0537  with  a  mean  error  of  +  0*0014,  and 
the  atomic  weight  of  the  same  metal  is  118'150  +  0*0089,  show  that  the  mean 
error  of  the  product  of  these  two  quantities  (Dulong  and  Petit's  law)  is  6*34  + 
0-1654. 


§  162.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.       529 

III.  The  probable  error  of  the  quotient  (B  -t-  A)  of  two  quantities 
A  and  B  respectively  affected  with  the  probable  errors  ±  a  and  ±bis 

•  ■l^S ,  .    ~m 

Examples. — (1)  It  is  known  that  the  atomic  ratio 

Cu  :  2Ag  =  100 :  339-411  ±  0-0089, 
what  is  the  atomic  weight   of  copper  given   Ag  =  107*108  +  0-0031  ?     Ansr. 
63-114  +  0-0020.     Hint. 


R  =x  ±  J/214-216x0-0039y+  (q.qq^  ~  339.4!!  m  ±  0-0020. 

Ou :  2  x  107-108  -=  100 :  339-411 ;  .-.  Cu  =  63*114. 
(2)  Suppose  that  the  maximum  pressure  of  the  aqueous  vapour,  f.v  in  the 
atmosphere  at  16°  is  found  to  be  8-2000,  with  a  mean  error  +  0*0024,  and  the 
maximum  pressure  of  aqueous  vapour,  /3,  at  the  dewpoint,  at  16°,  is  13-5000, 
with  a  mean  error  of  +  0-0012.  The  relative  humidity,  h,  of  the  air  is  given 
by  the  expression  h  =x  /j//2.  Show  that  the  relative  humidity  at  16°  is 
0-6074  +  0-0O02. 

IV.  The  probable  error  of  the  pboportion 
A  :  B  =  C  :  x, 
where  A,  B,  C,  are  quantities  respectively  affected  with  the  probable 
errors  ±  a,  ±  b,  ±  c,  is 


ff-±VV  A  J .        .      (13) 

Example.— Stas  found  that  AgCIO,  furnished  25-080  ±  0-0019  %  of 
oxygen  and  74*920  ±  0-0003  °/0  of  AgCl.  If  the  atomic  weight  of  oxygen  is 
15-879  ±  0-0003,  what  is  the  molecular  weight  of  AgCl  ?  Ansr.  142-303  ±  0-0066. 
Hints.  25-080 :  74-920  =  3  x  15879  :  x ;  .-.  x  =  142-303. 

/f/74-92x  47-637  x  0-001  \     ,  ^| 

*    y  j  y  25.Q8  J  +  (47-637  x  0-001)a  +  (74-92  x  3  x  0-0009)a  [ 

Bs=± V^'08  ' 

If  the  proportion  be 

A\B  =  G  +  x:D  +  x, 

the  probable  error  is  given  by 


■r.'iOSS-i^ 


(14) 


Example. — Stas  found  that  31-488  +  0-0006  grams  of  NH4G1  were  equiva- 
lent to  100  grams  of  AgNOs.  Hence  determine  the  atomic  weight  of  nitrogen, 
given  Ag=107-108  ±0-0031 ;  01  =  35-179  ±  0-0048;  H-l;  03  =  47-687  ±  0-0009. 
Ansr.  13*911  ±  0-0048. 

LL 


530  HIGHER  MATHEMATICS.  §  162. 

V.  The  probable  error  of  the  arithmetical  mean  of  a  series  of 
observations  is  inversely  as  the  square  root  of  their  number.  Let 
rv  r2,  .  .  .,  rn  be  the  probable  errors  of  a  series  of  independent 
observations  av  a^ . .  .,  an,  which  have  to  be  combined  so  as  to 
make  up  a  final  result  u.  Let  the  probable  errors  be  respectively 
proportional  to  the  actual  errors  dav  da2, . . .  da„.  The  final  result 
u  is  a  function  such  that 

u  ■  f(av  av  •  •  •>  an)- 
The  influence  of  each  separate  variable  on  the  final  result  may  be 
determined  by  partial  differentiation  so  that 

,         ~du  _  ~du  _  "^ 

** =  ss^  +  ^da*+ (15> 

where  dav  da2,  . . .  represent  the  actual  errors  committed  in 
measuring  av  a2, .  .  . ;  the  partial  differential  coefficients  determine 
the  effect  of  these  variables  upon  the  final  result  u\  and  du  re- 
presents the  actual  error  in  u  due  to  the  joint  occurrence  of  the 
errors  dav  da2, , . .  Put  B  in  place  of  du;  rl  in  place  of  dav  etc. ; 
square  (15)  and 

^-®^+®v+--  •  •  <«> 

since  cross  products  are  negligibly  small.  The  arithmetical  mean 
of  n  observations  is 


therefore, 

<)&!      'ba2       ' "      n '   '  '  n2 

But  the  observations  have  an  equal  degree  of  precision,  and  there- 
fore, r,2  =  r22  = . . .  =  rn2  =  r2. 

->*-  Wt^       •    •    (17) 

This  result  shows  how  easy  it  is  to  overrate  the  effect  of  multi- 
plying observations.  If  B  denotes  the  probable  error  of  the  mean 
of  8  observations,  four  times  as  many,  or  32  observations  must  be 
made  to  give  a  probable  error  of  ^B ;  nine  times  as  many,  or  72 
observations  must  be  made  to  reduce  B  to  \B,  etc. 

Examples. — (1)  Two  series  of  determinations  of  the  atomic  weight  of  oxygen 
by  a  certain  process  gave  respectively  15-8726  ±0-00058  and  15-8769  ±0-00058. 
Hence  show  that  the  atomic  weight  is  accordingly  written  15-87475 ±0-00041. 


§  163.  PROBABILITY  AND  THE  THEORY  OF  ERRORS.       531 

(2)  In  the  preceding  section,  §  161,  given  formula  (8)  deduce  (9).  Hint. 
Use  (17),  present  section. 

(3)  Deduce  Peter's  approximation  formulae  (10)  and  (11),  §  161.  Hint. 
Since  5(a")/n  =  2{v2)/(n  -  1),  page  524,  we  may  suppose  that  on  the  average 
5(<c):  sjn  =  2(v) :  sin  -  1,  etc.  2xjn  is  the  mean  of  the  errors,  and  if  2,x/n= 
probability  integral,  page  522,  =  1/&  \f*,  it  follows  from  (2),  page  523,  r  = 
0-8453  Sx/n,  etc.     See  also  (2)  page  516. 

(4)  Show  that  when  n  is  large,  the  result  of  dividing  the  mean  of  the 
squares  of  the  errors  by  the  square  of  the  mean  of  the  errors  is  constant. 
Hint.     Show  that 

*?♦(*>)■-*-»•<•   •         •     •    w> 

This  has  been  proposed  as  a  test  of  the  fidelity  of  the  observations,  and  of  the 
accuracy  of  the  arithmetical  work.  For  instance,  the  numbers  quoted  in  the 
example  on  page  554  give  2(u)=  55-53;  2(i>2)=  354*35;  n=14;  constant =1-60. 
The  canon  does  not  usually  work  very  well  with  a  small  number  of  observa- 
tions. 

(5)  Show  that  the  probable  (or  mean)  error  is  inversely  proportional  to 
the  absolute  measure  of  precision.     Hint.  From  (1)  and  (2) 

r  =  r-  x  constant;  .-. roc  ^  ,  (19) 

§  163.  Numerical  Values  of  the  Probability  Integrals. 

We  have  discussed  the  two  questions : 

1.  What  is  the  best  representative  value  of  a  series  of  measure- 
ments affected  with  errors  of  observations  ? 

2.  How  nearly  does  the  arithmetical  mean  represent  all  of  a 
given  set  of  measurements  affected  with  errors  of  observation  ? 

It  now  remains  to  inquire 

3.  How  closely  does  the  arithmetical  mean  approximate  to  the 
absolute  truth  ?  To  illustrate,  we  may  use  the  results  of  Crookes' 
model  research  on  the  atomic  weight  of  thallium  (Phil.  Trans.,  163, 
277,  1874).     Crooke's  determination  of  this  constant  gave 

203-628;  203-632;  203-636;  203-638;  203-639  1„  ™«  „,* 

■Mean:  203-642. 


\ 


203-642;  203-644;  203-649;  203-650;  203-666 
The  arithmetical  mean  is  only  one  of  an  infinite  number  of  possible 
values  of  the  atomic  weight  of  thallium  between  the  extreme  limits 
203-628  and  203-666.  It  is  very  probable  that  203-642  is  not  the 
true  value,  but  it  is  also  very  probable  that  203-642  is  very  near 
to  the  true  value  sought.  The  question  " How  near?"  cannot  be 
answered.  Alter  the  question  to  "  What  is  the  probability  that 
the  truth  is  comprised  between  the  limits  203*642  +  x?  ".  and  the 
answer  may  be  readily  obtained  however  small  we  choose  to  make 
the  number  x. 

LL  * 


532  HIGHER  MATHEMATICS.  §  163. 

First,  suppose  that  the  absolute  measure  of  precision,  h,  of  the 

arithmetical  mean  is  known.     Table  X.  gives  the  numerical  values 

of  the  probability  integral 

9    f** 
P=  -7=]   e~Md(hx), 

S/TT  JO 

where  P  denotes  the  probability  that  an  error  of  observation  will 
have  a  positive  or  negative  value  equal  to  or  less  than  x,  h  is  the 
measure  of  the  degree  of  precision  of  the  results. 

When  h  is  unity,  the  value  of  P  is  read  off  from  the  table 
directly.  To  illustrate,  we  read  that  when  x  —  +  0*1  P  =  '112 ; 
when  x  =  +  0*2  P  =  -223 ;  .  .  .,  meaning  that  if  1,000  errors  are 
committed  in  a  set  of  observations  with  a  modulus  of  precision 
h  =  1,  112  of  the  errors  will  lie  between  +  O'l  and  -  O'l,  223 
between  +  0'2  and  -  0*2,  etc.  Or,  888  of  the  errors  will  exceed 
the  limits  ±  0*1 ;  777  errors  will  exceed  the  limits  +  0*2 ;  .  .  .  When 

01  0-2 
h  is  not  unity,  we  must  use  -r-   ~j~,  .  .  .,  in  place  of  O'l,  0'2. 

Examples.— (1)  If  hx  =  0-64,  P,  from  the  table,  is  0-6346.  Hence  0*6346 
denotes  the  probability  that  the  error  x  will  be  less  than  0'64/ft,  that  is  to 
say,  63-46  %  °f  *ne  errors  will  lie  between  the  limits.  +  0-64/fc.  The  remaining 
36"54  °/0  will  lie  outside  these  limits. 

(2)  Required  a  probability  that  an  error  will  be  comprised  between  the 
limits  ±  0-3  (h  =  1).     Ansr.  0-329. 

(3)  Required  the  probability  that  an  error  will  lie  between  -  0-01  and 
+  0-1  of  say  a  milligram.  This  is  the  sum  of  the  probabilities  of  the  limits 
from  0  to  -  0-01  and  from  0  to  +  0*1  (h  -  1).    Ansr.  £(0*113  +  0*1125)  =0*0619. 

(4)  Required  the  probability  that  an  error  will  lie  between  +  1*0  and  + 
0*01.  This  is  the  difference  of  the  probabilities  of  errors  between  1*0  and  zero 
and  between  0'01  and  zero  (h  =  1).     Ansr.  £(0*8427  -  0*0113)  =  0*4157. 

This  table,  therefore,  enables  us  to  find  the  relation  between  the 
magnitude  of  an  error  and  the  frequency  with  which  that  error  will 
be  committed  in  making  a  large  number  of  careful  measurements. 
It  is  usually  more  convenient  to  work  from  the  probable  error  B 
than  from  the  modulus  h.  More  practical  illustrations  have,  in 
consequence,  been  included  in  the  next  set  of  examples. 

Second,  suppose  that  the  probable  error  of  the  arithmetical  mean 
is  known.  Table  XI.  gives  the  numerical  values  of  the  probability 
integral 


*v&:ra,<?> 


§  163.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      533 

where  P  denotes  the  probability  that  an  error  of  observation  of  a 
positive  or  negative  value,  equal  to  or  less  than  x,  will  be  com- 
mitted in  the  arithmetical  mean  of  a  series  of  measurements  with 
probable  error  r  (or  B).  This  table  makes  no  reference  to  h.  To 
illustrate  its  use,  of  1,000  errors,  54  will  be  less  than  ^$B ;  500 
less  than  B ;  823  less  than  2B ;  957  less  than  3B ;  993  less  than 
4jB  ;  and  one  will  be  greater  than  5B. 

Examples. — (1)  A  series  of  results  are  represented  by  6-9  with  a  probable 
error  ±  0*25.  The  probability  that  the  probable  error  is  less  than  0-25  is  $. 
What  is  the  probability  that  the  actual  error  will  be  less  than  0-75. 
Here  x/B  =  0-75/0-25  =  3.  From  the  table,  p  =  0-9570  when  x/B  =  3. 
This  means  that  95-7  °/0  of  the  errors  will  be  less  than  0-75  and  4*3  °/» 
greater. 

(2)  D.  Gill  finds  the  solar  parallax  to  be  8-802"  ±  0*005.  What  is  the 
probability  that  the  solar  parallax  may  lie  between  8-802"  +  0-025.  Here 
x/B  =  0-025  -f  0-005  =  5.  When  B  =  5,  Table  XI.,  P  =  0-9993.  This 
means  that  £9993  might  be  bet  in  favour  of  the  event,  and  £7,  against  the 
event. 

(3)  Dumas  has  recorded  the  following  19  determinations  of  the  chemical 
equivalent  of  hydrogen  (O  =  100)  using  sulphuric  acid  (H2S04)  with  some, 
and  phosphorus  pentoxide  (P206)  as  the  drying  agent  in  other  cases  : 

i.  HaS04  :  12-472,  12-480,  12-548,  12-489,  12-496,  12-522,  12-533,  12-546, 
12-550,  12-562 ; 

ii.  P206  :  12-480,  12-491,  12-490,  12-490,  12-508,  12-547,  12-490,  12-551, 
12-551.  J.  B.  A.  Dumas'  "  Recherches  sur  la  Composition  de  l'Eau,"  Ann. 
Chim.  Phys.,  [3],  8,  200,  1843.  What  is  the  probability  that  there  will  be  an 
error  between  the  limits  +  0015  in  the  mean  (12-515),  assuming  that  the 
results  are  free  from  constant  errors  ?  The  chemical  student  will  perhaps  see 
the  relation  of  his  answer  to  Prout's  law.  Hints.  x/B  =  t ;  B  =  0*005685 ; 
x  =  0-015 ;  .'.  t  =  2-63.  From  Table  XI.,  when  t  =  2-63,  P  =  0'969.  Hence 
the  odds  are  969  to  31  that  the  mean  12*515  is  affected  by  no  greater  error 
than  is  comprised  within  the  limits  +  0*015.  To  exemplify  Table  X., 
h  =  0-4769/22  =  102, .-.  hx  =  102  x  0*015  =  1*53.  From  the  Table,  P  =  0*924 
when  hx  =  1*53,  etc.  That  is  to  say,  96*9  °/0  of  the  errors  will  be  less  and 
3*1  0/o  greater  than  the  assigned  limits. 

(4)  From  W.  Crookes'  ten  determinations  of  the  atomic  weight  of  thallium 
(above)  calculate  the  probability  that  the  atomic  weight  of  thallium  lies  be- 
tween 203-632  and  203-652.  Here  x  =  ±  0'01 ;  B  ±  =  0*0023 ;  .*.  2=a*/i*=4*4. 
From  Table  XI.,  P= 0*997.  (Note  how  near  this  number  is  to  unity  indicating 
certainty.)  The  chances  are  332  to  1  that  the  true  value  of  the  atomic  weight 
of  thallium  lies  between  203*632  and  203*652.  We  get  the  same  result  by 
means  of  Table  X.  Thus  7t=0-4769  -r  0-002 3 =207  ;  .-.  ^=207x0*01  =  2*07. 
When  &b=2*07,  P= 0*997.  If  1,000  observations  were  made  under  the  same 
conditions  as  Crookes',  we  could  reasonably  expect  997  of  them  to  be  affeoted 
by  errors  numerically  less  than  0*01,  and  only  8  observations  would  be  affected 
by  errors  exceeding  these  limits. 


534  HIGHER  MATHEMATICS.  §  164. 

The  rules  and  formulae  deduced  up  to  the  present  are  by  no 
means  inviolable.  The  reader  must  constantly  bear  in  mind  the 
fundamental  assumptions  upon  which  we  are  working.  If  these 
conditions  are  not  fulfilled,  the  conclusions  may  not  only  be  super- 
fluous, but  even  erroneous.     The  necessary  conditions  are  : 

1.  Every  observation  is  as  likely  to  be  in  error  as  every  other 
one. 

2.  There  is  no  perturbing  influence  to  cause  the  results  to  have 
a  bias  or  tendency  to  deviate  more  in  some  directions  than  in 
others. 

3.  A  large  number  of  observations  has  been  made.  In  practice, 
the  number  of  observations  may  be  considerably  reduced  if  the 
second  condition  is  fulfilled.  In  the  ordinary  course  of  things  from 
10  to  25  is  usually  considered  a  sufficient  number. 

§  164.  Maxwell's  Law  of  Distribution  of  Molecular  Velocities. 

In  a  preceding  discussion,  the  velocities  of  the  molecules  of  a 
gas  were  assumed  to  be  the  same.  Can  this  simplifying  assump- 
tion be  justified  ? 

According  to  the  kinetic  theory,  a  gas  is  supposed  to  consist  of 
a  number  of  perfectly  elastic  spheres  moving  about  in  space  with  a 
certain  velocity.  In  case  of  impact  on  the  walls  of  the  bounding 
vessel,  the  molecules  are  supposed  to  rebound  according  to  known 
dynamical  laws.  This  accounts  for  the  pressure  of  a  gas.  The 
velocities  of  all  the  molecules  of  a  gas  in  a  state  of  equilibrium  are 
not  the  same.  Some  move  with  a  greater  velocity  than  others. 
At  one  time  a  molecule  may  be  moving  with  a  great  velocity,  at 
another  time,  with  a  relatively  slow  speed.  The  attempt  has  been 
made  to  find  a  law  governing  the  distribution  of  the  velocities  of 
the  motions  of  the  different  molecules,  and  with  some  success. 
Maxwell's  law  is  based  upon  the  assumption  that  the  same  relations 
hold  for  the  velocities  of  the  molecules  as  for  errors  of  observation. 
This  assumption  has  played  a  most  important  part  in  the  develop- 
ment of  the  kinetic  theory  of  gases.  The  probability  y  that  a  mole- 
cule will  have  a  velocity  equal  to  v  is  given  by  an  expression  of  the 
type  : 

y=TA«)e      •  ;-'■';;■:■   (1) 

Very  few   molecules   will   have   velocities   outside  a   certain   re- 


§  164.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.       535 


stricted  range.     It  is  possible  for  a  molecule  to  have  any  velocity 

whatever,  but  the  probability  of  the 

existence  of  velocities  outside  certain 

limits   is   vanishingly   small.      The 

reader  will  get  a  better  idea  of  the 

distribution  of  the  velocities  of   the 

molecules  by  plotting  the  graph  of 

the  above  equation  for  himself.     Re-      °  °*  v° 

member  that  the  ordinates  are  pro-  Fla-  1^2« 

portional   to   the  number  of  molecules,  abscissae  to  their  speed. 

Areas  bounded  by  the  rr-axis,  the  curve  and  certain  ordinates  will 

give  an  idea  of  the  number  of  molecules  possessing   velocities 

between  the  abscissae  corresponding  to  the  boundary  ordinates.     In 

Fig.  172  the  shaded  portion  represents  the  number  of  molecules 

with  velocities  lying  between  V0  and  1'5F0. 

Example. — By  the  ordinary  methods  for  finding  a  maximum,  show  from 
(1),  that  y  is  a  maximum  when  v  =  a. 

Returning  to  the  study  of  the  kinetic  theory  of  gases,  p.  504, 
the  number  of  molecules  with  velocities  between  v  and  v  -+•  dv  is 
assumed  to  be  represented  by  an  equation  analogous  to  the  ex- 
pression employed  to  represent  the  errors  of  mean  square  of  page 
525,  namely, 

-t£)v(i)'<#  ■    •    •  « 

where  N  represents  the  total  number  of  molecules,  a  is  a  constant 
to  be  evaluated. 

I.  To  find  a  value  for  the  constant  a  in  terms  of  the  average 
velocity  VQ  of  the  molecules.  Since  there  are  dN  molecules  with  a 
velocity  v,  the  sum  of  the  velocities  of  all  these  dN  molecules  is 
vdN,  and  the  sum  of  the  velocities  of  all  the  molecules  must  be 

j::: -"•••' ■-M:®'--H$'M-t. 

from  (2).     Where  has  N  gone  ?     The  average  velocity  V0  is  one 
Nth  of  the  sum  of  the  velocities  of  the  N  given  molecules.     Hence, 

a  =  £*W* (3) 

II.  To  find  the  average  velocity  of  the  molecules  of  a  gas.  By  a 
well-known  theorem  in  elementary  mechanics,  the  kinetic  energy  of 
a  mass  m  moving  with  a  velocity  v  is  %mv2.    Hence,  the  sum  of  the 


536  HIGHER  MATHEMATICS.  §  164. 

kinetic  energies  of  the  dN  molecules  will  be  %(mdN)v2t  because 
there  are  dN  molecules  moving  with  a  velocity  v.  From  (2),  there- 
fore, the  total  kinetic  energy  (T)  of  all  the  molecules  is 

T  =  imv* .  dN=  4^       v* .  e     °»&o  =  -ANm**  =  -AMaK 

•••a  =  2Vi  •     •     •     •    w 

where  M  =  J7m  =  total  mass  of  N  molecules  each  of  mass  m.  The 
total  kinetic  energy  of  N  molecules  of  the  same  kind  is 

T  =  \mv\  +  \mv>\  +  . . .  +  \mvl  =  %m{v-?  +  v22  +  . .  .  +  v%).  (5) 
The  velocity  of  mean  square,  U,  is  defined  as  the  velocity  whose 
square  is  the  average  of  the  squares  of  the  velocities  of  all  the  N 
molecules,  or, 

-?j2    4.   7)2    J.  «a  1  1 

from  (5).     Again,  from  (4)  and  (6),  we  have 

a  =  vr;F»=7^  =  0'9213a-    •    •    (7) 

Most  works  on  chemical  theory  give  a  simple  method  of  proving 
that  if  p  denotes  the  pressure  and  p  the  density  of  a  gas, 

p-ipU*.  .  '  .  .  .  (8) 
This  in  conjunction  with  (6)  allows  the  average- velocity  of  the 
molecules  of  a  gas  to  be  calculated  from  the  known  values  of  the 
pressure  and  density  of  the  gas,  as  shown  in  any  Textbook  on 
Physical  Chemistry. 

The  reader  is  no  doubt  familiar  with  the  principle  underlying 
Maxwell's  law,  and,  indeed,  the  whole  kinetic  theory  of  gases.  I 
may  mention  two  examples.  The  number  of  passengers  on  say 
the  3-10  p.m.  suburban  daily  train  is  fairly  constant  in  spite  of  the 
fact  that  that  train  does  not  carry  the  same  passenger  two  days 
running.  Insurance  companies  can  average  the  number  of  deaths 
per  1,000  of  population  with  great  exactness.  Of  course  I  say 
nothing  of  disturbing  factors.  A  bank  holiday  may  require  pro- 
vision for  a  supra-normal  traffic,  and  an  epidemic  will  run  up  the 
death  rate  of  a  community.  The  commercial  success  of  these 
institutions  is,  however,  sufficient  testimony  of  the  truth  of  the 
method  of  averages,  otherwise  called  the  statistical  method 
of  investigation.  The  same  type  of  mathematical  expression  is 
required  in  each  case. 


§  165.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      537 

It  will  thus  be  seen  that  calculations,  based  on  the  supposition 
that  all  the  molecules  possess  equal  velocities,  are  quite  admissible 
in  a  first  approximation.  The  net  result  of  the  "  dance  of  the  mole- 
cules "  is  a  distribution  of  the  different  velocities  among  all  the 
molecules,  which  is  maintained  with  great  exactness. 

G.  H.  Darwin  has  deduced  values  for  the  mean  free  path,  eto.,  from  the 
hypothesis  that  the  molecules  of  the  same  gas  are  not  all  the  same  size.  He 
has  examined  the  oonsequences  of  the  assumption  that  the  sizes  of  the  mole- 
cules are  ranged  aocording  to  a  law  like  that  governing  the  frequency  of  errors 
of  observation.  For  this,  see  his  memoir  "  On  the  mechanical  conditions  of  a 
swarm  of  meteorites  "  (Phil.  Trans.,  180, 1,  1889). 


§  165.  Constant  or  Systematic  Errors. 

The  irregular  accidental  errors  hitherto  discussed  have  this 
distinctive  feature,  they  are  just  as  likely  to  have  a  positive  as  a 
negative  value.  But  there  are  errors  whioh  have  not  this  character. 
If  the  barometer  vacuum  is  imperfect,  every  reading  will  be  too 
small;  if  the  glass  bulb  of  a  thermometer  has  contracted  after 
graduation,  the  zero  point  rises  in  such  a  way  as  to  falsify  all 
subsequent  readings ;  if  the  points  of  suspension  of  the  balance 
pans  are  at  unequal  distances  from  the  centre  of  oscillation  of  the 
beam,  the  weighings  will  be  inaccurate.  A  change  of  tempera- 
ture of  5°  or  6°  may  easily  cause  an  error  of  0*2  to  1*0  0/o  in  an 
analysis,  owing  to  the  change  in  the  volume  of  the  standard 
solution.  Such  defective  measurements  are  said  to  be  affected 
by  oonstant  errors.1  By  definition,  constant  errors  are  produced 
by  well-defined  causes  which  make  the  errors  of  observation  pre- 
ponderate more  in  one  direction  than  in  another.  Thus,  some  of 
Dumas'  determinations  of  the  atomic  weight  of  silver  are  affected  by 
a  constant  error  due  to  the  occlusion  of  oxygen  by  metallic  silver  in 
the  course  of  his  work. 

One  of  the  greatest  trials  of  an  investigator  is  to  detect  and  if 


1  Pergonal  error.  This  is  another  type  of  constant  error  which  depends  on  the 
personal  qualities  of  the  observer.  Thus  the  differences  in  the  judgments  of  the 
astronomers  at  the  Greenwich  Observatory  as  to  the  observed  time  of  transit  of  a  star 
and  the  assumed  instant  of  its  actual  occurrence,  are  said  to  vary  from  y^j-  to  -J-  of  a 
second,  and  to  remain  fairly  constant  for  the  same  observer.  Some  persistently  read 
the  burette  a  little  high,  others  a  little  low.  Vernier  readings,  analyses  based  on 
colorimetric  tests  (such  as  Nessler's  ammonia  process),  etc.,  may  be  affected  by 
personal  errors. 


538  HIGHER  MATHEMATICS.  §  165. 

possible  eliminate  constant  errors.  "  The  history  of  science  teaches 
air  too  plainly  the  lesson  that  no  single  method  is  absolutely  to  be 
relied  upon,  and  that  sources  of  error  lurk  where  they  are  least 
expected,  and  that  they  may  escape  the  notice  of  the  most  ex- 
perienced and  conscientious  worker."  \  Two  questions  of  the 
gravest  moment  are  now  presented.  How  are  constant  errors  to 
be  detected  ?  How  may  the  effect  of  constant  errors  be  eliminated 
from  a  set  of  measurements  ?  This  is  usually  done  by  modifying 
the  conditions  under  which  the  experiments  are  performed.  "  It  is 
only  by  the  concurrence  of  evidence  of  various  kinds  and  from  various 
sources,"  continues  Lord  Eayleigh,  "  that  practical  certainty  may 
at  least  be  attained,  and  complete  confidence  restored."  Thus  the 
magnitude  is  measured  under  different  conditions,  with  different 
instruments,  etc.  It  is  assumed  that  even  though  each  method  or 
apparatus  has  its  own  specific  constant  error,  all  these  constant 
errors  taken  collectively  will  have  the  character  of  accidental  errors. 
To  take  a  concrete  illustration,  faulty  "  sights  "  on  a  rifle  may  cause 
a  constant  deviation  of  the  bullets  in  one  direction ;  the  "  sights  " 
on  another  rifle  may  cause  a  constant  "  error  "  in  another  direc- 
tion, and  so,  as  the  number  of  rifles  increases,  the  constant  errors 
assume  the  character  of  accidental  errors  and  thus,  in  the  long 
run,  tend  to  compensate  each  other.  This  is  why  Stas  generally 
employed  several  different  methods  to  determine  his  atomic  weights. 
To  quote  one  practioal  case,  Stas  made  two  sets  of  determinations 
of  the  numerical  value  of  the  ratio  Ag  :  KOI.  In  one  set,  four  series 
of  determinations  were  made  with  KG  prepared  from  four  different 
sources  in  conjunction  with  one  specimen  of  silver,  and  in  the  other 
set  different  series  of  experiments  were  made  with  silver  prepared 
from  different  sources  in  conjunction  with  one  sample  of  KC1.  Un- 
fortunately the  latter  set  was  never  completed. 

The  calculation  of  an  arithmetical  mean  is  analogous  to  the 
process  of  guessing  the  centre  of  a  target  from  the  distribution  of 
the  "hits"  (Fig.  165).  If  all  the  shots  are  affected  by  the  same 
constant  error,  the  centre,  so  estimated,  will  deviate  from  the  true 
centre  by  an  amount  depending  on  the  magnitude  of  the  (presumably 
unknown)  constant  error.  If  this  magnitude  can  be  subsequently 
determined,  a  simple  arithmetical  operation  (addition  or  subtraction) 
will  give  the  correct  value.     Thus  Stas  found  that  the  amount  of 

1  Lord  Rayleigh's  Presidential  Address,  B.A.  Reports,  1884. 


§  166.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      539 

potassium  chloride  equivalent  to  100  parts  of  silver  in  one  case 
was  as 

Ag:K01  =  100:69-1209. 
The  KOI  was  subsequently  found  to  contain  000259  per  cent,  of 
silica.     The  chemical  student   will  see   that   0*00179    has   conse- 
quently to  be  subtracted  from  69-1209.     Hence, 

Ag  :  KC1  =  100  :  69-11903. 
After  Lord  Rayleigh  (Proc.  Boy.  Soc.,  43,  356,  1888)  had  proved 
that  the  capacity  of  an  exhausted  glass  globe  is  less  than  when  the 
globe  is  full  of  gas,  all  measurements  of  the  densities  of  gases 
involving  the  use  of  exhausted  globes  had  to  be  corrected  for 
shrinkage.  Thus  Regnault's  ratio,  1  :  15-9611,  for  the  relative 
densities  of  hydrogen  and  oxygen  was  "corrected  for  shrinkage" 
to  1  :  159105.  The  proper  numerical  corrections  for  the  constant 
errors  of  a  thermometer  are  indicated  on  the  well-known  "  Kew 
certificate,"  etc. 

If  the  mean  error  of  each  set  of  results  differs,  by  an  amount 
to  be  expected,  from  the  mean  errors  of  the  different  sets  measured 
with  the  same  instrument  under  the  same  conditions,  no  constant 
error  is  likely  to  be  present.  The  different  series  of  atomic  weight 
determinations  of  the  same  chemical  element,  published  by  the 
same,  or  by  different  observers,  do  not  stand  this  test  satisfactorily. 
Hence,  Ostwald  concludes  that  constant  errors  must  have  been 
present  even  though  they  have  escaped  the  experimenter's  ken. 

Example. — Discuss  the  following:  "Merely  increasing  the  number  of 
experiments,  without  varying  the  conditions  or  method  of  observation, 
diminishes  the  influence  of  accidental  errors.  It  is,  however,  useless  to 
multiply  the  number  of  observations  beyond  a  certain  limit.  On  the  other 
hand,  the  greater  the  number  and  variety  of  the  observations,  the  more 
complete  will  be  the  elimination  of  the  effects  of  both  constant  and  accidental 
errors." 

§  166.  Proportional  Errors. 

One  of  the  greatest  sources  of  error  in  scientific  measurements 
occurs  when  the  quantity  cannot  be  measured  directly.  In  such 
oases,  two  or  more  separate  observations  may  have  to  be  made  on 
different  magnitudes.  Each  observation  contributes  some  little 
inaccuracy  to  the  final  result.  Thus  Faraday  has  determined  the 
thickness  of  gold  leaf  from  the  weight  of  a  certain  number  of 
sheets.     Foucault  measures  time,  Le  Chatelier  measures  tempera- 


540  HIGHER  MATHEMATICS.  §  166. 

ture  in  terms  of  an  angular  deviation.  The  determination  of  the 
rate  of  a  chemical  reaction  often  depends  on  a  number  of  more  or 
less  troublesome  analyses.1 

For  this  reason,  among  others,  many  chemists  prefer  the  standard  0  =  16 
as  the  basis  of  their  system  of  atomic  weights.  The  atomic  weights  of  most 
of  the  elements  have  been  determined  directly  or  indirectly  with  reference  to 
oxygen.  If  H  =  1  be  the  basis,  the  atomic  weights  of  most  of  the  elements 
depend  on  the  nature  of  the  relation  between  oxygen  and  hydrogen — a 
relation  which  has  not  yet  been  fixed  in  a  satisfactory  manner.  The  best  de- 
terminations made  since  1887  vary  between  H  :  0  =  1 :  15*96  and  H  :  0  =  1 :  15*87. 
If  the  former  ratio  be  adopted,  the  atomic  weights  of  antimony  and  uranium 
would  be  respectively  119*6  and  239*0  ;  while  if  the  latter  ratio  be  employed, 
these  units  become  respectively  118*9  and  237*7,  a  difference  of  one  and  two 
units  I  It  is,  therefore,  better  to  contrive  that  the  atomic  weights  of  the 
elements  do  not  depend  on  the  uncertainty  of  the  ratio  H  :  O,  by  adopting 
the  basis :  O  =  16. 

If  the  quantity  to  be  determined  is  deduced  by  calculation  from 
a  measurement,  Taylor's  theorem  furnishes  a  convenient  means  of 
criticizing  the  conditions  under  which  any  proposed  experiment  is 
to  be  performed,  and  at  the  same  time  furnishes  a  valuable  insight 
into  the  effect  of  an  error  in  the  measurement  on  the  whole  result. 
It  is  of  the  greatest  importance  that  every  investigator  should 
have  a  clear  idea  of  the  different  sources  of  error  to  which  his 
results  are  liable  in  order  to  be  able  to  discriminate  between  im- 
portant and  unimportant  sources  of  error,  and  to  find  just  where 
the  greatest  attention  must  be  paid  in  order  to  obtain  the  best 
results.  The  necessary  accuracy  is  to  be  obtained  with  the  least 
expenditure  of  labour. 

I.  Proportional  errors  of  simple  measurements.  Let  y  be  the 
desired  quantity  to  be  calculated  from  a  magnitude  x  which  can  be 
measured  directly  and  is  connected  with  y  by  the  relation 

V  =  /<»• 
f(x)  is  always  affected  with  some  error  dx  which  causes  y  to  deviate 
from  the  truth  by  an  amount  dy.     The  error  will  then  be 
dy=  (y  +  dy)  -  y  =  f(x  +  dx)  -  f(x). 

1  Indirect  results  are  liable  to  another  source  of  error.  The  formula  employed 
may  be  so  inexact  that  accurate  measurements  give  but  grossly  approximate  results. 
For  instance,  a  first  approximation  formula  may  have  been  employed  when  the 
accuracy  of  the  observations  required  one  more  precise  ;  ir  =  -^  may  have  been  put 
in  place  of  ir  =  8*14159  ;  or  the  coefficient  of  expansion  of  a  perfect  gas  has  been 
applied  to  an  imperfect  gas.    Such  errors  are  called  errors  of  method. 


§166.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      541 

dx  is  necessarily  a  small  magnitude,  therefore,  by  Taylor's  theorem, 

fx  +  dx)  =  f(x)  +  f{x) .  dx  +  . . ., 
or,  neglecting  the  higher  orders  of  magnitude, 

dy  ~  f'(x) .  dp. 
The  relation  between  the  error  and  the  total  magnitude  of  y  is 

%      f(x) .  dx 

y  "     Ax)    •  '        '        '        W 

All  this  means  is  that  the  differential  of  a  function  represents  the 
change  in  the  value  of  the  function  when  the  variable  suffers  an 
infinitesimal  change.  The  student  learned  this  the  first  day  he 
attacked  the  calculus.  The  ratio  dy  :  y  is  called  the  proportional, 
relative,  or  fractional  error,  that  is  to  say,  the  ratio  of  the  error 
involved  in  the  whole  process  to  the  total  quantity  sought ;  while 
lOOdy  :  y  is  called  the  percentage  error.  The  degree  of  accuracy 
of  a  measurement  is  determined  by  the  magnitude  of  the  propor- 
tional error. 

_  ,  _  Magnitude  of  error 

Proportional  Error  =  =-— ; r^- — 

Total  magnitude  of  quantity  measured 

Students  often  fail  to  understand  why  their  results  seem  all 
wrong  when  the  experiments  have  been  carefully  performed  and 
the  calculations  correctly  done.  For  instance,  the  molecular 
weight  of  a  substance  is  known  to  be  either  160,  or  some  multiple 
of  160.  To  determine  whioh,  0*380  (or  w)  grm.  of  the  substance 
was  added  to  14-01  (or  w^)  grms.  of  acetone  boiling  at  B{  (or  3*50°) 
on  Beckmann's  arbitrary  scale,  the  temperature,  in  consequence, 
fell  to  $2°.  (or  3-36°) ;  the  molecular  weight  of  the  substance,  M,  is 
then  represented  by  the  known  formula 

M=  1670^f^7);  or'  M  "  imu^u  " 323' 

or  approximately  2  x  160.  Now  assume  that  the  temperature 
readings  may  be  ±  0*05°  in  error  owing  to  convection  currents, 
radiation  and  conduction  of  heat,  etc.     Let  0^  =  3  -55°  and  02°  =  3'31°, 

.-.  M  =  1670.     °/38°        -188. 
14*01  x  0*24 

This  means  that  an  error  of  ±  ^°  in  the  reading  of  the  thermometer 

would  give  a  result  positively  misleading.     This  example  is  by  no 

means  exaggerated.     The  simultaneous  determination  of  the  heat 

of  fusion  and  of  the  specific  heat  of  a  solid  by  the  solution  of  two 

simultaneous  equations,  and  the  determination  of  the  latent  heat 


542  HIGHER  MATHEMATICS.  §  166. 

of  steam  are  specially  liable  to  similar  mistakes.  A  study  of  the 
reduction  formula  will  show  in  every  case  that  relatively  small 
errors  in  the  reading  of  the  temperature  are  magnified  into  serious 
dimensions  by  the  method  used  in  the  calculation  of  the  final 
result. 

Examples. — (1)  Almost  any  text-book  on  optics  will  tell  you  that  the 
radius  of  curvature,  r,  of  a  lens,  is  given  by  the  formula 

f~a 

Let  the  true  values  of  /  and  a  be  respectively  20  and  15.  Let  /  and  a  be  liable 
to  error  to  the  extent  of  +  0-5,  say,  /  is  read  20*5,  and  a  14*5.  Then  the  true 
value  of  r  is  60,  the  observed  value  51'2.  Fractional  error  =  ^-.  This  means 
that  an  error  of  about  0*5  in  20,  i.e.,  2*5 °/0  in  the  determination  of  /and  a 
may  cause  r  to  deviate  15  °/0  from  the  truth. 
(2)  In  applying  the  formula 

for  the  influence  of  temperature  on  the  velocity,  V,  of  a  chemical  reaction 
show  that  an  error  of  1°  in  the  determination  of  2\,  at  about  300°  abs.,  will 
give  a  fractional  error  of  2*4  in  the  determination  of  V.  Hint.  Substitute 
Tj  =  300,  T0  =  273.  Use  Table  IV.  I  make  V  =  41*52.  Now  put  Tx  =  301. 
I  get  V  =  43*79,  etc.  Hence  an  error  of  1°  will  make  V  vary  about  6  °/0  from 
its  true  value. 

If  we  knew  that  an  astronomer  had  made  an  absolute  error  of 
100,000  miles  in  estimating  the  distance  between  the  earth  and 
the  sun,  and  also  that  a  physicist  had  made  an  absolute  error  of 

*ne  i oooo^oooooo^  °f  a  mile  m  measuring  the  wave  length  of  a 
spectral  line,  we  could  form  no  idea  of  the  relative  accuracy  of  the 
two  measurements  in  spite  of  the  fact  that  the  one  error  is  the 
ioo o,o oo 0^0,0 oo oo otn  Part  °f tne  other.  In  the  first  measurement 
the  error  is  about  TToVo  °f  tne  whole  quantity  measured,  in  the 
second  case  the  error  is  about  the  same  order  of  magnitude  as  the 
quantity  measured.  In  the  former  case,  therefore,. the  error  is  neg- 
ligibly small ;  in  the  latter,  the  error  renders  the  result  nugatory. 

It  is  therefore  important  to  be  able  to  recognise  the  weak  and 
strong  points  of  a  given  method  of  investigation;  to  grade  the 
degree*  of  accuracy  of  the  different  stages  of  the  work  so  as  to 
produce  the  required  result ;  so  as  to  have  enough  at  all  points, 
but  no  superfluity.  I  have  already  spoken  of  the  need  for 
"  scientific  perspective "  in  dealing  with  numerical  computations. 

Examples. — (1)  It  is  required  to  determine  the  capacity  of  a  sphere  from 


§  166.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      543 

the  measurement  of  its  diameter.  Let  y  denote  the  volume,  x  the  diameter, 
then,  by  a  well-known  mensuration  formula,  y  =  lirx*.  It  is  required  to  find 
the  effect  of  a  small  error  in  the  measurement  of  the  diameter  on  the  cal- 
culated volume.     Suppose  an  error  dx  is  committed  in  the  measurement,  then 

y  +  dy  =  £ir(aj  +  dx)3  =  \ir{x*  +  SxWx  +  3x(dx)*  +  (dx)9}. 

By  hypothesis,  dx  is  a  very  small  fraction,  therefore,  by  neglecting  the  higher 
powers  of  dx  and  dividing  the  result  by  the  original  expression 

y  +  dy     1  (x*  +  8x2dx\    dy  _    dx 
y     T\      \«r      )''  y  ~x' 

Or,  the  error  in  the  calculated  result  is  three  times  that  made  in  the 
measurement.  Henoe  the  necessity  for  extreme  precautions  in  measuring 
the  diameter.  Sometimes,  we  shall  find,  it  is  not  always  necessary  to  be  so 
careful*  The  same  result  could  have  been  more  easily  obtained  by  the  use  of 
Taylor's  theorem  as  described  above.  Differentiate  the  original  expression 
and  divide  the  result  by  the  original  expression.  We  thus  get  the  relative 
error  without  trouble. 

•  (2)  Criticize  the  method  for  the  determination  of  the  atomic  weight  of 
lead  from  the  ratio  Pb  :  0  in  lead  monoxide.  Let  y  denote  the  atomic 
weight  of  lead,  a  the  atomic  weight  of  oxygen  (known).  It  is  found  ex- 
perimentally that  x  parts  of  lead  combine  with  one  part  of  oxygen,  the 
required  atomio  weight  of  lead  is  determined  from  the  simple  proportion 

j/:a«=x:l;  or,  y=ax;  or,  dy~adx;  ,\dy\y—dx\x.  .      (2) 

Thus  an  error  of  1  °/0  in  the  determination  of  x  introduces  an  equal  error  in 
the  calculated  value  of  y.  Other  things  being  equal,  this  method  of  finding 
the  atomic  weight  of  lead  is,  therefore,  very  likely  to  give  good  results. 

(3)  Show  that  the  result  of  determining  the  atomic  weight  of  barium  by 
precipitation  of  the  chloride  with  silver  nitrate  is  less  influenced  by  experi- 
mental errors  than  the  determination  of  the  atomic  weight  of  sodium  in  the 
same  way.  Assume  that  one  part  of  silver  as  nitrate  requires  x  parts  of  sodium 
(or  barium)  chloride  for  precipitation  as  silver  chloride.  Let  a  and  b  be  the 
known  atomic  weights  of  silver  and  chlorine.  Then,  if  y  denotes  the  atomic 
weight  of  sodium,  y+b:  a=x:l;  oity  =  ax-b;  .  \  a=(y  +  b)/x.  Differentiate, 
and  substitute  y=23,  6  =  35*5. 

dy  a  y  +  b   dx  dx 

—  =  tCLX  =         *    .  — •  =  Jo*o4 — > 

y       ax  -  b  y        x  x' 

or  an  error  of  1  °/0  in  the  determination  of  chlorine  in  sodium  will  introduce 
an  error  of  2-5  °/0  in  the  atomic  weight  of  sodium.  Hence  it  is  a  disadvantage 
to  have  b  greater  than  y.  For  barium,  the  error  introduced  is  1*5  %  instead 
of  2-5  °/0. 

(4)  If  the  atomic  weight  of  barium  y  is  determined  by  the  precipitation  of 
barium  sulphate  from  barium  chloride  solutions,  and  a  denotes  the  known 
atomio  weight  of  chlorine,  b  the  known  combining  weight  of  S04,  then  when 
x  parts  of  barium  chloride  are  converted  into  one  part  of  barium  sulphate, 

,  a  ,   ,  i     dy  (b  -  2a)dx 

v  +  2a  :  v  +  b  =  x  :1    —  = 

y  y  '    y        (1  -  x)  (bx  -  2a)' 


644  HIGHER  MATHEMATICS.  §  166. 

(5)  An  approximation  formula  used  in  the  determination  of  the  viscosity 
of  liquids  is 

irptr4 

where  v  denotes  the  volume  of  liquid  flowing  from  a  capillary  tube  of  radius  r 
And  length  I  in  the  time  t ;  p  is  the  actual  pressure  exerted  by  the  column  of 
liquid.  Show  that  the  proportional  error  in  the  calculation  of  the  viscosity  t\ 
is  four  times  the  error  made  in  measuring  the  radius  of  the  tube. 

(6)  In  a  tangent  galvanometer,  the  tangent  of  the  angle  of  deflection  of 
the  needle  is  proportional  to  the  current.  Prove  that  the  proportional  error 
in  the  calculated  value  of  the  current  due  to  a  given  error  in  the  reading  is 
least  when  the  deflection  is  45°.  The  strength  of  the  current  is  proportional 
to  the  tangent  of  the  displaced  angle  x,  or 

„  .       _ .  ,        G.  dx  dy  dx 

y  =  fix)  =  C  tan  x :  .\  dv  — k—  :   or,  —  =  -s • 

9     JK  '  '       «*#       00S2X  i   w*.    y       sin  a;,  cos  a? 

To  determine  the  minimum,  put 


±(dy\ss 


sin2#  -  cos2a; 
sTn2a? .  oos2a:  =  ° '  •'•  sin  x  m  cosx>  or'  sin  x  m  cos  x' 


This  is  true  only  in  the  neighbourhood  of  45°  (Table  XIV.),  and,  therefore,  in 

this  region  an  error  of  observation  will  have  the  least  influence  on  the  final 

result.    In  other  words,  the  best  results  are  obtained  with  a  tangent  galvo- 

nometer  when  the  needle  is  deflected  about  45°. 

What  will  be  the  effect  of  an  error  of  0*25°  in  reading  a  deflection  of  42°, 

on  the  calculated  current  ?    Note  that  x  in  the  above  formula  is  expressed  in 

circular  or  radian  measure  (page  606).     Hence, 

_  x  0*25 
0*25(degrees)  m  — 18Q      =  0*00436(radians). 

,  dy  dx  2dx        Q-Q0872  _  n  nQ     .      Q  0/ 

•''  y  *  sin  x  .  cos  x  ~  sin  2x  =  sin  84°  "  °'09;  l-e''  9   /o' 

since,  from  a  Table  of  Natural  Sines,  sin  84°  =  0*9946. 

(7)  Show  that  the  proportional  error  involved  in  the  measurement  of  an 
electrical  resistance  on  a  Wheatstone's  bridge  is  least  near  the  middle  of  the 
bridge.  Let  B  denote  the  resistance,  I  the  length  of  the  bridge,  x  the  distance 
of  the  telephone  from  one  end.  .*.  y  =  Rxftl  +  x).  Proceed  as  above  and 
show  that  when  x  =••  £1  (the  middle  of  the  bridge),  the  proportional  error  is  a 
minimum. 

(8)  By  Newton's  law  of  attraction,  the  force  of  gravitation,  g,  between 
two  bodies  varies  directly  as  their  respective  masses— Wj,  m2 — and  inversely  as 
the  square  of  their  distance  apart,  r.  The  mass  of  each  body  is  supposed  to 
be  collected  at  its  centroid  (centre  of  gravity).  The  weight  of  one  gram  at 
Paris  is  equivalent  to  880-868  dynes.  The  dyne  is  the  unit  of  force.  Hence 
Newton's  law,  g  =  /im1w2/r2  (dynes),  may  be  written  w  =  ajr2  (grams),  where 
a  is  a  constant  equivalent  to  /*  x  mx  x  m^  x  980*868.  Hence  show  that  for 
small  changes  in  altitude  dwjw  =  -  2dr(r.  Marek  was  able  to  detect  a  differ- 
ence of  1  in  500,000,000  when  comparing  the  kilogram  standards  of  the 
Bureau  International  des  Poids  et  Mesures.  Hence  show  that  it  is  possible 
to  detect  a  difference  in  the  weight  of  a  substance  when  one  scale  pan  of  the 


§  166.     PEOBABILITY  AND  THE  THEORY  OF  ERRORS.      545 

balance  is  raised  one  centimetre  higher  than  the  other.    Hint.  Radius  of  earth 
=  r  =  637,130,000  cm.  ;  w  =  1  kilogrm.  ;  dr  =  1  cm. ; 

dw  _  2  1  1 

'"•  ~~w  ~  637,130,000  =  318,565,000'    This  1S  S188-*61  tnan  500,000"^00' 

As  a  further  exercise,  show  that  a  kilogram  will  lose  0-00003  grm.,  if  it  be 

weighed  10  cm.  above  its  original  position.      Hint.   Find  -dw  ;  r  has  its 

former  value ;  w  =  1000  grm. ;  dr  =  10  cm. 

II.  Proportional  error  of  composite  measurements.  Whenever  a 
result  has  to  be  determined  indirectly  by  combining  several  different 
species  of  measurements — weight,  temperature,  volume,  electro- 
motive force,  etc. — the  effect  of  a  percentage  error  of,  say,  1  per 
cent,  in  the  reading  of  the  thermometer  will  be  quite  different  from 
the  effect  of  an  error  of  1  per  cent,  in  the  reading  of  a  voltmeter. 

It  is  obvious  that  some  observations  must  be  made  with 
greater  care  than  others  in  order  that  the  influence  of  each  kind 
of  measurement  on  the  final  result  may  be  the  same.  If  a  large 
error  is  compounded  with  a  small  error,  the  total  error  is  not  ap- 
preciably affected  by  the  smaller.  Hence  Ostwald  recommends 
that  "  a  variable  error  be  neglected  if  it  is  less  than  one-tenth  of 
the  larger,  often,  indeed,  if  it  is  but  one-fifth  ". 

Examples. — (1)  Joule's  relation  between  the  strength  of  a  current  G 
(amperes)  and  the  quantity  of  heat  Q  (calories)  generated  in  an  electric  con- 
ductor of  resistance  B  (ohms)  in  the  time  t  (seconds),  is,  Q  =  0'24:C2Bt.  Show 
that  B  and  t  must  be  measured  with  half  the  precision  of  C  in  order  to  have 
the  same  influence  on  Q. 

(2)  What  will  be  the  fractional  error  in  Q  corresponding  to  a  fractional 
error  of  0*1  %  in  B  ?     Ansr.  0-001,  or  0-1  °/0. 

(3)  What  will  be  the  percentage  error  in  C  corresponding  to  0-02  °/0  m  Q  ? 
Ansr.  0-01  %. 

(4)  If  the  density  s  of  a  substance  be  determined  from  its  weights  (w,  Wj) 
in  air  and  water,  and  remembering  that  s  =  wll{w-w-^),  show  that 

ds  _     w      /dwj     dw\ 
s  ~  w-w\w1      w  )* 

(5)  The  specific  heat  of  a  substance  determined  by  the  method  of  mixtures 
is  given  by  the  formula 

7^0(02 -gj) 

m(e-62)  ' 
where  m  is  the  weight  of  the  substance  before  the  experiment ;  wx  the  weight 
of  the  water  in  the  calorimeter ;  c  the  mean  specific  heat  of  water  between 
62  and  0j ;  6  is  the  temperature  of  the  body  before  immersion ;  dx  the  maximum 
temperature  reached  by  the  water  in  the  calorimeter ;  02  the  temperature  of 
the  system  after  equalization  of  the  temperature  has  taken  place.  Supposing 
the  water  equivalent  of  the  apparatus  is  included  in  mv  what  will  be  the 
effect  of  a  small  error  in  the  determination  of  the  different  temperatures  on 
the  result? 

MM 


546  HIGHER  MATHEMATICS.  §  166. 

First,  error  in  0^  Show  that  ds/s  =  -  d0,/(02  -  0i)-  If  an  error  of  say  0*1° 
is  made  in  a  reading  and  02  -  dx  =  10°,  the  error  in  the  resulting  specific  heat 
is  about  1  °/0.  If  a  maximum  error  of  O'Ol  %  is  to  be  permitted,  the  tempera- 
ture must  be  read  to  the  0-0001°. 

Second,  error  in  0.  Show  that  dsfs  =  -  ^0/(0  -  02).  If  a  maximum  error  in 
the  determination  of  s  is  to  be  0*1  °/0,  when  0  -  02  =  50°,  0  must  be  read  to  the 
0*05°.  If  an  error  of  0*1°  is  made  in  reading  the  temperature  and  0  -  02  =  50°, 
show  that  the  resulting  error  in  the  specific  heat  will  be  0'2  °/0. 

Third,  error  in  02.  Show  that  ds/s  =  de.2l(62-01)  +  d02/(0-02).  If  the 
maximum  error  allowed  is  0-1  °/0  and  02-01  =  lO°,  0-02  =  50°,  show  that  02 
must  be  read  to  the  yfg-0  ;  while  if  an  error  of  0-1°  is  made  in  the  reading  of 
02,  show  that  the  resulting  error  in  the  specific  heat  is  l"2°/0. 

(6)  In  the  preceding  experiment,  if  mx  =  100  grams,  show  that  the 
weighing  need  not  be  taken  to  more  than  the  0*1  gram  for  the  error  in  s  to  be 
within  0*1  % ;  and  for  m,  need  not  be  closer  than  0*5  gram  when  m  is  about 
50  grams. 

Since  the  actual  errors  are  proportional  to  the  probable  errors, 
the  most  probable  or  mean  value  of  the  total  error  du,  is  obtained 
from  the  expression 

(*.)•- (5s;Ah)+(^)+ (3) 

from  (16),  §  162,  page  530.  Note  the  squared  terms  are  all  positive. 
Since  the  errors  are  fortuitous,  there  will  be  as  many  positive  as 
negative  paired  terms.  These  will,  in  the  long  run,  approximately 
neutralize  each  other.     Hence  (3). 

Examples. —  (1)  Divide  equation  (3)  by  u?t  it  is  then  easy  to  show  that 
(dQ\*      JdC\*      fdBy      /dty 

\-q)  -n-o)  +{-e)  +(v> 

from  the  preceding  set  of  examples.  Hence  show  that  the  fractional  error  in 
Q,  corresponding  to  the  fractional  errors  of  0-03  in  G,  0'02  in  B  and  0-03  in  t, 
is  0-07. 

(2)  The  regular  formula  for  the  determination  of  molecular  weight  of  a 
substance  by  the  freezing  point  method,  is  M  =  KwjQ,  where  K  is  a  constant, 
M  the  required  molecular  weight,  w  the  weight  of  the  substance  dissolved  in 
100  grams  of  the  solvent,  0  the  lowering  of  the  freezing  point.  In  an  actual 
determination,  w  =  0-5139,  0  =  0-295,  K  =  19  (Perkin  and  Kipping's  Organic 
Chemistry),  what  would  be  the  effect  on  If  of  an  error  of  0"01  in  the  deter- 
mination of  w,  and  of  an  error  of  0-01  in  the  determination  of  0?  Also  show 
that  an  error  of  0-01  in  the  determination  of  0  affects  M  to  an  extent  of 
-  3-39,  while  an  error  of  -01  in  the  determination  of  w  only  affects  M  to  the 
extent  of  0"91.  Hence  show  that  it  is  not  necessary  to  weigh  to  more  than 
8-01  of  a  gram. 

From  (16),  §  162,  page  530,  when  the  effect  of  each  observation 
on  the  final  result  is  the  same,  the  partial  differential  coefficients 


§  16G.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      547 
are  all  equal.     If  u  denotes  the  sum  of  n  observations,  av  a2t . . .,  an. 

But,  in  order  that  the  actual  errors  affecting  each  observation  may 
be  the"  same,  we  must  have,  from  (17),  page  530, 

dax  =  da2  =  . . .  =  dan  =  -^ ;         .        .        (4) 

with  the  fractional  errors  : 

daY    .  da2  dan  __du     1 

*  *   u    '  '    u  ^  •  •  •  =  Ijf  "  u  '  Jn     '         * 

Examples.— (1)   Suppose  the  greatest  allowable  fractional  -  error  in    Q 
(preceding  examples)  is  0-5  °/0,  what  is  the  greatest  percentage  error  in  each 
of  the  variables  G,  B,  t,  allowable  under  equal  effects  ?     Here, 
dC  _  dB  _  dt  _  0  005 
G~  B  *  t  ~    v'3  ' 
Ansr.  0-22  for  B  and  t ;  and  0-11  °/0  for  G. 

(2)  If  a  volume  v  of  a  given  liquid  flows  from  a  long  capillary  tube  of 

radius  r  and  length  I'm  t  seconds,  the  viscosity  of  the  liquid  is  t\  =  ttpr^tl'&vl, 

where  p  denotes  the  excess  of  the  pressure  at  the  outlet  of  the  tube  over 

atmospheric  pressure.     What  would  be  the  errors  drt  dv,  dl,  dt,  dp,  necessary 

under  equal  effects  to  give  r\  with  a  precision  of  0*1  °/0  ?    Here, 

dp      dt      Adr         dv         dl     0-001      _>/MWVir 

-=-r  =  4:—  = =  -  —  =  — —  =  0-0004:5. 

p        t         r  v  I        s/b 

It  is  now  necessary  to  know  the  numerical  values  of  p,  t,  v,  r,  I,  before 

dp,  dt,.. .  can  be  determined.     Thus,  if  r  is  about  2  mm.,  the  radius  must 

be  measured  to  the  0*00022  mm.  for  an  error  of  0*1  %  in  t\.     It  has  been  shown 

how  the  best  working  conditions  may  be  determined  by  a  study  of  the  formula, 

to  which  the  experimental  results  are  to  be  referred.    The  following  is  a  more 

complex  example. 

(3)  The  resistance  X  of  a  cell  is  to  be  measured.  Let  Gx,  G2  respectively 
denote  the  currents  produced  by  the  cell  when  working  through  two  known 
external  resistances  rx  and  r2,  and  let  Bx,  B2  be  the  total  resistances  of  the 
circuit,  E  the  electromotive  force  of  the  cell  is  constant.  Your  text-book  on 
practical  physics  will  tell  you  that 

y  _  C2r2  -  Cxrx (6) 


Ox  -  C2 

What  ratio  Cx :  C2  will  furnish  the  best  result  ?    As  usual,  by  partial  differ- 
entiation, (4)  above,  g 

<^2=(!^H!dc<)-  •  •  •  p> 

Find  values  for  ?)Xj'dCx  and  dXjdC2  from  (6)  ;  and  put  Bx  for  rx ;  B2  for  r2. 

From  Ohm's  law,  E  =  CB,  E  being  constant,  Gx  :  G2  =  B2:  Bv  Thus 
?>X  C2(r2  -  rx)  _      _BSB±      ,  dX       Gx(r2  -  rx)  _       BXB* 

Wx  =  "  (Cx  -  C2Y  "     E(B2-BX)>  VC2~  (Cx-C2)*      E(B2-Bx)'Kf 

Substitute  this  result  in  (7). 


548  HIGHER  MATHEMATICS.  §  167. 

(i)  If  a  mirror  galvanometer  is  used,  dCx  —  dC2  =  dC  (say)  =  constant. 

.  WRf  +  RfRj)  (dcy  _  iy(^  +  ^)  {dcf  /Q. 

by  substituting  x  =  R2  :  Rv     For  a  minimum  error,  we  have,  by  the  usual 
method, 

d  (    x4  -x*     \ 

d^\x^r^TJJ  =  ° ;  •'•  x*  ~  2x"  ~  1  =  ° ;  •*•  *  =  2'2  approx- 

Or,  R2  =  2-2Rx ;  or,  Gx  -  2*2C2,  from  Ohm's  law.     Substitute  this  value  of  x 
in  (9),  and  we  get 

j7QR*.dO\ 
dX== E »  ....       (10) 

which  shows  that  the  external  resistance,  Rlt  should  be  as  small  as  is  consistent 
with  the  polarization  of  the  battery. 

(ii)  If  a  tangent  galvanometer  is  used,  dG\G  is  constant.  The  above 
method  will  not  work.  Hence  substitute  Cx  =  ERX  and  C2  =  ER2  in  the  first 
of  equations  (8),  we  get 

R1      R2 
From  this  it  can  be  shown  there  ic  no  best  ratio  R2  :  Rv     From  the  last  ex- 
pression we  can  see  that  the  error  dX  decreases  as  Rx  diminishes,  and  as  R2 
increases.     Hence  R2  should  be  made  as  large  and  RY  as  small  as  is  consistent 
with  the  range  of  the  galvanometer  and  the  polarization  of  the  battery. 

You  can  easily  get  the  fractional  errors  in  each  case.  From  (10)  and 
(11)  respectively 

dG1=JL_    X  dX,  <W_1    X_  dX 
G1  ~>j26'Rl'  X'    G  "2'iV  X' 
assuming  in  the  latter  case  that  Gx  :  C2  =  3  :  1  ;  so  that  the  intermediate 
step  from  (11)  is  dX  =  si 2 .  3i?12/(3E1  -  Ex)  x  dC/G. 

§  167.  Observations  of  Different  Degrees  of  Accuracy. 

Hitherto  it  has  been  assumed  that  the  individual  observations 
of  any  particular  series,  are  equally  reliable,  or  that  there  is  no 
reason  why  one  observation  should  be  preferred  more  than  an- 
other. As  a  general  rule,  measurements  made  by  different 
methods,  by  different  observers,  or  even  by  the  same  observer  at 
different  times,1  are  not  liable  to  the  same  errors.     Some  results 

1 1  am  reminded  that  Dumas,  discussing  the  errors  in  his  great  work  on  the  gravi- 
metric composition  of  water,  alluded  to  a  few  pages  back,  adds  the  remarks  :  "The 
length  of  time  required  for  these  operations  compelled  me  to  prolong  the  work  far  into 
the  night,  generally  finishing  with  the  weighings  about  2  or  3  o'clock  in  the  morning. 
This  may  be  the  cause  of  a  substantial  error,  for  I  dare  not  venture  to  assert  that  such 
weighings  deserve  as  much  confidence  as  if  they  had  been  performed  under  more 
favourable  conditions  and  by  an  observer  not  so  worn  out  with  fatigue,  the  inevitable 
result  of  fifteen  to  twenty  hours  continued  attention." 


§  167.    PROBABILITY  AND  THE  THEORY  OF  ERRORS.      549 

are  more  trustworthy  than  others.  In  order  to  fix  this  idea, 
suppose  that  twelve  determinations  of  the  capacity  of  a  flask  by 
the  same  method,  gave  the  following  results  :  six  measurements 
each  1-6  litres  ;  four,  1*4  litres  ;  and  two,  1-2  litres.  The  numbers 
6,  4,  2,  represent  the  relative  values  of  the  three  results  1-6,  1*4. 
1-2,  because  the  measurement  1*6  has  cost  three  times  as  much 
labour  as  1*2.  The  former  result,  therefore,  is  worth  three  times 
as  much  confidence  as  the  latter.  In  such  cases,  it  is  customary 
to  say  that  the  relative  practical  value,  or  the  weight  of  these  three 
sets  of  observations,  is  as  6  :  4  :  2,  or,  what  is  the  same  thing,  as 
3:2:1^.  In  this  sense,  the  weight  of  an  observation,  or  set  of 
observations,  represents  the  relative  degree'  of  precision  of  that 
observation  in  comparison  with  other  observations  of  the  same 
quantity.  It  tells  us  nothing  about  the  absolute  precision,  h>  of 
the  observations. 

It  is  shown  below  that  the  weight  of  an  observation  is,  in 
theory,  inversely  as  its  probable  error ;  in  practice,  it  is  usual  to 
assign  arbitrary  weights  to  the  observations.  For  instance,  if  one 
observation  is  made  under  favourable  conditions,  another  under 
adverse  conditions,  it  would  be  absurd  to  place  the  two  on  the 
same  footing.  Accordingly,  the  observer  pretends  that  the  best 
observations  have  been  made  more  frequently.  That  is  to  say, 
if  the  observations  av  a2,  .  .  .,  an,  have  weights  pv  p2,  .  .  .,  pn, 
respectively,  the  observer  has  assumed  that  the  measurement  aY 
has  been  repeated  px  times  with  the  result  av  and  that  an  has  been 
repeated  pn  times  with  the  result  an. 

To  take  a  concrete  illustration,  Morley :  has  made  three  accurate 
series  of  determinations  of  the  density  of  oxygen  gas  with  the 
following  results : — 

I.  1-42879  ±  0000034 ;  II.  1-42887  ±  0-000048 ; 
III.  1-42917  ±  0-000048. 

The  probable  errors  of  these  three  means  would  indicate  that 
the  first  series  were  worth  more  than  the  second.  For  experimental 
reasons,  Morley  preferred  the  last  series,  and  gave  it  double  weight. 
In  other  words,  Morley  pretended  that  he  had  made  four  series  of 
experiments,  two  of  which  gave  1*42917,  one  gave  1-42879,  and  one 


lE.  Morley,  "On  the  densities  of  oxygen  and  hydrogen  and  on  the  ratio  of  their 
atomic  weights,"  Smithsonian  Contributions  to  Knowledge,  No.  980,  55,  1895. 


550  HIGHER  MATHEMATICS.  §  167. 

gave  1-42887.  The  result  is  that  1*42900,  not  1-42894,  is  given  as 
the  best  representative  value  of  the  density  of  oxygen  gas. 

The  product  of  an  observation  or  of  an  error  with  the  weight  of 
the  observation,  is  called  a  weighted  observation  in  the  former 
case,  and  a  weighted  error  in  the  other. 

The  practice  of  weighting  observations  is  evidently  open  to 
some  abuse.  It  is  so  very  easy  to  be  influenced  rather  by  the  differ- 
ences of  the  results  from  one  another,  than  by  the  intrinsic  quality 
of  the  observation.     This  is  a  fatal  mistake. 

I.  The  best  value  to  represent  a  number  of  observations  of  equal 
weight,  is  their  arithmetical  mean.  If  P  denotes  the  most  probable 
value  of  the  observed  magnitudes  av  a2,  .  .  .  an,  then  P  -  avP-  a2, 
.  .  .,  P  -  an,  represent  the  several  errors  in  the  n  observations. 
From  the  principle  of  least  squares  these  errors  will  be  a  minimum 
when 

(P  -  a2)2  +  (P  -  a2)2  +  .  .  .  +  (P  -  an)n  =  a  minimum. 

Hence,  from  the  regular  method  for  finding  minimum  values, 

p  =  a1  +  a2  +  ...  +  an 

n  *  ^  ' 

or  the  best  representative  value  of  a  given  series  of  measurements  of 
an  unknown  quantity,  is  an  arithmetical  mean  of  the  n  observations, 
provided  that  the  measurements  have  the  same  degree  of  confidence. 

II.  The  best  value  to  represent  a  number  of  observations  of 
different  weight,  is  obtained  by  multiplying  each  observation  by  its 
weight  and  dividing  the  sum  of  these  products  by  the  sum  of  their 
different  weights.  With  the  same  notation  as  before,  let  pv  p2i . . ., 
pn,  be  the  respective  weights  of  the  observations  av  a2, . . .,  an. 
From  the  definition  of  weight,  the  quantity  ax  may  be  considered 
as  the  mean  of  px  observations  of  unit  weight ;  a2  the  mean  of  p2 
observations  of  unit  weight,  etc.  The  observed  quantities  may, 
therefore,  be  resolved  into  a  series  of  fictitious  observations  all  of 
equal  weight.  Applying  the  preceding  rule  to  each  of  the  resolved 
observations,  the  total  number  of  standard  observations  of  unit 
weight  will  be  px  +  p2  +  . . .  +  pn ;  the  sum  of  the  p1  standard 
observations  of  unit  weight  will  be  pxax ;  the  sum  of  p2  standard 
observations,  p2a2,  etc.  Hence,  from  (1),  the  most  probable  value 
of  a  series  of  observations  of  different  weights  is 

p,  m  Pi^i  +  p2a2  +  . . .  +  pnan 

Pl  +  p2  +  . . .  +  pn  W 


§  167.   PROBABILITY  AND  THE  THEORY  OF  ERRORS.       551 

Note  the  formal  resemblance  between  this  formula  and  that  for 
finding  the  centre  of  gravity  of  a  system  of  particles  of  different 
weights  arranged  in  a  straight  line. 

Weighted  observations  are,  therefore,  fictitious  results  treated 
as  if  they  were  real  measurements  of  equal  weight.  With  this 
convention,  the  value  of  P'  in  (2)  is  an  arithmetical  mean  some- 
times called  the  general  or  probable  mean. 

III.  The  weight  of  an  observation  is  inversely  as  the  square  of 
its  probable  error.  Let  a  be  a  set  of  observations  whose  probable 
error  is  R  and  whose  weight  is  unity.  Let  pl9  p2, . . .,  pn  and  rv 
r2,  .  . .,  rn,  be  the  respective  weights  and  probable  errors  of  a  series 
of  observations  av  a2, . . .,  an,  of  the  same  quantity.  By  definition 
of  weight,  ax  is  equivalent  to  px  observations  of  equal  weight.  From 
(17),  page  530, 

R  R*  R2  111        " 

Examples. — (1)  If  n  observations  have  weights^,  jp2, . .  .,pn,  show  that 

B=±ik) w 

Differentiate  (2)  successively  with  respect  to  Oj,  <%,  . . .  and  substitute  the 
results  in  (16),  page  530. 

(2)  Show  that  the  mean  error  of  a  series  of  observations  of  weights,  pv  p2, 
•  ..,JP«,  is  

M=  +     /    2(jxc2) 

!(n-l)2(p)' 

Hint.  Proceed  as  in  §  161  but  use  px2  and  pv2  in  place  of  x2  and  v2  respectively. 
If  the  sum  of  the  weights  of  a  series  of  observations  is  2(p)=40,  and  the  sum 
of  the  products  of  the  weights  of  each  observation  with  the  square  of  its 
deviation  from  the  mean  of  nine  observations  is  2  [px2)  =0*3998,  show  that 
M  =  ±  0-035. 

(3)  The  probable  errors  of  four  series  of  observations  are  respectively  1*2, 
0'8,  0*9,  1*1,  what  are  the  relative  weights  of  the  corresponding  observations  ? 
Ansr.  7:16:11:8.     Use  (3). 

(4)  Determinations  of  the  percentage  amount  of  copper  in  a  sample  of 
malachite  were  made  by  a  number  of  chemical  students,  with  the  following 
results :  (1)  39*1  ;  (2)  38-8,  38-7,  38'6 ;  (3)  39-9,  391,  39-3  ;  (4)  37*7,  37-9.  If 
these  analyses  had  an  equal  degree  of  confidence,  the  mean,  38*8,  would  best 
represent  the  percentage  amount  of  copper  in  the  ore — formula  (1).  But  the 
analyses  are  not  of  equal  value.  The  first  was  made  by  the  teacher.  To  this 
we  may  assign  an  arbitrary  weight  10.  Sets  (2)  and  (3)  were  made  by  two 
different  students  using  the  electrolytic  process.  Student  (2)  was  more  ex- 
perienced than  student  (3),  in  consequence,  we  are  led  to  assign  to  the  former 
an  arbitrary  weight  6,  to  the  latter,  4.  Set  (4)  was  made  by  a  student  pre- 
cipitating the  copper  as  CuS,  roasting  and  weighing  as  CuO.     The  danger 


552  HIGHER  MATHEMATICS.  §  167. 

of  loss  of  CuS  by  oxidation  to  CuS04  during  washing,  leads  us  to  assign  to 
this  set  of  results  an  arbitrary  weight  2.  From  these  assumptions,  show  that 
38*91  best  represents  the  percentage  amount  of  copper  in  the  ore.  For  the 
sake  of  brevity  use  values  above  37  in  the  calculation.     From  formula  (2), 

-£■  =  1*91.  Add  37  for  the  general  mean.  It  is  unfortunate  when  so  fantastic 
a  method  has  to  be  used  for  calculating  the  most  probable  value  of  a  "  constant 
of  Nature,"  because  a  redetermination  is  then  urgently  required. 

(5)  H.  A.  Rowland  (Proc.  Amer.  Acad.,  15,  75,  1879)  has  made  an  exhaus- 
tive study  of  Joule's  determinations  of  the  mechanical  equivalent  of  heat,  and 
he  believes  that  Joule's  several  values  have  the  weights  here  appended  in 
brackets :  442-8  (0) ;  427*5  (2) ;  426*8  (10) :  428*7  (2) ;  429*1  (1) ;  428-0  (1) ; 
425-8  (2) ;  428-0  (3)  ;  427*1  (3) ;  426*0  (5)  ;  422-7  (1) ;  426*3  (1).  Hence  Kowland 
concludes  that  426-9  best  represents  the  result  of  Joule's  work.  Verify 
this.  Notice  that  Rowland  rejects  the  number  442-8  by  giving  it  zero 
weight. 

(6)  Encke  gives  the  8-60816"  +  0*037  as  the  value  of  the  solar  parallax ; 
D.  Gill  gives  8*802"  +  0*005.  Hence  the  merit  of  Encke's  work  is  to  the 
merit  of  Gill's  work,  as  (0-005)2 :  (0*037)2=25 :  1369  =  1 :  54-76.  Or  £54*76  may 
be  bet  in  favour  of  Gill's  number  against  £1  in  favour  of  Encke. 

IV.  To  combine  several  arithmetical  means  each  of  which  is 
affected  with  a  known  probable  (or  mean)  error,  into  one  general 
mean.     One  hundred  parts  of  silver  are  equivalent  to 

49*5365  ±  0*013  of  NH4C1,  according  to  Pelouze  ; 
49*523    +00055         „  „  Marignac ; 

49-5973  +  0-0005         „  „  Stas  (1867) ; 

49-5992  +  0*00039       „  „  Stas  (1882), 

where  the  first  number  represents  the  arithmetical  mean  of  a  series 
of  experiments,  the  second  number  the  corresponding  probable 
error.  How  are  we  to  find  the  best  representative  value  of  this 
series  of  observations  ?  The  first  thing  is  to  decide  what  weight 
shall  be  assigned  to  each  result.  Individual  judgment  on  the 
"  internal  evidence  "  of  the  published  details  of  the  experiments 
is  not  always  to  be  trusted.  Nor  is  it  fair  to  assign  the  greatest 
weight  to  the  last  two  values  simply  because  they  are  by  Stas. 

L.  Meyer  and  K.  Seubert,  in  a  paper  Die  Atomgewichte  der 
Elemente,  aus  der  Originalzahlen  neu  berechnet,  Leipzig,  1883, 
weighted  each  result  according  to  the  mass  of  material  employed 
in  the  determination.  They  assumed  that  the  magnitude  of  the 
errors  of  observation  were  inversely  as  the  quantity  of  material 
treated.  That  is  to  say,  an  experiment  made  on  20  grams  of 
material  is  supposed  to  be  worth  twice  as  much  as  one  made  on 
10  grams.     This  seems  to  be  a  somewhat  gratuitous  assumption. 


§  167.   PKOBABILITY  AND  THE  THEOKY  OF  ERKORS.      553 

One  way  of  treating  this  delicate  question  is  to  assign  to  each 
arithmetical  mean  a  weight  inversely  as  the  square  of  its  mean 
error.  F.  W.  Clarke  in  his  "  Recalculation  of  the  Atomic  Weights," 
Smithsonian  Miscellaneous  Collections,  1075,  1897,  employed  the 
probable  error.  Although  this  method  of  weighting  did  not  suit 
Morley  in  the  special  case  mentioned  on  page  549,  Clarke  con- 
siders it  a  safe,  though  not  infallible  guide.  Let  A,  B,  C, . . .,  be 
the  arithmetical  mean  of  each  series  of  experiments ;  a,  b,  c,. . ., 
the  respective  probable  (or  mean)  errors,  then,  from  (2), 

A      -H      £. 

General  Mean  =  ;       .  .  (5) 


1 

a*  + 

1 

+ 

1 
1 

+ ... 

Vs? 

+ 

1 
9 

+ 

1 

Probable  Error  =  ±     i  .  .  (6) 


Examples.— (1)  From  the  experimental  results  just  quoted,  show  that 
the  best  value  for  the  ratio 

Ag :  NH4C1  is  100 :  49-5983  ±  0-00031. 

Hint.  Substitute  A=  49-5365, a  =  0-013 ;  B  -  49-523,  b  =  0-0055  ;  G  =  49-5973, 
c  =  00005 ;  D  =  49-5992,  d  =  0-00039,  in  equations  (5). 

(2)  The  following  numbers  represent  the  most  trustworthy  results  yet  pub- 
lished for  the  atomic  weight  of  gold  (H=l) :  195-605+0-0099 ;  195-711  ±00224  ; 
195-808+0-0126;  195-624+0-0224;  195-896+0-0131;  195  -770 +0-0082.  Hence 
show  that  the  best  representative  value  for  this  constant  is  196-743  +  0-0049. 

(3)  In  three  series  of  determinations  of  the  vapour  pressure  of  water 
vapour  at  0°  Regnault  found  the  following  numbers  : 

I.  4-54;  4-54;  4-52;  4-54;  4-52;  4-54;  4-52;  4-50;  4-50;  4-54. 
II.  4-66;  4-67;  4-64;  4-62;  4'64 ;  4-66;  4-67;  4-66;  4-66. 
III.  4-54 ;  4-54 ;  4*54 ;  4-58 ;  4-58 ;  4-57  ;  4-58. 
Show  that  the  best  representative  value  of  series  I.  is  4*526,  with  a  probable 
error  +  0-0105 ;  series  II.,  4-653,  probable  error  +  0-0105 ;  series  III.,  4*561, 
probable  error  +  0-0127.     The  most  probable  value  of  the  vapour  pressure  of 
aqueous  vapour  at  0°  is,  therefore,  4*582,  with  an  equal  chance  of  its  possess- 
ing an  error  greater  or  less  than  0*0064. 

As  a  matter  of  fact  the  theory  of  probability  is  of  little  or 
no  importance,  when  the  constant,  or  systematic  errors  are  greater 
than  the  accidental  errors.  Still  further,  this  use  of  the  probable 
error  cannot  be  justified,  even  when  the  different  series  of  ex- 
periments are  only  affected  with  accidental  errors,  because  the 
probable  error   only  shows  how  unifobmly  an  experimenter  has 


554  HIGHER  MATHEMATICS.  §  167. 

conducted  a  certain  process,  and  not  how  suitable  that  process  is  for 
the  required  purpose.  In  combining  different  sets  of  determina- 
tions it  is  still  more  unsatisfactory  to  calculate  the  probable  error 
of  the  general  mean  by  weighting  the  individual  errors  according 
to  Clarke's  criterion  when  the  probable  errors  differ  very  consider- 
ably among  themselves.  For  example,  Clarke  (Z.c,  page  126) 
deduces  the  general  mean  136-315  ±  0*0085  for  the  atomic  weight 
of  barium  from  the  following  results  : 

136-271  ±  0-0106  ;  136-390  ±  0-0141 ;  135-600  ±  0-2711 ; 
136-563  ±  0-0946. 
The  individual  series  here  deviate  from  the  general  mean  more 
than  the  magnitude  of  its  probable  error  would  lead  us  to  suppose. 
The  constant  errors,  in  consequence,  must  be  greater  than  the 
probable  errors.  In  such  a  case  as  this,  the  computed  probable 
error  ±  0-0085  has  no  real  meaning,  and  we  can  only  conclude 
that  the  atomic  weight  of  barium  is,  at  its  best,  not  known  more 
accurately  than  to  five  units  in  the  second  decimal  place.1 

V.  Mean  and  probable  errors  of  observations  of  different  degrees 
of  accuracy.  In  a  series  of  observations  of  unequal  weight  the 
mean  and  probable  errors  of  a  single  observation  of  unit  weight 
are  respectively 

The  mean  of  a  series  of  observations  of  unequal  weight  has  the 
respective  mean  and  probable  errors 


m = *  V(»  -vkp)  ->™iB-±  °6745  Vf 


2(?n>2) 


l)3(j>)-     ™ 

Example. — A.n  angle  was  measured  under  different  conditions  fourteen 
times.  The  observations  all  agreed  in  giving  4°  15',  but  for  seconds  of  arc 
the  following  values  were  obtained  (the  weight  of  each  observation  is  given  in 
brackets) :  45"-00  (5) ;  31"-25  (4) :  42"-50  (5) ;  45"-00  (3) ;  37"-50  (3) ;  38"-33  (3) ; 
27"-50  (3) :  43"-33  (3) :  40"-63  (4) ;  36"-25  (2) ;  42"-50  (3) ;  39"-17  (3) ;  45"-00  (2) ; 
40"-83  (3).  Show  that  the  mean  error  of  a  single  observation  of  unit  weight 
s  +  9"-475,  the  mean  error  of  the  mean  39"-78  is  l"-397.  Hint.  2(p)  =  46 
2(pv)a  =  1167-03 ;  n  =  14 ;  2(pa)  =  1830-00. 

The  mean  and  probable  errors  of  a  single  observation  of  weight 
p  are  respectively 

-■±V^-"-±°Wi^-   (9) 

1 W.  Ostwald'smYigweon  Clarke's  work  (I.e.) in  the  Zeit.  phys.  Cheni.,  23, 187,1897. 


§  168.    PKOBABILITY  AND  THE  THEORY  OF  ERRORS.      555 

Example. — In  the  preceding  examples  show  that  the  mean  error  of  an 
observation  of  weight  (2)  is  ±  6"-70;  of  weight  (3)  is .+  5"-47;  of  weight  (4) 
±  4"-74 ;  and  of  weight  (5)  ±  4"-24. 

VI.  The  principle  of  least  squares  for  observations  of  different 
degrees  of  precision  states  that  "the  most  probable  values  of  the 
observed  quantities  are  those  for  which  the  sum  of  the  weighted 
squares  of  the  errors  is  a  minimum,"  that  is, 


PiW  +  JPa  V  +  •  •  •  +  VnV, 


minimum. 


An  error  v  is  the  deviation  of  an  observation  from  the  arithmetical 
mean  of  n  observations ;  a  "weighted  square"  is  the  product  of 
the  weight,  p,  and  the  square  of  an  error,  v. 

§  168.  Observations  Limited  by  Conditions. 

On  adding  up  the  results  of  an  analysis,  the  total  weight  of  the 
constituents  ought  to  be  equal  to  the  weight  of  the  substance  itself ; 
the  three  angles  of  a  plane  triangle  must  add  up  to  exactly  180° ; 
the  sum  of  the  three  triangles  of  a  spherical  triangle  always  equal 
180°  +  the  spherical  excess ;  the  sum  of  the  angles  of  the  nor- 
mals on  the  faces  of  a  crystal  in  the  same  plane  must  equal  360°. 
Measurements  subject  to  restrictions  of  this  nature,  are  said  to 
be  conditioned  observations.  The  number  of  conditions  to  be 
satisfied  is  evidently  less  than  the  number  of  unknown  quantities, 
i.e.,  observations,  otherwise  the  value  of  the  unknown  could  be 
deduced  from  the  conditions,  without  having  recourse  to  measure- 
ment. 

In  practice,  measurements  do  not  come  up  to  the  required 
standard,  the  percentage  constituents  of  a  substance  do  not  add 
up  to  100  ;  the  angles  of  a  triangle  are  either  greater  or  less  than 
180°.  Only  in  the  ideal  case  of  perfect  accuracy  are  the  conditions 
fulfilled.  It  is  sometimes  desirable  to  find  the  best  representative 
values  of  a  number  of  imperfect  conditioned  observations.  The 
method  to  be  employed  is  illustrated  in  the  following  examples. 

Examples. — (1)  The  analysis  of  a  Compound  gave  the  following  results : 
37*2  0/o  of  carbon,  44-1  %  of  hydrogen,  19*4  °/0  of  nitrogen.  Assuming  each 
determination  is  equally  reliable,  what  is  the  best  representative  value  of  the 
percentage  amount  of  each  constituent  ?  Let  C,  H,  N,  respectively  denote 
the  percentage  amounts  of  carbon,  hydrogen,  and  nitrogen  required,  then 
C  +  H  =  100- Nee  100-0- 19-4  =  80-6.  .-.  2C  +  H  =  117-8;  C  +  2H  =  124-7. 
Solve  the  last  two  simultaneous  equations  in  the  usual  way.  Ansr.  C  =  36*97  °/0 ; 
H=43-86  °/0;  N  =  19-17  %.      Note  that  this  result  is  quite  independent  of 


556  HIGHER  MATHEMATICS.  §  168. 

any  hypothesis  as  to  the  structure  of  matter.  The  chemical  student  will 
know  a  better  way  of  correcting  the  analysis.  This  example  will  remind  us 
how  the  atomic  hypothesis  introduces  order  into  apparent  chaos.  Some 
analytical  chemists  before  publishing  their  results,  multiply  or  divide  their 
percentage  results  to  get  them  to  add  up  to  100.  In  some  cases,  one  consti- 
tuent is  left  undetermined  and  then  calculated  by  difference.  Both  practices 
are  objectionable  in  exact  work. 

(2)  The  three  angles  of  a  triangle  A,  B,  G,  were  measured  with  the  result 
that  ^  =  51°;  5=94°  20';  0=34°  56'.  Show  that  the  most  probable  values 
of  the  unknown  angles  are  4  =  51°  56' ;  5=94°  15' ;  0  =  34°  49. 

(3)  The  angles  between  the  normals  on  the  faces  of  a  cubic  crystal  were 
found  to  be  respectively  a  =  91°  13' ;  0  =  89°  47' ;  y  =  91°  15' ;  8  =  89°  42'. 
What  numbers  best  represent  the  values  of  the  four  angles  ?  Ansr.  a  =  90° 
43'  45" ;  £  =  89°  17'  45"  ;  y  =  90°  0'  45" ;  8  =  89°  57'  45". 

(4)  The  three  angles  of  a  triangle  furnish  the  respective  observation 
equations  :  A  =  36°  25'  47"  ;  B  =  90°  36'  28" ;  C  =  52°  57'  57" ;  the  equation 
of  condition  requires  that  A  +  B  +  O  -  180°  =  0.  Let  xv  x2t  x3,  respectively 
denote  the  errors  affecting  A  ,B,  G,  then  we  must  have 

x1  +  x2+xs=  - 12 (1) 

I.  If  the  observations  are  equally  trustworthy,  x1  =  x2  =  x3  =  k,  say.  Sub- 
stitute this  value  of  xlf  x2  x3,  in  (1),  and  we  get  3k  +  12  =  0;  or,  k  =  -  4 ; 

.-.  A  =  36°  25'  43"  ;  B  =  90°  36'  24"  ;  G  =  53°  57'  53". 
The  formula  for  the  mean  error  of  each  observation  is 


±Vu 


*W 


w  +  q 

where  w  denotes  the  number  of  unknown  quantities  involved  in  the  n  ob 
servation  equations ;  q  denotes  the  number  of  equations  of  condition  to  be 
satisfied.  Consequently  the  w  unknown  quantities  reduce  to  w  -  q  inde- 
pendent quantities.  2(v2)  denotes  the  sum  of  the  squares  of  the  differences 
between  the  observed  and  calculated  values  of  A,  B,  G.  Hence,  the  mean 
error  =  ±  n/T8  =  ±  6"-93. 

II.  If  the  observations  have  different  weights.  Let  the  respective  weights 
of  A,  B,  C,  be  px  =  4 ;  p2  =  2  ;  ps  =  3.  It  is  customary  to  assume  that  the 
magnitude  of  the  error  affecting  each  observation  will  be  inversely  as  its 
weight.  (Perhaps  the  reader  can  demonstrate  this  principle  for  himself.) 
Instead  of  xx  =  x2  =  x3  =  k,  therefore,  we  write  xx  =  J/c ;  x2  =  \k ;  x3  =  %k. 
From  (1),  therefore,  13k  +  144  =  0 ;  ft=  - 11*07  ;  xx  =  -  2" '11 ;  x2  =  -  5"'54 ; 
x&  =  -  3" -69. 


/    2{pv2) 
m  =  Mean  error  =  +  V — 


w  +  q 

orm=+  11*52.     The  mean  errors  mv  m2,  w3,  respectively  affecting  a,  6,  c,  are 
m  m  m 

mi=±_.  m2=±  _.  TO3=±__. 

Hence 

A  =  36°  25'  44//-23±5"-76 ;  £=90°  36'  22"-46±8"-15 ;  C=52°  57'  53"-31±6"-65. 
It  is,  of  course,  only  permissible  to  reduce  experimental  data  in 


§  169.    PKOBABTLITY  AND  THE  THEOEY  OF  ERROKS.       557 

this  manner  when  the  measurements  have  to  be  used  as  the  basis 
for  subsequent  calculations.  In.  every  case  the  actual  measure- 
ments must  be  stated  along  with  the  "  cooked  "  results. 

§  169.  Gauss'  Method  of  Solving  a  Set  of  Linear  Observation 

Equations. 

In  continuation  of  §  108,  page  328,  let  x,  y,  z,  represent  the 
unknowns  to  be  evaluated,  and  let  alt  a2,  .  .  .,  bv  b2,  .  .  .,  clt  c2, 
Bv  B2,  .  .  .,  represent  actual  numbers  whose  values  have  been 
determined  by  the  series  of  observations  set  forth  in  the  following 
observation  equations : 

OjX  +  bxy  +  cxz  =  B1  ;\ 


(1) 


a2x  +  b2y  +  c2z  =  B2 ; 
a^x  +  b3y  +  czz  =  Bz ; 
a4x  +  b±y  +  c40  =  B±.  \ 

If  only  three  equations  had  been  given,  we  could  easily  calculate 
the  corresponding  values  of  x,  y}  z,  by  the  methods  of  algebra,  but 
these  values  would  not  necessarily  satisfy  the  fourth  equation. 
The  problem  here  presented  is  to  find  the  best  possible  values  of 
x,  y,  z,  which  will  satisfy  the  four  given  observation  equations. 
We  have  selected  four  equations  and  three  unknowns  for  the  sake 
of  simplicity  and  convenience.  Any  number  may  be  included  in 
the  calculation.  But  sets  involving  more  than  three  unknowns  are 
comparatively  rare.  We  also  assume  that  the  observation  equa- 
tions have  the  same  degree  of  accuracy.  If  not,  multiply  each 
equation  by  the  square  root  of  its  weight,  as  in  example  (3)  below. 
This  converts  the  equations  into  a  set  having  the  same  degree  of 
accuracy. 

I.  To  convert  the  observation  equations  into  a  set  of  normal 
equations  solvable  by  ordinary  algebraic  'processes.  Multiply  the  first 
equation  by  av  the  second  by  a2,  the  third  by  a3,  and  the  fourth  by 
a4.  Add  the  four  results.  Treat  the  four  equations  in  the  same 
way  with  bv  b2,  b3,  64,  and  with  cv  c2,  c3,  c4.  Now  write,  for  the 
sake  of  brevity, 

[aa\  =  a*  +  a*  +  a*  +  a2 ;  [bb\  =  b2  +  b2  +  V  +  V  \ 

[ab\  =  albl  +  a2b2  +  asb3  +  a^ ;       [ac\  =  aYcY  +  a2c2  +  asc3  +  a4c4  j 

[aB\  =  axBx  +  a2B2  +  azB3  +  a4i?4;  [bB\  =  bYBx  +  b2B2  +  b2B2  +  64i24 ; 

and  likewise  for  [cc]v  [bc]v  [cB]v     The  resulting  equations  are 


558  HIGHER  MATHEMATICS.  §  169. 

[aa\x  +  [ab\y  +  [ac\z  =  [aB\ ;' 

[ab\x  +  [bb\y  +  [bc\z  =  [bB], ;  -       .         .       (2) 

[ac^x  +  [bc\y  +  [cc^z  =  [cB]v  , 

These  three  equations  are  called  normal  equations  (first  set)  in 
x,  y,  z. 

II.  To  solve  the  normal  equations.  We  can  determine  the 
values  of  x,  y}  z,  from  this  set  of  simultaneous  equations  (2)  by 
any  method  we  please,  determinants  (§  179),  cross-multiplication, 
indeterminate  multipliers,  or  by  the  method  of  substitution.1  The 
last  method  is  adopted  here.  Solve  the  first  normal  equation  for 
x,  thus 

x  =  A^ky  -\f£kz  +    \^k.  .  .  (3) 

[aa]i         [aa]x         [aa\x  v  ' 

Substitute  this  value  of  x  in  the  other  two  equations  for  a  second 
set  of  normal  equations  in  which  the  term  containing  x  has  dis- 
appeared. 

(m-  ^*i>f+  (eh  -  $&ty  -  H  -  ^4> 

For  the  sake  of  simplicity,  write 

The  second  set  of  normal  equations  may  now  be  written : 

[bb\y  t  [&>],*  =  [«q, ;\  ... 

[6o]22/  +  [cc\z  =  [ci?]2.  J     '        >        •    -W 
Solve  the  first  of  these  equations  for  y, 

y~     t»].  I+  lbbV        ■        ■  (5) 

Substitute  this  in  the  second  of  equations  (4),  and  we  get  a  third 
set  of  normal  equations, 

(m  -  fUt*].)*  -  ([«bl  -  gj^i). 

1  The  equations  cannot  be  solved  if  any  two  are  identical,  or  can  be  made  identical 
by  multiplying  through  with  a  constant. 


§  169.   PROBABILITY  AND  THE  THEORY  OF  ERRORS.       559 


(6) 


which  may  be  abbreviated  into  [cc]^z{  =  [cB]z.     Hence, 

k*      ;    •    ' 

[bb]2,  [bc]2, . . .,  [cc]3, ...  are  called  auxiliaries.     Equations  (3) 
(5),  (7),  collectively  constitute  a  set  of  elimination  equations  : 
_     [ab]lt       [ac]u   t   [aBl ,  \ 


x  = 


[aa]iy 


y=- 


W 
Mi 

[bb]2 


z  + 


z  + 


z  — 


(7) 


The  last  equation  gives  the  value  of  z  directly;  the  second  gives 
the  value  of  y  when  z  is  known,  and  the  first  equation  gives  the 
value  of  x  when  the  values  of  y  and  z  are  known. 

Note  the  symmetry  of  the  coefficients  in  the  three  sets  of  normal 
equations.  Hence  it  is  only  necessary  to  compute  the  coefficients 
of  the  first  equation  in  full.  The  coefficients  of  the  first  horizontal 
row  and  vertical  column  are  identical.  So  also  the  second  row  and 
second  column,  etc.  The  formation  and  solution  of  the  auxiliary 
equations  is  more  tedious  than  difficult.  Several  schemes  have 
been  devised  to  lessen  the  labour  of  calculation  as  well  as  for  test- 
ing the  accuracy  of  the  work.     These  we  pass  by. 

IV.  The  weights  of  the  values  of  x,  y,  z.  Without  entering 
into  any  theoretical  discussion,  the  respective  weights  of  zf  y,  and  x 
are  given  by  the  expressions : 

r    ^  ihh\  [ca]Jbb]r 


'[ccj. 


\cc\lbb\  -  [bcy[bc]{ 


(8) 


III 


The  mean  errors  affecting  the  values  of  x,  y,  z. 
axx  +  b^tj  +  c^z  -  B1  =  vl ; 
a2x  +  b2y  +  c2z  -  B2  =  v2; 


Let 


Let  M  denote  the  mean  error  of  any  observed  quantity  of  unit 
weight, 


M 
M 


-  \  n  -  io 


n  -  w 


for  equal  weights ; 


for  unequal  weights 


(9) 


where  n  denotes  the  number  of  observation  equations,  w  the  number 


560 


HIGHER  MATHEMATICS. 


§  169. 


of  quantities  x,  y,  z,  .  .  .  Here  w  =  3,  n  =  4.     Let  M„  Myi  M„  re- 
spectively denote  the  mean  errors  respectively  affecting  x,  y,  z. 

M  MM 

Mx=±-^;  My=±—',  Mt=±~JT.    .         (10) 


Jp» 


JPy 


slPz 


Examples. — (1)  Find  the  values  of  the  constants  a  and  b  in  the  formula 

y  =  a  +  bx, (11) 

from  the  following  determinations  of  corresponding  values  of  x  and  y  : — 
y  =  3-5,  5-7  8-2  10-3,  .  .  . ; 

x  =  0,  88         182,  274,.  .  . 

We  want  to  find  the  best  numerical  values  of  a  and  b  in  equation  (11).  Write 
x  for  a,  and  y  for  b,  so  as  to  keep  the  calculation  in  line  with  the  preceding 
discussion.     The  first  set  of  normal  equations  is  obviously 

\aa\x  +  [ab\y  =  [aB\ ;  and  [ab\x  +  \bb\y  =  [bR\. 

"x~      [aaly  +  [aa\ '  *  *  y  ~  [  bb]2' 
Again,  [aa\  =  4 ;  [bb\  =  115,944  ;  [ab\  =  544 ;  \aB\  =  27-7  ;  [bR\  =  4,816-2 ; 
\bb\  =  4,853-67  ;  [6E]2  =  115,951-4.    x  =  3-52475  ;  y  =  0-02500  ;  or,  reconvert- 
ing x  into  a,  and  y  into  b,  (11)  is  to  be  written, 
y  =  3-525  +  0-025z. 


a. 

b. 

Difference  between 

Calculated  and 

Observed. 

Square  of  Difference 

between  Calculated 

and  Observed. 

Calculated. 

Observed. 

0 

88 
182 
274 

3-525 

5-725 

8-075 

10-375 

3-5 

5-7 

8-2 

10-3 

+  0-025 
+  0-025 
-  0-125 
+  0-075 

0-000625 
0-000625 
0-015625 
0-005625 

.  0-0225 

.-.  M  =  ±  0-106.  

Weight  oib=p„  =  [66]2  =  41,960  ;  M„=±  0-106/  \/41,960  =  ±  0*0004. 

Weight  of  a  =  px  =  ^f  =  1-5  ;  Ma=±  0-106/  -s/l-5  =  ±  0-087. 

(2)  The  following  equations  were  proposed  by  C.  F.  Gauss  in  his  Theoria 
motus  corporum  coelestium  (Hamburg,  1809 ;  Gauss'  Werke,  7,  240,  1871)  to 
illustrate  the  above  method : 

x  -    y  +  2z  =  3  ;      4cc  +    y  +  4a  =  21 
3a;  +  2y  -  hz  =  5 ;  -  x  +  Sy  +  3z 
Hence  show  that  x  =  +  2-470  ;  y  =  +  3-551 ;  z  =  +  1-916  ;   2(u2) 
M  =  ±  284 ;  px  =  246  ;  py  =  136  ;  pM  =  539  ;  If*  =  ±  0-057  ;   My  -- 
Mg  =  +  0-039.     Hint.  The  first  set  of  normal  equations  is 

27a;  +  6y  =  88  ;  6x  +  15y  +  e  =  70;  y  +  54a  =  107. 

(3)  The  following  equations  were  also  proposed  by  C.  F.  Gauss  (I.e.)  to 
illustrate  his  method  of  solution ;  x  -  y  +  2z = 3,  with  weight  1 ;  3x  +  2y-5z=5, 


=  21;, 
=  14.  J 


.       (12) 

0-0804  ; 
-  0-077 : 


§  169.   PROBABILITY  AND  THE  THEORY  OF  ERRORS.       561 

with  weight  1 ;  4sc  +  y  +  ±z  =  21,  with  weight  1 ;  -  2a?  +  6y  +  6z  =  28,  with 
weight  £.  By  the  rule,  multiply  the  last  equation  by  \/j  =  £  and  we  get 
set  (12).  Show  that  x  =  +  2*47  with  a  weight  24*6  ;  y  =  +  3*55  with  a  weight 
13*6 ;  and  z  =  +  1*9  with  a  weight  53*9.  It  only  remains  to  substitute  these 
values  of  x,  y,  z,  in  (14)  to  find  the  residuals  v.  Hence  show  that  M  =  ±  295. 
Proceed  as  before  for  Mx,  My,  Mz. 

(4)  The  length,  Z,  of  a  seconds  pendulum  at  any  latitude  L,  may  be  re- 
presented by  A.  0.  Clairaut's  equation :  I,  =  L0  +  A  sin2.L,  where  L0  and  A  are 
constants  to  be  evaluated  from  the  following  observations  : 

L  =  0°0't         18°  27',      48°  24',      58°  15',      67°  4' ; 
I  m  0-990564,  0-991150,  0-993867,  0-994589,  0-995325. 
Hence  show  that  I  =  0-990555  +  0*005679  sin2£.      Hint.   The  normal  equa- 
tions are, 

x  +  0-44765  y  =  0-993099 ;  x  +  070306  y  =  0-994548. 

(5)  Hinds  and  Callum  (Journ.  Amer.  Chem.  Soc,  24,  848,  1902)  represent 
their  readings  of  the  percentage  strength,  y,  of  a  solution  of  iron  with  the 
photometric  readings,  x,  of  the  intensity  of  transmitted  light  by  the  formula 
y(x  +  b)  =  a.     The  readings  were 

x  =3-8,     4-3,     4-7,     5-3,     6-0,     6-7,     7*4,     8-1,     8-7,     9-7; 

y  x  102  =  8-64,  7-57,  6-92,  6-06,  5-28,  4-70,  4-22,  3-79,  3-52,  3-13. 
The   authors   state  that  a  =  0-2955 ;  b  =  0*375.     The  probable   error  of  one 
determination  of  y  is  given  as  0-000034,  or  as  3  parts  in  10,000,000.     Use  (9). 

The  above  is  based  on  the  principle  of  least  squares.  A  quicker 
method,  not  so  exact,  but  accurate  enough  for  most  practical  pur- 
poses, is  due  to  Mayer.  We  can  illustrate  Mayer's  method  by- 
equations  (12). 

First  make  all  the  coefficients  of  x  positive,  and  add  the  results 
to  form  a  new  equation  in  x.  Similarly  for  equations  in  y  and  z. 
We  thus  obtain, 

9#  -  y  -  2z  =  15 ;  5x  +  ly  =  37 ;  x  +  y  +  14s  =  33. 
Solve   this  set  of  simultaneous   equations   by  algebraic  methods 
and  we  get  x  =  2-485;    y  =  3-511;    z  =  1'929.      Compare  these 
values  of  x,  y,  z,  with  the  best   representative  values  for   these 
magnitudes  obtained  in  Ex.  (2),  above. 

V.  Errors  affecting  two  or  more  dependent  observations.  There 
is  a  tendency  in  computing  atomic  weights  and  other  constants  for 
all  the  errors  to  accumulate  upon  the  constant  last  determined.  The 
atomic  weight  of  fluorine  is  obtained  from  the  ratio  :  CaF2 :  CaS04. 
The  calculation  not  only  includes  the  experimental  errors  in  the 
measurement  of  this  ratio,  but  also  the  errors  in  the  atomic  weight 
determinations  of  calcium  and  sulphur.  It  has  been  pointed  out 
by  J.  D.  van  der  Plaats  (Gompt.  Bend.,  116,  1362,  1893)  that  with 
sufficient  experimental  data  the  given  ratio  can  be  made  to  furnish 

NN 


562  HIGHER  MATHEMATICS.  §  169 

three  atomic  weights  over  which  the  errors  of  observation  are 
equally  distributed,  and  not  accumulated  upon  a  single  factor. 
F.  W.  Clarke  (Amer.  Chem.  Journ.,  27,  32,  1902)  illustrates  the 
method  by  calculating  the  seven  atomic  weights :  silver,  chlorine, 
bromine,  iodine,  nitrogren,  sodium  and  potassium — given  O  =  16 ; 
H  =  1-0079 — from  thirty  ratios  arranged  in  the  form  of  thirty 
linear  equations,  thus, 

Ag  :  Br  =  100  :  74-080 ;  .-.  100  Br  =  74-080  Ag ; 
KC103 :  03  =  100  :  39-154  •  .-.  39-154  K  +  39-154  CI  =  2920-608  ; 

These  thirty  linear  equations  are  reduced  to  seven  normal  equa- 
tions as  indicated  above.  By  solving  these,  the  atomic  weights  of 
the  seven  elements  are  obtained  with  the  errors  of  observation 
evenly  distributed  among  them  according  to  the  method  of  least 
squares. 

When  two  observed  quantities  are  afflicted  with  errors  of  ob- 
servation and  it  is  required  to  find  the  most  probable  relation 
between  the  quantities  concerned,  we  can  proceed  as  indicated  in 
the  following  method.     The  observed  quantities  are,  say, 

y  =  0-5,         0-8,         1-0,         1-2; 
x  =  0-4,        0-6,        0-8,        0-9, 

and  we  want  to  find  the  best  representative  values  for  a  and  b  in 
the  equation 

y  =  ax  +  b. 
You  can  get  approximate  values  for  a  and  b  by  the  graphic  method 
of  page  355  ;  or,  take  any  two  of  the  four  observation  equations 
and  solve  for  a  and  b.     Thus,  taking  the  first  and  third, 

0-5  =  0-4a  +  b ;  1-0  m  0'8a  +  b  ;  .-.  a  =  1-25  ;  b  =  0. 
Let  a  and  ft  be  the  corrections  required  to  make  these  values 
satisfy  the  conditions  of   the  problem   in   hand.      The   required 
equation  is,  therefore, 

y  =  (1-25  +  a)x  +p. 

Insert  the  observed  values  of  x  and  y,  so  as  to  form  the  four 
observation  equations  : 

0-5  =  (1-25  +  a)0-4  +  p;  1-0  =  (1-25  +  a)0-8  +  /?; 
0-8  =  (1-25  +  a)0-6  +  p ;  1-2  =  (1-25  +  o)0-9  +  fi; 

From  these  we  get  the  two  normal  equations 


§  170.   PROBABILITY  AND  THE  THEORY  OF  ERRORS.       563 

0-1250  -  2-70a  +  4-0/?;  0-0975  m  l-97a  +  2-7/?. 
.-.  a  =  +  0-089;  p  =  -  0-029. 
And  finally 

a  =  1-25  +  0-089  =  +  1-339 ;  b  =  0-000  -  0-029  =  -  0-029. 
The  best  representative   equation  for  the  above  observations  is 
therefore, 

y  =  l-339z  -  0-029. 
See  A.  F.  Ravenshear,  Nature,  63,  489,  1901.  The  above  method  is  given 
by  M.  Merriman  in  A  Textbook  on  the  Method  of  Least  Squares,  New  York, 
127,  1891 ;  W.  H.  Keesom  has  given  a  more  general  method  in  the  Com- 
munication's from  the  Physical  Laboratory  at  the  University  of  Leiden,  Suppl. 
No.  4,  1902. 

§  170.  When  to  Reject  Suspected  Observations. 

'  There  can  be  no  question  about  the  rejection  of  observations 
which  include  some  mistake,  such  as  a  wrong  reading  of  the 
eudiometer  or  burette,  a  mistake  in  adding  up  the  weights,  or  a 
blunder  in  the  arithmetical  work,  provided  the  mistake  can  be 
detected  by  check  observations  or  calculations.  Sometimes  a 
most  exhaustive  search  will  fail  to  reveal  any  reason  why  some 
results  diverge  in  an  unusual  and  unexpected  manner  from  the 
others.  It  has  long  been  a  vexed  question  how  to  deal  with 
abnormal  errors  in  a  set  of  observations,  for  these  can  only  be 
conscientiously  rejected  when  the  mistake  is  perfectly  obvious. 
It  would  be  a  dangerous  thing  to  permit  an  inexperienced  or 
biassed  worker  to  exclude  some  of  his  observations  simply  because 
they  do  not  fit  in  with  the  majority.  "  Above  all  things,"  said 
S.  W.  Holman  in  his  Discussion  on  the  Precision  of  Measure- 
ments, New  York,  1901,  "  the  integrity  of  the  observer  must  be 
beyond  question  if  he  would  have  his  results  carry  any  weight 
and  it  is  in  the  matter  of  the  rejection  of  doubtful  or  discordant 
observations  that  his  integrity  in  scientific  or  technical  work 
meets  its  first  test.  It  is  of  hardly  less  importance  that  he  should 
be  as  far  as  possible  free  from  bias  due  either  to  preconceived 
opinions  or  to  unconscious  efforts  to  obtain  concordant  results." 

Several  criteria  have  been  suggested  to  guide  the  investigator 
in  deciding  whether  doubtful  observations  shall  be  included  in  the 
mean.  Such  criteria  have  been  deduced  by  W.  Chauvenet,  Hagen, 
Stone,  Pierce,  etc.  None  of  these  tests  however  is  altogether 
satisfactory.     Chauvenet's  criterion  is  perhaps  the  simplest  to 

NN* 


564 


HIGHER  MATHEMATICS. 


§170. 


understand  and  most  convenient   to  use.      It  is  an  attempt  to 
show,  from  the  theory  of    probability,  that  reliable   observations 
will  not  deviate  from  the  arithmetical  mean  beyond  certain  limits. 
We  have  learned  from  (2)  and  (6),  page  523, 

04769 


r  = 


=  0-6745 


If  x  =  rt,  where  rt  represents  the  number  of  errors  less  than  x 
which  may  be  expected  to  occur  in  an  extended  series  of  observa- 
tions when  the  total  number  of  observations  is  taken  as  unity,  r 
represents  the  probable  error  of  a  single  observation.  Any  mea- 
surement containing  an  error  greater  than  x  is  to  be  rejected.  If 
n  denotes  the  number  of  observations  and  also  the  number  of 
errors,  then  nP  indicates  the  number  of  errors  less  than  rt,  and 
w(l  -  P)  the  number  of  errors  greater  than  the  limit  rt.  If  this 
number  is  less  than  J,  any  error  rt  will  have  a  greater  probability 
against  than  for  it,  and,  therefore,  may  be  rejected. 

The  criterion  for  the  rejection  of  a  doubtful  observation  is, 
therefore, 

1        n 


* 


P);  .-vP- 


2n 


l-Me^dt.       (l) 

s/ttjo 


By  a  successive  application  of  these  formulae,  two  or  more  doubt- 
ful results  may  be  tested.  The  value  of  t,  or,  what  is  the  same 
thing,  of  P,  and  hence  also  of  n,  can  be  read  off  from  the  table  of 
integrals,  page  622  (Table  XL).  Table  XII.  contains  the  nu- 
merical value  of  xjr  corresponding  to  different  values  of  n. 

Examples. — (1)  The  result  of  13  determinations  of  the  atomic  weight  of 
oxygen  made  by  the  same  observer  is  shown  in  the  first  column  of  the  sub- 
joined table.  Should  19-81  be  rejected?  Calculate  the  other  two  columns  of 
the  table  in  the  usual  way. 


Observation. 

X. 

A 

Observation. 

X. 

*. 

15-96 

-0-26 

0-0676 

15-88 

-0-34 

0-1156 

19-81 

+  3-59 

12-8881 

15-86 

-0-36 

0-1296 

15-95 

-0-27 

0-0729 

16-01 

-0-21 

0-0441 

15-95 

-0-27 

0-0729 

15-96 

-0-26 

0-0676 

15-91 

-0-31 

0-0961 

15-88 

-0-34 

0-1156 

15-88 

-0-34 

0-1156 

15-93 

-0-29 

0-0841 

15-91 

-0-31 

0-0961 

Mean  of  13  observations 

=  16-22 ;  2(a?2j 

=  13-9659 

§  170.   PROBABILITY  AND  THE  THEORY  OF  ERRORS.       565 

The  deviation  of  the  suspected  observation  from  the  mean,  is  3 "59.  By 
Chauvenet's  criterion,  probable  error  =  r  =  0*7281,  n  =  13.  From  Table  XII., 
x/r  =  3-07,  .*.  x  =  3-07  x  0-7281  =  22-7.  Since  the  observation  19-81  deviates 
from  the  mean  more  than  the  limit  22-7  allowed  by  Chauvenet's  criterion, 
that  observation  must  be  rejected. 

(2)  Should  16*01  be  rejected  from  the  preceding  set  of  observations  ? 
Treat  the  twelve  remaining  after  the  rejection  of  19*81  exactly  as  above. 

(3)  Should*  the  observations  0*3902  and  0*3840  in  F.  Rudberg's  results, 
page  527,  be  retained  ? 

(4)  Do  you  think  203*666  in  W.  Crookes'  data,  page  531,  is  affected  by 
some  "  mistake  "  ? 

(5)  Would  H.  A.  Rowland  have  rejected  the  "  442*8  "  result  in  Joule's 
work,  page  552,  if  he  had  been  solely  guided  by  W.  Chauvenet's  criterion  ? 

(6)  Some  think  that  "  4*88  "  in  Cavendish's  data,  page  527,  is  a  mistake. 
Would  you  reject  this  number  if  guided  by  the  above  criterion  ? 

These  examples  are  given  to  illustrate  the  method  of  applying 
the  criterion.  Nothing  more.  Any  attempt  to  establish  an  arbi- 
trary criterion  applicable  to  all  cases,  by  eliminating  the  knowledge 
of  the  investigator,  must  prove  unsatisfactory.  It  is  very  question- 
able if  there  can  be  a  better  guide  than  the  unbiassed  judgment 
and  common  sense  of  the  investigator  himself.  The  theory  you 
will  remember  is  only  "  common  sense  reduced  to  arithmetic  ". 

Any  observation  set  aside  by  reason  of  its  failure  to  comply 
with  any  test  should  always  be  recorded.  As  a  matter  of  fact,  the 
rare  occurrence  of  abnormal  results  serves  only  to  strengthen  the 
theory  of  errors  developed  from  the  empirical  formula,  y  =  ke~  *2*2. 
There  can  be  no  doubt  that  as  many  positive  as  negative  chance 
deviations  would  appear  if  a  sufficient  number  of  measurements 
were  available.1  "  Every  observation,"  says  G.  L.  Gerling  in  his 
Die  Ausgleichungs-Bechnungen  der  praktischen  Geometrie,  Ham- 
burg, 68,  1843,  "  suspected  by  the  observer  is  to  me  a  witness  of 
its  truth.  He  has  no  more  right  to  suppress  its  evidence  under  the 
pretence  that  it  vitiates  the  other  observations  than  he  has  to  shape 
it  into  conformity  with  the  majority."  The  whole  theory  of  errors 
is  founded  on  the  supposition  that  a  sufficiently  large  number  of 
observations  has  been  made  to  locate  the  errors  to  which  the 
measurements  are  susceptible.  When  this  condition  is  not  ful- 
filled, the  abnormal  measurement,  if  allowed  to  remain,  would 
exercise  a  disproportionate  influence  on  the  mean.     The  result 


i  F.  Y.  Edgeworth  has  an  interesting  paper  "  On  Discordant  Observations  "  in  the 
Phil.  Mag.  [5],  23,  364,  1887. 


566  HIGHER  MATHEMATICS.  §  170. 

would  then  be  less  accurate  than  if  the  abnormal  deviation  had 
been  rejected.  The  employment  of  the  above  criterion  is,  therefore, 
permitted  solely  because  of  the  narrow  limit  to  the  number  of  ob- 
servations. It  is  true  that  some  good  observations  may  be  so  lost, 
but  that  is  the  price  paid  to  get  rid  of  serious  mistakes. 

It  is  perhaps  needless  to  point  out  that  a  suspected  observation 
may  ultimately  prove  to  be  a  real  exception  requiring  further 
research.  To  ignore  such  a  result  is  to  reject  the  clue  to  a  new 
truth.  The  trouble  Lord  Eayleigh  recently  had  with  the  density  of 
nitrogen  prepared  from  ammonia  is  now  history.  The  "  ammonia" 
nitrogen  was  found  to  be  f^th  part  lighter  than  that  obtained 
from  atmospheric  air.  Instead  of  putting  this  minute  "  error  "  on 
one  side  as  a  "  suspect,"  Lord  Eayleigh  persistently  emphasized 
the  discrepancy,  and  thus  opened  the  way  for  the  brilliant  work  of 
W.  Ramsay  and  M.  W.  Travers  on  "  Argon  and  Its  Companions  ". 


CHAPTEE  X. 
THE  CALCULUS  OF  VARIATIONS. 

"  Natura    operatur  per    modos    faciliores    et    expeditiones." — P.    de 
Fermat.1 

§  171.  Differentials  and  Variations. 

Nearly  two  hundred  years  ago  Maupertius  tried  to  show  that 
the  principle  of  least  action  was  one  which  best  exhibited  the 
wisdom  of  the  Creator,  and  ever  since  that  time  the  fact  that 
a  great  many  natural  processes  exhibit  maximum  or  minimum 
qualities  has  attracted  the  attention  of  natural  philosophers.  In 
dealing  with  the  available  energy  of  chemical  and  physical  phen- 
omena, for  example,  the  chemist  seeks  to  find  those  conditions 
which  make  the  entropy  a  maximum,  or  the  free  energy  a  mini- 
mum, while  if  the  problems  are  treated  by  the  methods  of  ener- 
getics, Hamilton's  principle  : 

"  If  a  system  of  bodies  is  at  A  at  the  time  tv  and  at  B  at  the  time 
t2,  it  will  pass  from  A  to  B  by  such  a  path  that  the  mean  value  of  the 
difference  between  the  kinetic  and  potential  energy  of  the  system  in 
the  interval  t2-  tx  is  a  minimum  " 

is  used.  Problems  of  this  nature  often  require  a  more  powerful 
mathematical  tool  than  the  differential  calculus.  The  so-called 
calculus  of  variations  is  used. 

If  it  be  required  to  draw  a  curve  of  a  certain  fixed  length  from 
0  to  A  (Fig.  173)  so  that  the  area  bounded  by  OB,  BA,  and  the 
curve  may  be  a  maximum.  The  inquiry  is  directed  to  the  nature 
of  the  curve  itself.  In  other  words,  we  want  the  equation  of  the 
curve.      This   is   a  very  different    kind   of   problem    from   those 

1  "  Nature  works  by  the  easiest  and  readiest  means." — P.  de  Fermat  in  a  letter  to 
M.  de  la  Chambre,  1662. 

567 


568 


HIGHER  MATHEMATICS. 


§172. 


hitherto  considered  where  we  have  sought  what  special  values 
must  be  assigned  to  certain  variables  in  a  given  expression  in 
order  that  this  function  may  attain  a  maximum  or  minimum 
value. 

Whatever  be  the  equation  of  the  curve,  we  know  that  the  area 
must  be  furnished  by  the  integral  jydx  ;  or  jf(x)dx.  The  problem 
now  before  us  is  to  find  what  must  be  the  form  of  f(x)  in  order 
that  this  integral  may  be  a  maximum.  It  is  easy  to  see  that  if 
the  form  of  the  function  y  =  f(x)  is  variable,  the  value  of  y  can 
change  infinitesimally  in  two  ways,  either 

(i)  By  an  increment  in  the  value  of  the  independent  variable 
x;  or 

(ii)  By  a  change  in  the  form  of  the  function  as  it  passes  from 
the  shape  f(x)  to,  say,  the  shape  0(#) ;  or,  to  be  more  explicit,  say 
from  y  =  sin  x  to,  say,  y  =  tan  x. 

The  first  change  is  represented  by  the  ordinary  differential  dy  j 
the  second  change  is  called  a  variation,  and  is  symbolized,  in 
Lagrange's  notation,  by  8y.     Consequently,  the  differential 

dy  =  f(x  +  dx)  -f(x); 
while  the  variation 

Sy  =  <p(x)  -  f(x).  .         .        .        (1) 

Care  must  be  taken  that  the  symbol  "  8  "  is  only  applied  to  those 
measurements  which  are  produced  by  a 
change  in  the  form  of  the  function.  The 
change,  dy,  is  represented  in  Fig.  173 
by  dy  =  NQ  -  MP ;  the  change  &y  by 
Sy  =  M  F  -MP;  dx  =  MN;  Sx  =  MM'. 
It  is  not  difficult  to  show  from  the  above 
diagram  that  the  symbols  of  differentia- 
tion and  variation  are  interchangeable,  so  that 

dSy  =  My (2) 


Fig.  173. 


§  172.  The  Variation  of  a  Function. 

To  find  the  variation — not  the  differential — of  a  function.  Let 
y  be  the  given  function.  Write  y  +  8y  in  place  of  y,  and  subtract 
the  new  function  from  the  old,  and  there  you  have  it.  We  at  once 
recognize  the  formal  analogy  of  the  operation  with  the  process  of 
differentiation.     Thus,  if 

u  =  yn, 


§  173.  THE  CALCULUS  OF  VARIATIONS.  569 

the  variation  of  u  is 

du 
8u  =  (y+8yy-y  =  ^fy,         .,        .        (3) 

by  Taylor's  theorem,  neglecting  the  higher  order  of  infinitesimals. 
Let  us  adopt  Newton's  notation,  and  write  y  for  dy/dx ;  y  for 
d2y/dx2 ;  . . .,  then,  if 

when  y  changes  to  y  +  8y,  y  becomes  y  +  S#.     Accordingly 


hu  = 


dun         dun,  .         dun  du  n/dy\ 

r/y +  tf»  3  *> 8u  -  Ty*y  +  t^\£)>    w 

■fay 

by  the  extension  of  Taylor's  theorem,  neglecting  the  higher  powers 
of  small  magnitudes.  You  will  remember  that  "<$,''  on  page  19, 
was  used  to  represent  a  small  finite  change  in  the  value  of  the  in- 
dependent variable,  while  here  "8"  denotes  an  infinitesimal 
change  in  the  form  of  the  function. 

To  evaluate  8y,  8y,  hy  you  follow  exactly  the  same  methods. 

.*:?    $(dy\    d(y  +  w    dy-d*y.  ,^_^%.      _ 

ty  ~  \dx)  dx  dx~  dx  >  ddx2       dx2  ""    W 

So  far  as  I  know  the  verb  "  to  variate  "  or  u  to  vary,"  meaning  to 
find  the  variation  of  a  function  in  the  same  way  that  M  to  differ- 
entiate "  means  to  find  the  differential  of  a  function,  is  not  used. 

§  173.  The  Yariation  of  an  Integral  with  Fixed  Limits. 

Let  it  be  required  to  find  the  variation  of  the  integral 

U  =  \lV.dx       ....        (6) 

=f{x>y> %%•-)*> or'  7=  fa y>y>y>-- )•      (7) 

The  value  of  U  may  be  altered  either  by 

(i)  A  change  in  the  limits  xx  and  x0 ;  or, 
(ii)  A  change  in  the  form  of  the  function. 
We  have  already  seen  that  if  the  end  values  of  the  integral  are 
fixed,  any  change  in  the  independent  variable  x  does  not  affect  the 
value  of  U.  Let  us  assume  that  the  limits  are  fixed  or  constant. 
The  only  way  that  the  value  of  U  can  now  change  is  to  change 
the  form  of  V  =  /(. .  .).  But  the  variation  of  V  is  SV,  and,  by  the 
above-mentioned  rule, 


where 


570  HIGHEE  MATHEMATICS.  §  174. 

*^ +  %**%** ^ 

For  the  sake  of  brevity,  let  us  put 

P  =  fy'>  Q  =  d$>  B=df'> (9) 

and  we  get 

Let  us  now  integrate  term  by  term.  We  know  of  old  (A),  page 
205,  that 

,AV  i  \  du  ^ 

so  that  if  we  put  Q  =  u;  dQ  =  du;  dSy  =  dv  ;  v  =$y,  then 

similarly,  by  a  double  application  of  the  method  of  integration  by 
parts,  we  find  that 

and  consequently,  after  substituting  the  last  two  results  in  (10), 
we  get 

*H:(*-S+SW(«-IM*th  <«> 

The  last  two  terms  do  not  involve  any  integrations,  and  depend 
upon  the  form  of  the  function  only.  Let  I0  represent  the  aggregate 
of  terms  formed  when  x0  is  put  for  x ;  and  Ix  the  aggregate  of 
terms  when  x1  is  put  for  x ;  then  (11)  assumes  the  form 

8U  =IX  -  I0  +  \  KByte,    .        .        .       (12) 
where  K  has  been  put  in  place  of  the  series 

s-'-S-g.    •    ■    ■  « 

The  variation  when  the  function  V  includes  higher  derivatives  than 
y,  is  found  in  a  similar  manner. 

§  175.  Maximum  or  Minimum  Values  of  a  Definite  Integral. 

Perhaps  the  most  important  application  of  the  calculus  of  varia- 
tions is  the  determination  of  the  form  of  the  function  involved  in  a 
definite  integral  in  such  a  manner  that  the  integral,  say, 


174.  THE  CALCULUS  OF  VARIATIONS.  571 


V.dx,  .        .        .      (14) 

shall  have  a  maximum  or  a  minimum  value.  In  order  to  find  a 
maximum  or  a  minimum  value  of  a  function,  we  must  find  such  a 
value  of  x  that  a  small  change  in  the  value  of  x  will  produce  a  change 
in  the  value  of  the  function  which  is  indefinitely  small  in  com- 
parison with  the  value  of  x  itself.     We  must  have 

hU  =  0 ;  and  I,  - 10  +  P  KSydx  =  0.  .       (15) 

This  requires  that 

I±  _  IQ  =  0  ;  and  |    KBydx  =  0,        .         .       (16) 

Jxo 

for  if  each  member  did  not  vanish,  each  would  be  determined  by 
the  value  of  the  other.  Since  Sy  is  arbitrary,  the  second  condition 
can  only  be  satisfied  by  making 

dQ      d2B 

Most  of  your  troubles  in  connection  with  this  branch  of  the  calculus 
of  variations  will  arise  from  this  equation.  It  is  often  very  re- 
fractory ;  sometimes  it  proves  too  much  for  us.  The  equation  then 
remains  unsolved.  The  nature  of  the  problem  will  often  show 
directly,  without  any  further  trouble,  whether  it  be  a  maximum  or 
a  minimum  value  of  the  function  we  are  dealing  with  ;  if  not,  the 
sign  of  the  second  differential  coefficients  must  be  examined.  The 
second  derivative  is  positive,  if  the  function  is  a  minimum  ;  and 
negative,  if  the  function  is  a  maximum.  But  you  will  have  to  look 
up  some  text-book  for  particulars,  say  B.  Williamson's  Integral 
Calculus,  London,  463,  1896. 

Examples. — (1)  What  is  the  shortest  line  between  two  points  ?  A  straight 
line  of  course.  But  let  us  see  what  the  calculus  of  variations  has  to  say  about 
this.  The  length  of  a  curve  between  two  points  whose  absciss®  are  xx  and  x0, 
is,  page  246, 

fcf^W* •  <18> 

This  must  be  a  minimum.  Here  7"  is  a  function  of  y.  Hence  all  the  terms 
except  dQldx  vanish  from  (17),  and  we  get 

g  =  0;or,Q  =  C,  .         ,         .         .         .       (19) 

where  C  is  constant.     But,  by  definition  (9), 

9_f£,_*_  =0,        ...        ,       (20) 


572  HIGHER  MATHEMATICS.  §  174. 

since  V  -  \/l  +  #2 ;  .•.  dV  =  (1  +  yap*  #  .  d#.     Accordingly, 

2/  =  (l+^)Cf2;  .-.^(l-O-l;  .-.£  =  a,  .  .  (21) 
where  a  must  be  constant,  since  G  is  constant.  Hence,  by  integrating  y  =  a, 
we  get 

y  =  ax  +  b, (22) 

where  b  is  the  constant  of  integration.  The  required  curve  is  therefore  a 
straight  line  (8),  page  90.     Again,  from  (16)  and  (20), 

^-^iTir^-TO^  •    •    •    •  <28> 

If  the  two  given  points  are  fixed,  5^  =  0,  and  Sy0  =  0,  hence  Ix  - 10  vanishes. 
Let  x0,  y0,  and  xlt  ylt  be  the  two  fixed  points.     Then, 

y0  =  ax0  +  b;  y1  =  ax1  +  b (24) 

If  only  x0,  and  a^  are  given,  so  that  y0  and  y1  are  undetermined,  we  have,  by 
the  differentiation  of  (24),  yx  =  a.     Hence,  by  substitution  in  (23), 
dy  a 

£  =  a'>  •'•  -JFfr  Wi  -  *v°)  =  °«      =      •      •     <25) 

Since  8y  and  Zy  are  arbitrary,  (25)  can  only  be  satisfied  when  a  =  0.  The 
straight  line  is  then  y  =  b.  This  expresses  the  obvious  fact  that  when  two 
straight  lines  are  parallel,  the  shortest  distance  between  them  is  obtained  by 
drawing  a  straight  line  perpendicular  to  both. 

(2)  To  find  the  "  curve  of  quickest  descent "  from  one  given  point  to 
another.  Or,  as  Todhunter  puts  it,  "  suppose  an  indefinitely  thin  smooth  tube 
connects  the  two  points,  and  a  heavy  particle  to  slide  down  this  tube ;  we 
require  to  know  the  form  of  the  tube  in  order  that  the  time  of  descent  may  be 
a  minimum  ".  This  problem,  called  the  brachistochrone  (brachistos  =  shortest ; 
chronos  =  time),  was  first  proposed  by  John  Bernoulli  in  June,  1696,  and  the 
discussion  which  it  invoked  has  given  rise  to  the  calculus  of  variations.  Any 
book  on  mechanics  will  tell  you  that  the  velocity  of  a  body  which  starts  from 
rest  is,  page  376,  Ex.  (4), 

£  =  ^2gy,  5        .        c  (26) 

where  the  axis  y  is  measured  vertically  downwards,  and  the  <c-axis  starts  from 
the  upper  given  part.     The  time  of  descent  is  therefore 

Hm-^^-kfr^  ■  •  <27> 

as  you  will  see  by  glancing  at  page  569,  (6).     Accordingly,  we  take 

T-$p?   .    .    -    .    .   m 

so  that  V  only  involves  y  and  y.     Hence,  for  a  minimum,  we  have 
_    dQ      .  dV       d  fdV\      _• 

When  "Pdoes  not  contain  x  explicitly,  the  complete  differential  of  the  function 

F-/(y,$,f,...), (30) 

is  evidently 

dV_dVdy  ,dVdy            _p%,    0dl  +  Bdl^  ,ft1* 

lx~~  dy'dx*  c#*<^      dx  +  Vdx+     dx  +  "->         (61> 


§  175.  THE  CALCULUS  OF  VARIATIONS.  573 

as  indicated  on  page  72.       Multiply  (17)  through  with  dyjdx,  and  subtract 
the  result  from  (81).     The  P  terms  vanish,  and 

1M^!^)-(H-!)-^---- 

remains.     This  may  be  written  more  concisely, 

dV=d  /    dy\  _JL/dR  dy  _        \ 
dx      dlc\^dx)       dx\dx'dx         *) 

which  becomes,  on  integration, 

v_Qdy_dRtd^dPy  (32) 

v  ~  Vdx      dx    dx      ndx*  +  '  *  *  +  u§  v    ' 

where  C  is  the  constant  of  integration.     Particular  cases  occur  when  P,  Q,  or 

R  vanish.     The  most  useful  case  occurs,  as  here,  when  V  involves  only  y  and 

y.    In  that  case,  (29)  reduces  to 

V-&  +  0.      .        .        .        ■        .        03) 


tfrom  (28)  we  get 

a/TTF 

^j/o+W1"*"  "Vy(i  + jf8) 


F-  J1+F-    1    V*        +  <?•    •  1  C;      .        (34) 


Consequently, 

2/(1  +  y2)  =  constant,  say  =  2a.         .        .        •        (35) 

.  f^Y  _  2a  ~  y         ?x  _  /      y      \t=         y  *        (36) 

' '  \<W  ~       y      '  "dy-\2a-yj      J2ay  -  y*' 

On  integration,  using  (17),  page  193, 

x  =  o  vers  -  |-  si  2y  -  y*  +  6,        (37) 

where  6  is  an  integration  constant.     This  is  the 

well-known  equation  called  the  cycloid  (Fig.  174). 

The  base  of  the  cycloid  is  the  as-axis,  and  the 

curve  meets  the  base  at  a  distance  6,  or,  Ox, 

Fig.  174,  from  the  origin.     When  6  =  0,  the  origin  is  at  the  upper  point  so 

that  x  =  0,  when  6  =  0.     Now 

But  the  extreme  points  are  fixed  so  that  5y0  and  5^  vanish,  hence,  Ix  -  I0  also 
vanishes.  If  only  the  abscissa  of  the  lower  point  is  given,  not  the  ordinates, 
J0  vanishes,  as  before,  and  therefore, 

*-»  •....        (39) 

But  8yx  is  arbitrary,- hence,  if  Ix  is  to  vanish,  yx  must  be  zero.  This  means 
that  the  tangent  to  the  cycloid  at  the  lower  limiting  point  must  be  horizontal 
with  the  cc-axis. 


§  175.  The  Variation  cf  an  Integral  with  Variable  Limits. 

The  preceding  problem  becomes  a  little  more  complex  if  we  as- 
sume that  we  have  two  given  curves,  and  it  is  required  to  find  "  the 


574  HIGHER  MATHEMATICS.  §  175. 

curve  of  quickest  descent"  from  the  one  given  curve  to  the  other. 
Here  we  have  not  only  to  find  the  path  of  descent,  but  also  the 
point  at  which  the  particle  is  to  leave  one  curve  and  arrive  at  the 
other.  The  former  part  of  the  question  is  evidently  work  for  the 
calculus  of  variations,  and  the  latter  is  readily  solved  by  the  differ- 
ential calculus :  given  the  curve,  to  find  its  position  to  make  t  a 
minimum.     The  value  of  the  integral 


U 


=  \XlVdx,     .        .        .        .        (40) 


not  only  changes  when  y  is  changed  to  y  +  Sy,  but  also  when  the 
limits  tfj  and  x0  become  x1  +  dxv  and  x0  +  dx0  respectively.  The 
change  of  the  limits  augments  U  by  the  amount 

1*1 +  ^l  f  *o  +  dxo 

V.dx  -  V0.dx;  .         (41) 

~  *i  r  J  *o 

or,  neglecting  the  higher  powers  of  dxx  and  dx0,  U  receives  the 

increment 

dU  =  V1dx1  ~  V0dx0.       .        .        .       (42) 

The  total  increment  of  U  is  therefore 

Total  incr.  U  =  dU  +  SU  =  Vldx1  -  V0dxQ  +8    V.  dx.       (43) 

In  words,  the  total  increment  which  a  quantity  receives  from  the 
operation  of  several  effects  is  the  sum  of  the  increments  which  each 
effect  would  produce  if  it  acted  separately.  This  is  nothing  but  the 
principle  of  the  superposition  of  small  motions,  pages  70  and  400, 
under  another  guise. 

The  maximum-minimum  condition  is  that  the  total  increment 
be  zero.  This  can  only  obtain  when  V1  =  0,  and  V0  =  0.  We 
thus  have  two  new  conditions  to  take  into  consideration  besides 
those  indicated  in  the  preceding  section. 

Example. — Find  the  "  curve  of  quickest  descent "  from  one  given  curve  to 
another.  Ex.  (2),  page  572,  has  taught  us  that  the  "  curve  of  quickest  descent" 
is  a  cycloid.  The  problem  now  before  us  is  to  find  the  relation  between  the 
cycloid  and  the  two  given  curves.  We  see  from  (15)  and  (43)  that  the  maxi- 
mum-minimum condition  is 


Vldx1  -  Tq^o  +  I1-  IQ+  jKSydx  =  0.  .        .        .      (44) 

J  XQ 

re  can  use  the  results  of  Ex.  (2),  page  573,  equations  (33)  and  (38),  there- 
tie  maximum-minimum  condition  becomes 

i~~i  \tyi(l  +  2/i2)      VZ/o^  +  ^/o2)      Jx°\         axJ 


§  176.  THE  CALCULUS  OF  VAKIATIONS.  575 

As  before,  (29)  holds  good,  consequently, 

n/2/i(1  +  2/2)  =  >/2a.       ....        (46) 
...  Vxdxx  -  VQdx0  +  ^gj*  -  |>  )  =  0.  .        .        (47) 

Eemembering  that  the  end  values  of  the  curve  are  x0,  y0,  and  a^,  2/l5  let  y 
suffer  a  variation  Sy  so  that 

r  =  y+5y, (48) 

with  fixed  limits,  and,  at  the  same  time,  x0,  y0,  and  xx,  yx,  respectively  become 
x0,  F0,  and  xx,  Yx.     Let  us  find  how  5yQ  and  8yx  are  affected  when  the  values 
of  x  change  respectively  to  x0  +  dxQ,  and  xx  +  dxx.    By  Taylor's  theorem, 
instead  of  yx  becoming  Yx,  we  have  Yx  changed  to 
dY,  1    dT, 

Y^  dxxdx^W^dx^  + <49> 

or,  from  (4)  and  (48), 

to  +  8yJ  +  (i^  +  WiSpl ' <fal) +  •  • '      •'"    '    (50) 

Neglecting  the  higher  powers  of  dxx  and  the  product  Syv  dxx, 

Yx  becomes  yx  +  8yx  +  yxdxx,     .        .        .  (51) 

as  a  result  of  the  variation  and  of  the  change  of  xx  into  xx  +  dxx. 
Let  the  equation  of  one  of  the  curves  be 

2/=/(*i) (52) 

then  the  abscissa  of  the  end  value  of  Yx  is  changed  into  f(xx  +  dxx)  after  the 
variation.    Consequently,  after  variation, 

Vi  +  tyi  +'yidxx  =f(xx  +  dxj  =f{xx)  +f'{xx)dxxt 
by  Taylor's  theorem.     From  (50)  we  can  cancel  out  the  y'a  and 

ty  =  {/(«i)  -  yi)dx> (53) 

remains.     A  similar  relation  holds  good  between  Sy0  and  dx0. 

Let  us  return  after  this  digression  to  (47),  and,  in  order  to  fix  our  ideas, 
let  the  two  given  curves  be 

yx  =  ma^  +  a  ;  y0  =  mx0  +  b;  .-.  yx  =  m  ;  y0  =  n.         .      (54) 
From  (53)  we  have 

Syx  =  {m  -  ^x)dxx ;  Sy0  =  (n  -  y0)dx0.     .        .        .      (55) 
Substitute  these  values  in  (47),  and 

{7l+  Js(m  "  *l)}dXl  ~{Vo  +  Wa{n  ~  **}**  =  °'  '  (56) 
Since  dxj  and  dx0  are  arbitrary,  the  coefficients  of  dxx  and  dxQ  must  be 
separately  zero  in  order  that  (56)  may  vanish. 

.M+.fcm-0;l  +  fe.-0;«.g— i;g~±.      .      (57) 

Now  compare  this  result  with  (18),  page  96,  and  you  will  see  that  the  two 
given  curves  are  at  right  angles  with  the  "  curve  of  quickest  descent". 

§  176.  Relative  Maxima  and  Minima. 

After  the  problem  of  the  brachistochrone  had  been  solved, 
James  Bernoulli,  brother  of  John,  proposed  another  variety  of 
problem — the  so-called  isoperimetrical  problem — of  which  the  fol- 


676  HIGHER  MATHEMATICS.  §  176. 

lowing  is  a  type  :  Find  the  maximum  or  minimum  values  of  a  certain 
integral,  Uv  when  another  integral,  U2,  involving  the  same  vari- 
ables has  a  constant  value.  The  problem  proposed  at  the  beginning 
of  this  chapter  is  a  more  concrete  illustration.  Here,  8^  must 
not  only  vanish,  but  it  must  vanish  for  those  values  of  the  vari- 
ables which  make  U2  constant.  It  will  be  obvious  that  if  U-^  be 
a  maximum  or  a  minimum,  so  will  U1  +  aU2  also  be  a  maximum 
or  a  minimum  ;  a  is  an  arbitrary  constant.  The  problem  therefore 
reduces  to  the  determination  of  the  maximum  or  minimum  values 
of  Ux  +  aU2.     If 

U1=\*1V1dx;    U2  =   \X1V2dx;    .       '.        (58) 
J  *o  J  *o 

Ux  +  aU2  will  be  a  maximum  or  a  minimum  when 


J 


(Vi  +  V2a)dx  =  0,      .        .        .        (59) 
*o 

is  a  maximum  or  a  minimum.  When  U2  is  known,  a  can  be 
evaluated. 

Example. — Find  the  curve  of  given  length  joining  two  fixed  points  so 
that  the  area  bounded  by  the  curve,  the  aj-axis,  and  the  ordinates  at  the  fixed 
points  may  be  a  maximum.    Here  we  have 

TJX  =  j\dx  ;  U,  =/^V1  +  (i)^'  '        '        •        (6°) 
as  indicated  on  page  246.    Here  then 

Vx  +  aV2  =  y  +  ajl  +  f- (61) 

We  require  the  maximum  value  of  the  integral 

^/?„f-W^Wi_+IIH-  •  •  <62> 

7  is  a  function  of  y  and  y,  hence  from  (19)  we  must  have 

V=P$  +  C; (63) 

i ay2  a 

...y  +  ajrTJ^jTr~+Cli.:y+:jrTj^cl,.     (64) 

By  a  transposition  of  terms, 

1  +  \dx)  ~  (y  -  CJ*'  -\dx)  ~  a*-(y-  C,)2'       ■       '      t65) 

which  becomes,  on  integration, 

x-C2=  J  a2-  {y-  CJ2;  or,  {x  -  C2)2  +  {y  -  Ctf  =  a\      .      (66) 

This  is  obviously  the  equation  of  a  circular  line.  The  limits  are  fixed,  and 
therefore  Ix  -  I0  =  0.  The  constants  a,  Clt  and  C2  can  be  evaluated  when 
the  fixed  points  and  the  length  of  the  curve  are  known. 


§  178.  THE  CALCULUS  OF  VARIATIONS.  577 

§  177.  The  Differentiation  of  Definite  Integrals. 

I  must  now  make  a  digression.  I  want  to  show  how  to  find  the 
differential  coefficient,  du/da,  of  the  definite  integral  u  =  \f  (x,  a)dx 
between  the  limits  yx  and  y0,  when  yx  and  y0  are  functions  of  a. 
Letf(x,  a)dx  become  f(x,  a)  after  integration,  we  have  therefore 

u  =  I    f{x,  a)dx  =  f(yv  a)  -  f(y0,  a).        .        (67) 

Hence,  on  partial  differentiation   with  respect   to  yv  when  yQ  is 
constant ;  and  then  with  respect  to  yQ,  when  y1  is  constant,  we  get 
7)u       d  ~du       d 

ay, = jjfte»  a>  *■/<*»  a)'~wr  ¥o/(y°' a) =fiy'" ay  (68) 

Now  suppose  that  a  suffers  a  small  increment  so  that  when  a  be- 
comes a  +  h,  u  becomes  u  +  k,  then,  keeping  the  limits  constant, 
iucT.v  =  j{f'(x,a  +  h)-f(x,a)}dx.      .         .       (69) 
Dividing  by  8a,  and  passing  to  the  limit,  we  have 

mcr.  u  =  [y.Ax,a+h)-f{xta)     .    ■  dM[y,  df'(x,a) 

Incr.  a    )y0  h  '  '  '  da    )y       da     a      ^iK)} 

If  both  yx  and  y0  are  functions  of  a,  then  du/da  must  be  the 

sum  of  three  separate  terms,  (i)  the  change  due  to  a  ;  (ii)  the  change 

due  to  y1 ;  and  (iii)  the  change  due  to  y0.     These  separate  effects 

have  been  evaluated  in  equations  (68)  and  (70),  consequently, 

£4J>^ -£*£*•*£ £-&•&  ft 

The  higher  derivatives  can  be  obtained  by  an  application  of  the 
same  methods. 

§  178.    Double  and  Triple  Integrals. 

We  now  pass  to  double  integrals,  say, 

U=jjVdxdy,      ....       (73) 
where  V  is  a  function  of  x,  y,  z,  p,  and  q,  and 

dz  dz  /.,. 

*-»/*""*•  '  •  •  <74> 
We  apply  the  same  general  methods  as  those  employed  for  single 
integrals,  but  there  are  some  difficulties  in  connection  with  the 
limits  of  integration  of  multiple  integrals.  Let  8z  denote  the  varia- 
tion of  z  which  occurs  when  the  form  of  the  function  connecting  z 
with  x,  and  y  is  known,  x  and  y  remaining  constant  during  the 

00 


(78) 


578  HIGHER  MATHEMATICS.  §  178. 

variation.  Further,  let  SV  denote  the  variation  of  V,  and  8  U  the 
variation  of  U,  when  z  becomes  8z ;  then  by  the  preceding  methods, 

w=^+%^+Tqh=Pte+QSp+%h,     •     (75) 

where  we  have  put  for  the  sake  of  convenience, 

*-&«-?"-?■  ■  •  <- 

We  therefore  write,  from  (75), 

8 17= fyvdxdy  =  jj(p&  +Q^+  B^)dxdy.      .      (77) 
Still  keeping  on  the  old  track, 

-oi(*-s*f)«** 

The  differential  coefficients  with  respect  to  x  and  y  are  complete. 
We  get,  on  integration  with  respect  to  y, 

[*'["' ^(R&z)dxay=\'1\Rte~\'Xdx,       .        .      (79) 

JxoJyou"L  J*oL         J0O 

where     RBz     ,  as  on  page  232,  represents  the  value  of  RBz  wheu 

y1  and  y0  are  each  substituted  in  place  of  y,  and  the  latter  then 
subtracted  from  the  former.  Again,  from  (70)  followed  by  a  trans- 
position of  terms,  we  get 

r-r*- if.** -[<*>!]:■  ■  m 

where     (QSzJy       denotes  the  value  of  (QSzJy  when  yl  and  y0  are 

each  substituted  in  place  of  y,  in  Qdx,  and  the  latter  subtracted  from 
tha  former.     Hence,  we  may  write 

JT/^«-M:-J>>i]::-  <»» 

By  substituting  (79)  and  (81)  in  place  of  (78),  we  get 

If  the  limits  yx  and  y0  are  constant,  yx  and  y0  vanish,  and  we  can 
therefore  neglect  the  last  term.     If  the  limits  also  change,  we  must 


§  178.  THE  CALCULUS  OF  VARIATIONS.  679 

add  on  a  new  term  in  accordance  with  the  principles  laid  down  in 
§175. 

For  the  maximum -mini  mum  condition,  BU  of  (82)  can  only 
vanish  when  the  coefficient  of  Sz,  namely, 

*-2-S-*    ■   •   •  « 

The  solution  of  this  partial  differential  equation  furnishes  z  in  terms 
of  x,  y,  and  arbitrary  functions ;  the  latter  must  be  so  determined 
that  the  remaining  terms  of  (82)  vanish. 
For  the  triple  integral 

U  m  fSJVdxdydz,  .        .        .      (84) 

where  V  is  a  given  function  of  u,  x,  y,  z,  p,q,r;  and  u  is  a  function 
such  that 

du  du  du  ,„_% 

•  (86) 

•  (87) 

•  (88) 


We  have  also 

8(7  =  iSJBVdxdydz. 

•         • 

As  before, 

«~      ,»„         ^d8u       „d8u 

dSu 

8F=WSM  +  P_+g_ 

***>• 

where 

dV            dV            dV 

T>           dV 

*  ~  du'>  F  ~  dp'  *  ~  dq'> 

B=fo> 

and  the  variation  works  out  to 

^=IJI(^-S-f-S>«^+ 


(89) 


For  the  maximum-minimum  condition,  we  must  solve  the 
partial  differential  equation 

>T      dP      dQ      dB     _ 

N-dx--H  +  !z--°>  '  •  •  (9°) 
and  fit  the  arbitrary  constants  so  that  the  remaining  terms  of  8U 
vanish.  A  complete  exposition  of  the  subject  would  be  quite  outside 
the  limits  of  this  volume.  J.  H.  Jellet's  An  Elementary  Treatise  on 
the  Calculus  of  Variations,  Dublin,  1850,  is  a  good  text-book; 
O.  Bolza,  in  his  Lectures  on  the  Calculus  of  Variations,  Chicago, 
1904,  has  a  review  of  modern  theory. 

J.  H.  van  der  Waals  seeks  the  maximum  value  of  a  triple  integral  in  his 
Bintire  Qemische,  Leipzig,  34, 1900,  but  the  physical  conditions  of  the  problem 
enable  the  solution  of  (90)  to  be  obtained  in  a  simple  manner. 

OO* 


CHAPTER  XI. 

DETERMINANTS. 

"  Operations  involving  intense  mental  effort  may  frequently  be  re- 
placed by  tbe  aid  of  other  operations  of  a  routine  character, 
with  a  great  saving  of  both  time  and  energy.  By  means  of  the 
theory  of  determinants,  for  example,  certain  algebraic  opera- 
tions can  be  solved  by  writing  down  the  coefficients  according  to 
a  prescribed  scheme  and  operating  with  them  mechanically." — 
E.  Mach. 

§  179.    Simultaneous  Equations. 

This  chapter  is  for  the  purpose  of  explaining  and  illustrating  a 
system  of  notation  which  is  in  common  use  in  the  different  branches 
of  pure  and  applied  mathematics. 

I.  Homogeneous  simultaneous  equations  in  two  unknowns. 

The  homogeneous  equations, 

aYx  +  bxy  =  0 ;  a2x  +  b2y  m  0,  .        .      (1) 

represent  two  straight  lines  passing  through  the  origin.  In  this 
case  (§  29),  x  =  0  and  y  =  0,  a  deduction  verified  by  solving  for 
x  and  y.  Multiply  the  first  of  equations  (1)  by  b2,  and  the  second 
by  bv  Subtract.  Or,  multiply  the  second  of  equations  (1)  by  av 
and  the  first  by  a2.     Subtract.     In  each  case,  we  obtain, 

x(a1b2  -  a2b{)  =  0  ;  y{a2bx  -  axb2)  =  0.    .         .      (2) 
Hence,  x  =  0  ;  and  y  =  0  ;  or, 

a-J)2  -  a2\  =  0  ;  and  a2bY  -  ajb2  =  0.  .      (3) 

The  relations  in  equations  (3)  may  be  written, 

av    bx  I  =  0  ;  and  \a2,    b2 1  =  0,        .         .       (4) 
a2,    b2\  \av    &J 

where  the  left-hand  side  of  each  expression  is  called  a  determinant. 
This  is  nothing  more  than  another  way  of  writing  down  the  differ- 
ence of  th%  diagonal  products.     The  letters  should  always  be  taken 

580 


§  179.  DETERMINANTS.  581 

in  cyclic  order  so  that  b  follows  a,  c  follows  b,  a  follows  c.  In  the 
same  way  2  follows  1,  3  follows  2,  and  1  follows  3. 

The  products  axb^  a^,  are  called  the  elements  of  the  determinant ; 
Oj,  fej,  a^  o2,  are  the  constituents  of  the  determinants.  Commas  may  or  may 
not  be  inserted  between  the  constituents  of  the  horizontal  rows.  When  only 
two  elements  are  involved,  the  determinant  is  said  to  be  of  the  second  order. 

From  the  above  equations,  it  follows  that  only  when  the  de- 
terminant of  the  coefficients  of  two  homogeneous  equations  in  x  and 
y  is  equal  to  zero  can  x  and  y  possess  values  differing  from  zero. 
II.  Linear  and  homogeneous  equations  in  three  unknowns. 
Solving  the  linear  equations 

axx  +  bxy  +  cx  =  0 ;  a^x  +  b2y  +  c2  =  0,  .        .     (5) 
for  x  and  y,  we  get 

=  bxc2  -  b2cx  (      =  0la9  -  c2ax 

aYb2  -  bxa2 '  y       axb2  -  bxa2  '       ^  ' 

If  axb2  -  bxa2  =  0,  x  and  y  become  infinite.  In  this  case,  the  two 
lines  represented  by  equations  (5)  are  either  parallel  or  coincident. 
When 

x  =  _ =  qo  .    y  m   _ =  Q^ 

the  lines  intersect  at  an  infinite  distance  away.      Reduce  equations 

(5)  to  the  tangent  form,  page  90, 

but  since  a-fi2  -  bxa2  =  0,  a1/b1  =  a2/b2  =  the  tangent  of  the  angle 
of  inclination  of  the  lines  ;  in  other  words,  two  lines  having  the 
same  slope  towards  the  #-axis  are  parallel  to  each  other.1 

When  the  two  lines  cross  each  other,  the  values  of  x  and  y  in 

(6)  satisfy  equations  (5).     Make  the  substitution  required. 

ax{ bxc2  -  b2cY)  +  b^c^  -  c2ax)  +  c^a^  -  a2bx)  =  0, 

a2(  V2  ~  Vi)  +  k*(cia2  ~  <¥*i)  +  c20  A  -  a2&i)  -  0, 

or,  writing 

X  Y 

x  =-g  ;  and  y  =  ^, .         .         .         .       (8) 

we  get  a  pair  of  homogeneous  equations  in  X,  Y,  Z,  namely, 

aYX  +  bxY  +  cxZ  =  0 ;  a2X  +  b2Y  +  c2Z  =  0.  (9) 

1Thus  the  definition,  "  parallel  lines  meet  at  infinity,"  means  that  as  the  point 
of  intersection  of  two  lines  goes  further  and  further  away,  the  lines  become  more  and 
more  nearly  parallel. 


682  HIGHER  MATHEMATICS.  §  179. 

Equate  coefficients  of  like  powers  of  the  variables  in  these  identical 
equations. 

.*.  ax :  bx :  ox  =  a2 :  b2  :  c2, 
or,  from  (8)  and  (6), 

X  :Y:  Z  =  bxo2  —  b2o1 :  cxa2  —  c2aY :  a1b2  -  a2blt 

=  \bi  cihlci   <h|:K   M-         •        •      (10) 


(11) 


\b2   c2\   \c2   a2\   \a2    b(jL 
The  three  determinants  on  the  right,  are  symbolized  by 

Iai    bi    ci  II 
a2   b2   c2  1 

where  the  number  of  columns  is  greater  than  the  number  of  rows.1 

The  determinant  (11),  is  called  a  matrix.      It  is  evaluated,  by 

taking  the  difference  of  the  diagonal  products  of  any  two  columns. 

The  results  obtained  in  (10)  are  employed  in  solving  linear 

equations. 

Examples.— (1)  Solve  4<e  +  5y  =  7 ;  Bx  -  lOy  =  19. 

X:Y:Z  =  i      5,  -  71:1-  7,   4  l :  I  4,      51; 
|-10,   -19  I    1-19,    3|    |  3,-10 1 
=  -165:55:  -55;  or  x  =  +  3  and  y=  -1. 

(2)  Solve  20x  -  19y  =  23  ;  19a;  -  20y  =  16.    Ansr.  x  =  4,  y  =  3. 

(3)  Sqtye  the  observation  equations : 

-5x-'2y  =  -4 ;  -14a;  +  'By  =  1-18.     Ansr.  x  m  2,  y  =  3. 

(4)  Solve  \x- \y  =  6  ;  $x - %y  =  - 1.     Ansr.  x  =  24,  y  =  18. 

The  condition   that   three  straight  lines   represented  by  the 
equations 

aYx  +  b±y  -r  ^  =  0 ;  «b2x  +  b2y  +  c2  =  0 ;  a^x  +  b3y  +  c3  =  0,  (12) 
may  meet  in  a  point,  is  that  the  roots  of  any  two  of  the  three 
lines  may  satisfy  the  third  (§  32).  In  this  case  we  get  a  set  of 
simultaneous  equations  in  X,  Y,  Z. 

a1X-^b1Y+c1Z=  a2X+b2Y+c2Z  =  asX+bsY+c3Z  =  0,      (13; 
by  writing  x  =  XI Z  and  y  =  Y\Z  in  equations  (12).     From  the 

last  pair, 

Y:Z  =  \b2,    c2\:\g2,    a2\:\a2,    b2\.         .       (14} 


^8>     C3 1     I  C3>     a3  I     I  a&      °S 


But  these  values  of  x  and  y,  also  satisfy  the  first  of  equations  (3), 
hence,  by  substitution, 

h\K    c2|  +  Mc2,    aJ  +  cJag,    62|  =  0,  .      (15) 

K    cal         «c3»    as\        la8>    hi 


1  It  is  customary  to  call  the  vertical  columns,  simply  "  columns  "  ;  the  horizontal 
rows,  "rows". 


§180. 


DETERMINANTS. 


which  is  more  conveniently  written 


ax 

bi   *i 

<*>2 

b2    c2 

<h 

h    c3 

=0, 


583 


(16) 


a  determinant  of  the  third  order. 

It  follows  directly  from  equations  (13),  (14),  (16),  only  when 
the  determinant  of  the  coefficients  of  three  homogeneous  equations 
in  x,  y,  z,  is  equal  to  zero,  can  x,  y,  z,  possess  values  differing  from 
zero.  This  determinant  is  called  the  eliminant  of  the  equations. 
Each  determinant  in  (14)  is  called  a  subdeterminant,  or  minor 
of  (16). 

§  180.    The  Expansion  of  Determinants. 

It  follows  from  (15)  and  (16),  that 


ai  bi 


3    bS 


-  a^jCg  +;o.263c1+  o,^  -  axbzc2  -  a2&i0s  ~  aAcr   C1?) 


A  determinant  is  expanded,  by  taking  the  product  of  one  letter 
in  eaoh  horizontal  row  with  one  letter  from  each  of  the  other  rows. 
The  first  element,  called  the  leading  element,  is  the  product  of 
the  diagonal  constituents  from  the  top  left-hand  corner,  i.e.,  alb2c3 ; 
its  sign  is  taken  as  positive.  The  signs  of  the  other  five  terms  1 
are  obtained  by  arranging  alphabetically,  and  observing  whether 
they  can  be  obtained  from  the  leading  element  by  an  odd  or  an 
even  number  of  changes  in  the  subscripts ;  if  the  former,  the 
element  is  negative,  if  the  latter,  positive.  For  example,  a2b1c$, 
is  obtained  by  one  interchange  of  the  subscripts  2  and  1  in  the 
leading  element ;  ajbxcz  is,  therefore,  a  negative  element ;  ajb^ 
requires  two  suoh  transformations,  2  and  1,  and  2  and  3,  hence 
its  sign  is  positive. 

Examples.— -(1)  Show  12  2  21  =  2  + 12 +  8-4-6-8  =  4. 


(2)  Show 


0  b  e 
b  0  a 
o  a  0 


3 

I  i 

=  2abc 


1  The  number  of  elements  in  a  determinant  of  the  second  order  is  2  x  1,  or  2 ! 
of  the  third  order  8  x  2  x  1,  or  3  !,  of  the  fourth  order,  4  I,  etc. 


584 


HIGHER  MATHEMATICS. 


§181. 


§  181.   The  Solution  of  Simultaneous  Equations. 

Continuing  the  discussion  in  §  179,  let  the  equations 
aix  +  \V  +  <>iZ  =  di ;  a2x  +  b2y  +  c2z  =  d2 ;  a3x  +  b$  +  czz  =  dz,    (18) 
be  multiplied  by  suitable  quantities,  so  that  y  and  z  may  be  elimi- 
nated.    Thus  multiply  the  first  equation  by  Av  the  second  by  A2, 
the  third  by  Az,  where  Av  A2,  Az,  are  so  chosen  that 

Mi  +  M2  +  Mi  =  °  [  Mi  +  <V*2   +Ms  =  0.         (19) 
Hence,  by  substitution, 

x{aYAx  +  a2^2  +  a3A3)  =  d^  +  d2A2  +  d343.  (20) 

Equations  (19)  being  homogeneous  in  Av  A2,  Az,  we  get,  from  (10), 
A1:A2:As  =  \b2   b9l:'\b3   M  :  I  6X   b2 
\c2    c3 1     I  C3    Cj  I     I  Cj    c2 
Substituting  these  values  of  Av  A2,  A3,  in  equations  (20),  we  get, 
as  in  equations  (14),  (15),  (16), 

(21) 
a2   bs 


cl 

= 

dx   bx  Cj 

<v 

d2   b2  c2 

C2 

d3   b3  c3 

In  the  same  way,  on  multiplying  by  Bv  B2,  Bz,  and  by  Cv  C2,  Cs 


y 


h 


4* 


=    a, 


do  d, 


3     ^3 


('l 

;  z 

ax   bx   cY 

= 

<H 

a2   o2   c2 

c3 

az  o3  c3 

bl  dl 

a2   b2  d2 
a3   bs  d3 


.(22) 


Examples. — Solve  the  following  set  of  equations : 

(1)  5x  +  3y  +  3z  =  48;  2x  +  6y  -  3z  =  18;  8a;  -  3y  +  2z-. 


21.     From  (21) 


=  3 


x  =  I  48       3 

18       6    - 
I  21    -3 

similarly  y  =  5 ;  z  =  6. 

(2)  H.  E.  Roscoe  and  C.  Schorlemmer  (Treatise  on  Chemistry,  1,  704, 
London,  1878)  use  the  following  set  of  equations  in  the  analysis  of  a  mixture 
of  gases  containing  C2H4,  C3H6,  and  06H6  gases : 

x  +  y  +  z  =  a ;  2x  +  3y  +  6z  =  b ;  2x  +  %y  +  %z  =  c, 
where  a,  b,  c,  are  numbers  obtained  from  the  gas  burette.     Solve  for  x,  y,  z. 

(3)  Field's  process  (Jour.  Chem.  Soc,  10,  234,  1858)  for  the  determination 
of  chlorine,  bromine,  and  iodine  when  mixed  in  solution  involves  the  equations 

x  +  y  +  z  =  a ;  1-31&  +  y  +  z  =  b ;  l-637«  +  l-25y  +  z  =  c, 
where  a,  b,  and  c  are  numbers  determined  by  analysis.     Similar  equations 
arise  in  the  indirect  process  of  analysis  of  a  mixture  of  sodium  and  potassium 
salts.    Find  x,  y,  z. 

(4)  Solve  the  observation  equations : 

•3a?  +  '2y  +  -5z  =  3-2 ;  -2x  +  -3y  +  '4z  =  2*9 ;  -4a;  +  -3y  +  -&z  =  3*7. 
Ansr.  x  =  2,  y  =  3,  z  =  4. 


§  182.  DETERMINANTS.  585 

(5)  To  illustrate  the  solution  of  simultaneous  equations  "  by  the  writing 
down  of  the  coefficients  according  to  a  fixed  scheme  and  operating  upon  them 
according  to  a  prescribed  scheme,"  take  the  proof  of  (2),  from  (1),  page  444, 
as  an  exercise.  In  equation  (1)  take  z  as  an  independent  variable  and  solve 
the  two  simultaneous  equations  for  dx/dz  and  dyfdz.     Hence, 

\QX    Bx  B1    Px 

4?  - 1  #2  -^2     <ty     -^2  P2  dx  dy  dz 

«fe  ~    P,    Q1    ''  dz=  Px    Ql   ;  •'•  I  Qx    Rx  |=prPTr  I  P~Qx\ 
|P2     Q2  P2     Q2  \Q2    B2\    \B2    P2|    |P2     Q2\ 

which  has  the  same  form  as  (2),  page  445. 

§  182.  Test  for  Consistent  Equations. 

It  is  easy  to  find  values  of  x  and  y  in  the  two  equations 
axx  +  bYy  +  cx  =  0 ;  a2x  +  b2y  +  c2  =  0, 
as  shown  in  §  179,  (6) ;  similarly,  values  for  x,  yf  and  z  in  the 
equations 

axx  +  bxy  +  cYz  +  dx  =  0 ;  a^x  +  b2y  +  c2z  +  d2  ==  0 ; 
a&  +  bzy  +  csz  +  dz  =  0, 
can  be  readily  obtained,  and  generally,  in  order  to  find  the  value 
of  1,  2,  3,  .  .  .  unknowns,  it  is   necessary  and   sufficient  to  have 
1,  2,  3,  .  .  ,  independent  relations  (equations)  respectively  between 
the  unknowns. 

If  there  are,  say,  three  equations  and  only  two  unknowns,  it  is 
possible  that  the  values  of  the  unknowns  found  from  any  two  of 
these  equations  will  not  satisfy  the  third.    For  example,  take  the  set 

3a?  -  2y  =  4 ;  2x  +  ±y  =  24 ;  x  +  y  =  2. 
On  solving  the  first  two  equations,  we  get  x  =  4,  y  =  4.  But  these 
values  of  x  and  y  do  not  satisfy  the  third  equation.  On  solving  the 
last  two,  x  =  -  8,  y  =  10,  and  these  values  of  x  and  y  do  not  satisfy 
the  remaining  equation.  In  other  words,  the  three  equations  are 
inconsistent.  Consequently,  it  is  a  useful  thing  to  be  able  to  find 
if  a  number  of  equations  are  consistent  with  each  other ;  in  other 
words,  to  find  if  values  of  x  and  y  can  be  determined  to  satisfy  all 
of  a  given  set  of  equations.  For  instance,  is  the  set 
axx  +  bxy  +  cx  =  0 ;  a2x  •+  b2y  +  c2  =  0 ;  a^c  +  b3y  +  c3  =  0,  (23) 
consistent?     From  the  first  two  equations,  page  581,  we  get 

ajbi  -  bYa2  >  y      aYb2  -  \a2      '        '       v"*> 
Substitute  these  values  of  x  and  y  in  the  last  of  equations  (23),  the 


586  HIGHER  MATHEMATICS.  §  182. 

two  unknowns  disappear,  and,  if  the  equations  are  consistent, 

as{bxc2  -  b2cx)  +  bz(cxa2  -  c2ax)  +  0,(0^  -  a2bx)  =  0, 
remains.     But  this  result  is  obviously  the  expansion  of  the  de- 
terminant 

=  0,         .        .        .       (25) 


al 

h 

°l\ 

a2 

h 

C2 

az 

K 

cz\ 

and  this  in  consequence  called  the  eliminant  of  the  three  given 
equations.  Hence  we  conclude  that  three  equations  are  consistent 
with  each  other  only  when  the  determinant  of  the  coefficients  and 
absolute  term  of  the  three  linear  equations  in  x,  y,  z,  is  equal  to 
zero. 

Examples. — (1)  Show  that  the  following  equations  are  consistent  with 
one  another, 

x  +  y-z-Q\  x-y  +  z  =  2;  y  +  z-x  —  ^\  x  +  y  +  z  =  6. 
Hint.  The  eliminant  is  1      1-1      0=0. 

1-112 
1114 
1116 

(2)  A  point  oscillates  freely  in  space  under  the  action  of  a  force  directed 
from  the  origin  of  the  coordinates.    The  equations  of  motion  are 

cPx  d*y  d2z 

Find  the  path  of  the  point.     First  solve  the  equations  as  in  Ex.  (4),  page  442. 

x=Gx  cos  qt+C4  sin  qt;  y=G2  cos  qt+G5  sin  qt;  z=sCs  cos  qt+G6  sin  qt. 
Now  eliminate  t,  because  time  does  not  determine  the  form  or  position  of 
the  path.  Now  cos  qt,  and  sin  qt,  may  be  regarded  as  independent  variables 
to  be  treated  as  "  unknowns  ".  Of  course  two  equations  would  be  sufficient 
for  the  elimination,  but  three  are  given  and  all  must  be  satisfied.  For  con- 
sistency we  must  have 

Ix    Ox    G4 1  =  0. 
y  o2  gA 
z     Gs    G6\ 
When  the  determinant  is  expanded,  the  result  is  a  linear  homogeneous  equa- 
tion in  x,  y,  and  z  which  is  the  equation  of  a  plane  passing  through  the 
origin,  and  whose  position  is  determined  by  the  constants  O.     Suppose  the 
plane  to  be  rotated  so  that  it  coincides  with  the  xy-nl&ne,  then  z  =  0.     Solve 
for  cos  qt,  and  sin  qt,  and  substitute  the  results  in  the  well-known  equation  (19), 
page  611,  sin2  gtf  +  cos2  qt  —  1.     The  equation  is  of  the  second  degree.    Expand 
and  put  C22  +  C52  -  a  ;  2CA  -  C4C5  =  b;  d2  +  042  =  c ;  {0YOb  -  C204)2  =  h. 

.'.  ax-bxy  +  cy  =  h. 
In  the  discriminant  b2  -  iac,  a  and  c  are  necessarily  positive,  consequently 
the  curve  is  either  an  ellipse  or  a  circle. 


§183. 


DETERMINANTS. 


587 


§  188.  Fundamental  Properties  of  Determinants. 

The  student  will  get  an  idea  of  the  peculiarities  of  determinants 
by  reading  over  the  following  : — 

I.  The  value  of  a  determinant  is  not  altered  by  changing  the 
columns  into  rows,  or  the  rows  into  columns. 

It  follows  direotly,  by  simple  expansion,  that 


and 


<h 

W 

c, 

(h 

K 

4 

a3 

\ 

c, 

(26) 


It  follows,  as  a  corollary,  that  whatever  law  is  true  for  the  rows  of  a 
determinant  is  also  true  for  the  columns,  and  conversely. 

II  The  sign,  not  the  numerical  value,  of  a  determinant  is  altered 
by  interchanging  any  two  rows,  or  any  two  columns. 
By  direct  calculation, 


M= 

-|*i  ail;  K  &i 

M 

\b2  oj    Uj  62 

l«3     &3 

bx    Oj    Cjl 


(27) 


b2  «2  Ca 
68  <h  C3I 


JZT.  2/"  faao  rows  or  faoo  columns  of  a  determinant  are  identical, 
the  determinant  is  equal  to  zero. 

If  two  identical  rows  or  columns  are  interchanged,  the  sign,  not  the  value, 
of  the  determinant,  is  altered.  This  is  only  possible  if  the  determinant  is 
equal  to  zero.     The  same  thing  can  be  proved  by  the  expansion  of,  say, 

°i  ^  <a|=0. 

<h  <*3  <h\ 

IV.  When  the  constituents  of  two  rows  or  two  columns  differ  by 
a  constant  factor,  the  determinant  is  equal  to  zero. 

Thus  by  expansion  show  that 

14   1    o|  =  4|l   1  51  =  4x0  =  0.       .         ,         .       (28) 

8  2  6  2  2  61 

12  8  7 1         1 3  8  7 1 

V.Ifa  determinant  has  a  row  or  column  of  cyphers  it  is  eqtial 
to  zero. 


This  is  illustrated  by  expansion, 

0  \  % 
0  6.  c. 


0. 


(29) 


VI.  In  order  to  multiply  a  determinant  by  any  factor,  multiply 
zach  constituent  in  one  row  or  in  one  column  by  this  factor. 


588 


HIGHER  MATHEMATICS. 


$183. 


This  is  illustrated  by  the  expansion  of  the  following : 


a2 


h  cs\ 


ma-L 

h 

<h 

ma^ 

h 

c2 

ma3 

b. 

% 

(30) 


VII.  In  order  to  divide  a  determinant  by  any  factor,  divide  each 
constituent  in  one  row  or  in  one  column  by  that  factor. 

This  follows  directly  from  the  preceding  proposition.     It  is  conveniently 
used  in  the  reduction  of  determinants  to  simpler  forms.     Thus, 

=  9.6.2.2    112.         .       (31) 
111 
4  3   1 

VIII.  If  the  sign  of  every  constituent  in  a  row  or  column  is  changed, 
the  sign  of  the  determinant  is  changed. 


16     9  8 

=  9.6.2 

114 

12   18  4 

2   2  2 

1  24   27   2 

4   3   1 

<h   &i  ci  I  =  ~  I  ~  <h   bi  <a 


<h  b2  c2 


-  <%  b2  Cjj 


-<h 

-  by  Cj 

-  <h 

-    h    C2 

-  <h 

~h    % 

(32) 


I  «3     °3    °S  I  I  -  <h     °3    C3 

IX.  One  row  t)r  column  of  any  determinant  can  be  reduced  to 
unity  (Dostor's  theorem). 

This  will  need  no  more  explanation  than  the  following  illustration  : 

3461=  12  1111 1 (33) 

2  8  81  1   3   2 

6  7  9|  |8  7  6| 

X.  If  each  constituent  of  a  row  or  column  can  be  expressed  as 
the  sum  or  difference  of  two  or  more  terms,  the  determinant  can  be 
expressed  as  the  sum  or  difference  of  two  other  determinants. 

This  can  be  proved  by  expanding  each  of  the  following  determinants,  and 
rearranging  the  letters. 


<h±2> 
(h  ±r> 


K 

Cl|  = 

=  |<h 

bx  c, 

± 

t>2, 

«2| 

«2 

h  c2 

Bfc 

Csl 

\(h 

h    C3 

p  h 

q    b 
r   b 


i    Cjl. 

2  cj 

3  C3I 


(34) 


In  general,  if  each  constituent  of  a  row  or  column  consists  of  n  terms,  the 
determinant  can  be  expressed  as  the  sum  of  n  determinants. 


Example. 


-Show  by  this  theorem,  that 

Ib  +  c,    a  -  b,  a  I  =  Sabc 
c  +  a,    b  -  c,    b\ 


b3  -c*. 


a  +  b, 


XI.  The  value  of  a  determinant  is  not  changed  by  adding  to  or 
subtracting  the  constituents  of  any  row  from  the  corresponding  con- 
stituents of  one  or  more  of  the  other  rows  or  columns. 


§184. 


DETERMINANTS. 


589 


from  X.  and  III., 

<h  ±  &i>   bi>  ci  1  = 

=  1  ax   bx  cx  1  + 

£ 

^ 

»2  ±   &2»     &2>     °2  1 

I  <X2     &2    C2  1 

&> 

h 

^3  ±   &3>     &3>     C3  1 

1^3     &3    C3| 

&i 

h 

(35) 


which  proves  the  rule,  because  the  determinant  on  the  right  vanishes.     This 
result  is  employed  in  simplifying  determinants. 
Examples. — (1)  Show  II,  x,  y  +  z  1  =  0. 

1 1,  y,  z  +  x  I 

1 1,   *,    x  +  y  I 
Add  the  second  column  to  the  last  and  divide  the  result  by  x  +  y  +  z.    The 
determinant  vanishes  (3). 


(2)  Show 


x  y  z 
z  x  y 
y  z  x 


{x  +  y  +  z)  1 1  1  1 
I  z  x  y 
\y  z  x 


Add  the  second  and  third  rows  to  the  first  and  divide  by  x  +  y  +  z. 


(3)  Why  is  I  ^   bx  Cj 
I  Ojj  b2  Cg 


not  equal  to 


(4)  Show 


(h  +  K 


bx  +  av 

&3  +  «3> 


XII.  If  all  but  one  of  the  constituents  of  a  row  or  column  are 
cyphers,  the  determinant  can  be  reduced  to  the  product  of  the  one 
constituent,  not  zero,  into  a  determinant  whose  order  is  one  less  than 
the  original  determinant. 


For  example, 

la    6  1-1 
0  a,  bA 


<h  bi 


<h  h 


0    &J    &2 


'I- 


(36) 


0       0-7 

5       6-2 

-  3       1        8 

The  converse  proposition  holds.  The  order  of  a  determinant  can  be  raised 
by  similar  and  obvious  transformations. 

XIII.  If  all  the  constituents  of  a  determinant  on  one  side  of  the 
diagonal  from  the  top  left-hand  corner  are  cyphers,  the  determinant 
reduces  to  the  leading  term. 


Thus, 


The  determinant 
tuent  a,. 


\ix\ 


(h  h  <h    =  «i  I  h  <h  |  =  <hh&>         •  •      (37) 

0    U  |0    c3 

0    0    c3 
-2  % 1  is  called  the  co-factor  or  complement  of  the  oonsti- 
0    c, 


§  184.    The  Multiplication  of  Determinants. 

This  is  done  in  the  following  manner : 

!   bl  I  x  I  dx   ex  I  =  I  axdx  +  bxev   axd2  +  bxe2  I. 


'2     b2 


a2dx  +  b2ev   a2d2  +  b2e2 


(38) 


590 


HIGHEK  MATHEMATICS. 


§185. 


The  proof  follows  directly  on   expanding   the  right  side  of   the 
equation.     We  thus  obtain, 


aYdv    \e2 

d\e2  I  a 


a2  fc 
(dxe2  -  d2ex)  I  a1  \ 


+  I  bxev   aLd2 


b2ev 


a2dCj 


»2     u2 


=    a, 


*x 


Since  the  value  of  a  determinant  is  not  altered  by  writing  the 
columns  in  rows  and  the  rows  in  columns,  the  product  of  two 
determinants  may  be  written  in  several  equivalent  forms  which  all 
give  the  same  result  on  expansion.  Thus,,  instead  of  the  right 
side  of  (38),  we  may  have 

etc. 


axdx  +  \d2y  axex  +  bxe2 
Mi  +  hdv   <Vi  +  he2 

1;  \axdx 

1  Ua 

+  #2^2»     aiei  +  a2e2 
+  M2>     Vl   +   V2 

Examples.— (1)  1  Oj  bx  I2  =  1  a\  +  b\,       a^a^  +  bxb2  |. 

(2)  Multiply  1  %  6a  Cj  1  and 

1^3    *3    C3l 

*i  «i  A 

4s  e2  /a 

The  answer  may  be  written  in  several  different  forms  ;  one  form  is 

Ia1d1  +  bjj^  +  6jv  a^dz  +  bxe2  +  c^,   a^d^  +  bxez  +  Cj/j 
a^  +  62e1  +  Cg/j,  a2d2  +  b#2  +  C2/2,  Ms  +  b#3  +  c2/3 
Mi  +  Vi  +  GJv  Ma  +  Va  +  Cs/a.    Ms  +  &se3  +  csfs 
This  can  be  verified  by  the  laborious  operation  of  expansion.      There  are 
twenty-seven  determinants  all  but  six  of  which  vanish. 

When  two  constituents  of  a  determinant  hold  the  same  relative 
position  with  respect  to  the  leading  constituents,  they  are  said  to  be 
conjugate,  Thus  in  the  last  of  the  determinants  in  (34)  bY  and 
q  are  conjugate,  so  are  bz  and  c2,  r  and  cr  If  the  conjugate 
elements  are  equal,  the  determinant  is  symmetrical,  if  equal  but 
opposite  in  sign,  we  have  a  skew  determinant.  The  square  of  a 
determinant  is  a  symmetrical  determinant 

§  185.    The  Differentiation  of  Determinants. 

Suppose  that  the  constituents  of  a  determinant  are  independent 

D  =  \xY  yx\  =xYy2-x2yv 
*s   2/2 


and  that 


§186. 


DETERMINANTS. 


591 


then,     d(D)  =  xxdy2  +  y2dxx  -  x2dyl  -  yYdx2 ; 

=  (y2dxx  -  yxdx2)  +  {xxdy2  -  x2dyx) ; 

=  |^i  2/il  +  K  dVi\  •        •         •      (39) 

\dx2  y2\      \x2  dy2\ 

If  the  constituents  of  the  determinant  are  functions  of  an  in- 
dependent variable,  say  t,  then,  writing  xx  for  dx/dt}  y2  for  dyjdt 
and  so  on,  it  can  be  proved,  in  the  same  way, 


£>«K  fill  d(D)/dt=\x 
x2  V* 


Examples.— (1)  Show  that  if  D  = 


d(D)  =  I  dx1  Vl  zx  I  + 


+  K  Vi 


|*i  0i  *i| 

k>  y2  *2 

Us  2/s  *sl 

xx 

<*Vl     *1  I  + 

*» 

%2     *2 

*3 

<%3    *3I 

^2    V2     Z1 

dx3  y3  *3 

It  =  I  «,  y,  zx  I  +  I  a^  yx  »x 

«a  2/2  «a|        «a  2/2  «2 

*3    2/3    *3|         1*8    2/3    *3 


a?!  yx  dzx 

xz  y%  d»2 
*3  v*  d% 

*1  2/!  *1 
%2  2/2  *2 
*8    V%     «3 


(2)  If  Oj,  tj,  Cj,  Oa,  62»  •  •  •>  are  constants,  show  that 


d  I  OjX  bxy  c^z  I  = 

]a,jc  b2y  c%z\ 
I*    b3y  czz\ 


axdx  bxy  c^z 
a2dx  b$  c^z 

*>3y  <h» 


+  ,  etc,  =  dx  I  bxy  (^z  I,  eto. 

\b*y  w\ 


(40) 


§  186.  Jaoobians  and  Hessians. 

I.  Definitions.    If  u,  v,  w,  be  functions  of  the  independent  vari- 
ables, x,  y,  zt  the  determinant 

.      (41^ 


~bu    ~bu 

Hu 

Dx    ty 

dl 

~bv     ~bv 

Dv 

Dx    ty 

$z 

"bw    ~dw 

~dw 

~&x  7)y 

Dz 

is  called  a  Jacobian  and  is  variously  written, 

~d(u,  v,  w)  _ 

^  yf  Z}  >  or  J(U,  V,  w)  ;  or  simply  7, 


(42) 


when  there  can  be  no  doubt  as  to  the  variables  under  consideration. 
In  the  special  case,  where  the  functions  11,  v,  w  are  themselves 
differential  coefficients  of  the  one  function,  say  u,  with  respect  to 
x}  y  and  z,  the  determinant 


692 


HIGHEE  MATHEMATICS. 


§186 


Dx2 


~b2u 


bxbz 
~d2U 

~dy~dz 
~d2U 


(43) 


~dxdy 

l*u 

7)y~dx     7>y2 

~d2u      ~d2u 

'dz'bx  'bybz 

is  called  a  Hessian  of  u  and  written  H(u),  or  simply  H.  The 
Hessian,  be  it  observed,  is  a  symmetrical  determinant  whose 
constituents  are  the  second  differential  coefficients  of  u  with 
respect  to  x,  y,  z.  In  other  words,  the  Hessian  of  the  primitive 
function  u,  is  the  Jacobian  of  the  first  differential  coefficients  of  u, 
or  in  the  notation  of  (42) , 

(~bu    ~du    ^u\ 
ydx'  ~by    ^z) 


H(u)  = 
v  '  *(x,  y,  z) 

II.  Jacobia?is  and  Hessians  of  interdependent  functions. 


(44) 


If 


~bu  ~dv 

?>X  =  *  (v'lx 


~dll  ■.     3fl 


Eliminate  the  function /(«)  as  described  on  page  449. 

~du    bv        ~du    *dv 

~bx  '  ~oy      ~dy'~dx~    ' 


or 


0. 


(45) 


~du  ~du 

~dx  ~oy 

7)v  *bv 

~dx  "by 

That  is  to  say,  ifu  is  a  function  of  v,  the  Jacobian  of  the  functions 
of  u  and  v  with  respect  to  x  and  y  will  be  zero. 

The  converse  of  this  proposition  is  also  true.  If  the  relation  (45) 
holds  good,  u  will  be  a  function  of  v. 

In  the  same  way,  it  can  be  shown  that  only  when  the  Hessian 
ofuis  not  equal  to  zero  are  the  first  derivatives  of  u  with  respect  to 
x  and  y  independent  of  each  other. 

Examples.— (1)  If  the  denominators  of  (9),  et  seq.,  page  453,  that  is,  if  P,  Q, 
and  B  vanish,  show  that  u  can  be  expressed  as  a  function  of  -y,  or,  u  and  v  are 
not  independent.  Ansr.  The  expression  is  a  Jacobian.  If  u  is  a  function  of 
v,  the  Jacobian  vanishes.  B  vanishes  if  either  u  or  v  is  a  function  of  z  only  ; 
P,  Q,  and  B  all  vanish  if  u  is  a  function  of  v ;  and  f(u,  v)  =  0  can  be  re- 
presented by  v  =  c  which  contains  no  arbitrary  function. 


§186. 


DETERMINANTS. 


693 


(2)  Show  that  ~,  '    '    .  =  0  is  a  condition  that  *  =  0  shall  be  an  in- 
o\Jy,  y,  ») 

tegral  of  Pdz/dx  +  Qdz/dy  =  B.    Hint.  <p  is  a  function  of  x,  y,  z,  and  can  be 
expressed  as  a  function  of  u  and  v. 

(8)  If  P,  Q,  and  B  are  given,  the  Jacobian  of  u  and  v  must  be  pro- 
portional to  P,  Q  and  B.     This  follows  from  the  equations  on  page  453. 

III.  The  Jacobian  of  a  function  of  a  function.     If  uv  u2)  are 
functions  of  x1  and  x2t  and  xY  and  x2  are  functions  of  y1  and  y2, 


bux 


bux    bx1       bu^    bx2 .  bu±  _  d'^    bx±       b^    ()iC2 


da^    ^2/j       bx2    by1     by2      bxx    by 


bx* 


1 


By  the  rule  for  the  multiplication  of  determinants, 


bu-±  bu-L 

an 

bux  bux 

bx-^   bxx 

tyl    <^2 

bxx  by1 

*Vl     <^2 

bu2   bu2 

bu2   bu2 

~bx2   bx2 

tyi  ty* 

bx2  by2 

Wi    W2 

(46) 


b(uv  u2)  ^  <>K,  u2)    b(xv  x2) 
%i>  2/2)      ^0*1.  «a) '  <%i>  2/2)' 
This  bears  a  close  formal  analogy  with  the  well-known 

bu  _  bu   by 
bx~  by  '  bx' 
IV.  The  Jacobian  of  implicit l  functions.     If  u  and  v,  instead  of 
being  explicitly  connected  with  the  independent  variables  x  and  y, 
are  so  related  that 

V  mMx>  V>  u>  v)  =  0;  q  =  /2(ff,  y,  u,  v)  =  0, 
u  and  v  may  be  regarded  as  implicit  functions  of  x  and  y.     By 
differentiation 

bu    bu       bv    bv        _      bp       bp    bu 

by 

bu 

by  "*  bv 

and  by  the  rule  for  the  multiplication  of  determinants, 

(47) 


bx 

bp       bp 


bu        bp    bv 

—  -+■  -=- .  —  =  0  • 

bx        bv    bx  y 


by       bu 


M 


ix  + 


h 

bu 


bu 
IX  + 


bq 
bv 


Dv 


bq 
bx**0*  by  + 


bq 

bu 


bp 

bv 


+  *v 


0; 


bv 
by 

bx      U' 


bp   bp 

X 

bu  bv 

=  - 

bp   bp 

bu  bv 

bx   bx 

bx  by 

bu   bv 

bu   bv 
by  by 

bg  jr 

bx  by 

1 A  function  is  said  to  be  explicit  when  it  can  be  expressed  directly  in  terms  of 
the  variable  or  variables,  e.g.,  z  is  an  explicit  function  of  a;  in  the  expression  :  z  =  x*; 
z  +  a  =  bx*.  A  function  is  implicit  when  it  cannot  be  so  expressed  in  terms  of  the 
independent  variable.    Thus  x2  +  xy  =  y2  ;  x  +  y  =  0*,  are  implicit  functions. 

PP 


594  HIGHER  MATHEMATICS.  §  187. 

0r  Xp>  g)Hu,  v)  _  _  ^(p,  q) 

1)(u,  v)"d(x,  y)         5(5;  y)' 
A  result  which  may  be  extended  to  include  any  number  of  inde- 
pendent relations. 

§  187.  Illustrations  from  Thermodynamics. 

Determinants,  Jacobians  and  Hessians  are  continually  appear- 
ing in  different  branches  of  applied  mathematics.  The  following 
results  will  serve  as  a  simple  exercise  on  the  mathematical  methods 
of  some  of  the  earlier  sections  of  this  work.  The  reader  should 
find  no  difficulty  in  assigning  a  meaning  to  most  of  the  coefficients 
considered.  See  J.  E.  Trevor,  J  own.  Phys.  Chem.r39  523,  573, 
1899 ;  10,  99,  1906  ;  also  E.  B.  Baynes'  Thermodynamics,  Oxford, 
95,  1878. 

If  U  denotes  the  internal  energy,  <£  the  entropy,^  the  pressure, 
v  the  volume,  T  the  absolute  temperature,  Q  the  quantity  of  heat  in 
a  system  of  constant  mass  and  composition,  the  two  laws  of  thermo- 
dynamics state  that 

dQ  =  dU  +  p.dv;  dQ  =  Td<j>, ...      (1) 

pages  80  and  81.     To  find  a  value  for  each  of  the  partial  derivatives 
(14>\     /ty\      /7>4>\      (o<f>\     (ocf>\     /ty\ 

IspJ:  vw  M;  w):  w;  w; 

/0V\       fdv\        /0V\      fbv\       fdv\      fdv\ 

v^y;  \*ph  \w;  wv  WA/Wa' 

in  terms  of  the  derivatives  of  U. 

I.  When  v  or  <f>  is  constant.     From  (1), 

-pm  lUftv;   and  T  =  ~dU/o<f>.  .  .  (2) 

First,  differentiate  each  of  the  expressions  (2),  with  respect  to  <f>  at 
constant  volume. 

/op\         7)*U  /oT\       7>*U  /Q. 

"  [foj.  "  SSK* ;  and  \b+)9  =  W  '       '       (3) 
o2U 

By  division,  -  (^J  -  ^j (4) 

Next,  differentiate  each  of  equations  (2)  with  respect  to  v  at  constant 
entropy. 


§  188.  DETEKMINANTS.  695 

By  division,  -   Qj)+  =  J£.  .         .         .         (6) 

~dcf>dv 

II.  When  either  p  or  T  is  constant.     We  know  that 

dp  -  s£fc  +  ^~dp;    ana  dT  -  jjfo  +  ^>.  .        (7) 

First,  when  p  is  constant,  eliminate  dv  or  d<£  between  equations 
(7).     Hence  show  that 

dv     _d<f>  _d6 
^p      Tp    ,  J1 
<)<£       7)v 
where  J  denotes  the  Jacobian  ~d(p,  Tjl'd^v,  <f>).     If  H  denotes  the 
Hessian  of  U,  show  that 

AchA    _  o2U  ^2Tj  ^r/ 

dfl2 

Finally,  if  T  is  constant,  show  that 

d2*7  W  vu 

§  188.  Study  of  Surfaces. 

Just  as  an  equation  of  the  first  degree  between  two  variables 
represents  a  straight  line  of  the  first  order,  so  does  an  equation  of 
the  first  degree  between  three  variables  represent  a  surface  of  the 
first  order.     Such  an  equation  in  its  most  general  form  is 

Ax  +  By  +  Gz  +  D  =  0, 
the  equation  to  a  plane. 

An  equation  of  the  second  degree  between  three  variables  re- 
presents a  surface  of  the  second  order.  The  most  general 
equation  of  the  second  degree  between  three  variables  is 

Ax2  +  By2  +  Gz2  +  Dxy  +  Eyz  +  Fzx  +  .  .  .  +  N  =  0. 
All  plane  sections  of  surfaces  of  the  second  order  are  either  circular, 
parabolic,  hyperbolic,  or  elliptical,  and  are  comprised  under  the 
generic  word  conicoids,  of  which  spheroids,  paraboloids,  hyperboloids 
and  ellipsoids  are  special  cases. 

J.  Thomson  (Phil.  Mag.,  43,  227,  1871)  developed  a  surface  of 

PP* 


HIGHER  MATHEMATICS. 


§188. 


the  second  degree  by  plotting  from  the  gas  equation 
f(p,v,  T)  =  0;  or pv  =  BT, 

by  causing  p,  v  and  T  to  vary  simultaneously.  The  surface  pabv 
(Fig.  175)  was  developed  in  this  way. 

Since  any  section  cut  perpendicular  to  the  T-  or  0-axis  is  a 
rectangular  hyperbola,  the  surface  is  a  hyperboloid.  The  iso- 
thermals  T,  T2,  T3,  .  .  .  (Fig.  29,  page  111)  may  be  looked  upon  as 
plane  sections  cut  perpendicular  to  the  0-axis  at  points  correspond- 
ing to  Tv  T2,  .  .  .,  and  then  projected  upon  the  _py-plane.  In 
Fig.  176,  the  curves  corresponding  to  pv  and  ab  have  been  so 
projected. 

As  a  general  rule,  the  surface  generated  by  three  variables  is 
not  so  simple  as  the  one  represented  by  a  gas  obeying  the  simple 
laws  of  Boyle  and  Charles. 

Yan  der  Waals'  "\J/"  surfaces  are  developed  by  using  the 
variables  \j/}  x,  v,  where  \J/  denotes  the  thermodynamic  potential  at 


Fig.  175.— /^-surface.  Fig.  176.— Two  Isothermals. 

constant  volume  (U  -  Tq)  ;  x  the  composition  of  the  substance; 
v  the  volume  of  the  system  under  investigation.  The  "i/^"  surface 
is  analogous  to,  but  not  identical  with,  pabv  in  the  above  figure. 

The  so-called  thermodynamic  surfaces  of  Gibbs  are  obtained 
in  the  same  way  from  the  variables  v,  U,  tf>  (where  v  denotes  the 
volume,  U  the  internal  energy,  and  0  the  entropy)  of  the  given 
system. 

The  solubility  of  a  double  salt  may  be  studied  with  respect  to 
three  variables — temperature,  0,  and  the  concentrations  s1  and  s2  of 
each  component  in  the  presence  of  its  own  solid.     Thus  a  mixed 


§188. 


DETERMINANTS. 


597 


solution  of  magnesium  sulphate,  MgS04,  and  potassium  sulphate, 
K2S04,  will  deposit  the  double  salt,  MgS04.K2S04.6H20,  under 
certain  conditions.     The  surface  so  obtained  is  called  a  surface 
of  solubility.     The  solution  can  also  deposit  other  solids  under 
certain  conditions.     For  example,  we  may  also  have  crystals  of 
MgS04.7H20  deposited   in  such  a  manner  as   to  form   another 
surface  of  solubility.     This  is  not  all.     The  above  system  may 
deposit   crystals    of  the   hydrate   MgS04.6H20,  the   double   salt 
MgS04.K2S04.4H20,    or   the   separate    components.      The    final 
result  is  the  set  of  surfaces  shown  in  Fig.  177. 
The  surface  which  is  represented  by  the  equation, 
J\x,  y,z)  =  0;  ot,z  =  <$>{x,  y), 
will  exhibit  the  characteristic  properties  of  any  substance  with 
respect  to  the  three  variables  x,  y,  and  z.     The  surface,  in  fact, 


Pig.  177. 

will  possess  certain  geometrical  peculiarities  which  depend  upon 
the  nature  of  the  substance.  It  is  therefore  necessary  to  be  able 
to  study  the  nature  of  the  surface  at  any  point  when  we  know  the 
equation  of  the  surface. 

I.  The  tangent  line,  and  tangent  'plane.    Let  the  point  P(xv  yl9  zY) 
be  upon  the  surface 

u  =  f(x,y,z) (1) 

The  equations  of  a  line  through  the  point  P  are,  page  131, 
x-xx  __  y-y1  _  z-zx 
I      ' 


=  r, 


m  n 

and  where  the  line  meets  the  surface 

u  =  f(xi  +  ^r»  Vi  +  mr»  z\  +  nr)  =  o. 
By  Taylor's  theorem, 

/  du-         du  du      \  r2/   d  d  d  \2 

\  dxx         dyx         dzx      )  2  \  dxx         dyx         dzj  U  + 


=  0. 


(2) 
(3) 

(4) 


598  HIGHER  MATHEMATICS.  §  188. 

One  value  of  r  must  be  zero  since  P  is  on  the  surface ;  and  if  we 
choose  the  line  so  that 

7du         du  du 

^+w^  +  ra^1-°>  ;•    •    •  <5> 

another  value  of  r  will  vanish  ;  so  that  for  this  direction  another 
point,  Q,  will  coincide  with  P  and  the  line  will  be  a  tangent  line. 
Equation  (6)  gives  the  relation  between  the  direction  cosines  of  a 
tangent  line  to  the  surface  at  the  point  P(xv  yv  Zj).  Eliminating 
I,  m,  and  n,  between  (2)  and  (5),  we  get 

,  .~du       .  ,~bu       .  .^U 

(*-*^^-^  +  (*-^  =  °-       •    <6> 

This  equation  being  of  the  first  degree  in  x,  y,  and  z  represents  a 
plane  surface.     All  the  tangent  lines  lie  in  one  plane.     Equation 
(6)  is  the  equation  of  a  tangent  plane  at  the  point  (xv  yv  zj. 
If  the  surface  had  the  form 

z=f(x>y)  ....       (7) 

equation  (6)  would  have  been 

at  the  point  {xv  yv  zj. 

Examples. — (1)  Show  that  the  tangent  plane  of  the  sphere  xx* + yx2  +  zx2 = r2 
at  the  point  (x,  y,  z)  is  xx1+yy1  +  zz1=r*.  Hint.  duldxl  =  2x1;  >bujdyi  =  2y1; 
dujdzi = 2zv  Substitute  in  (6)  and  it  follows  that  xxl  +  yyx  +  zzx  =  x2+y2  +  z2 =r2. 

(2)  The  equation  of  the  tangent  plane  at  the  point  (xv  yv  zx)  on  the  para- 
boloid 

£  +  %  m  4px ;  is  ^  +  ^  =  2p(z  +  zx). 

(3)  Show  that  the  tangent  plane  to  the  surface  (1)  or  (6)  above  is  hori- 
zontal, that  is,  parallel  to  the  £?/-plane.  z-zx  —  0.  Hint.  When  a  line  is 
parallel  to  the  sc-axis,  the  angle  it  makes  with  the  axis  is  zero,  and  tan  0°=  0, 
hence  we  must  have  'dul'dx  and  'duj'dy  both  zero ;  and  'duj'dz  not  zero. 

II.  The  normal.  By  analogy  with  (1),  page  106,  or  by  more 
workmanlike  proofs  which  the  student  can  discover  for  himself,  we 
can  write  the  condition  that  a  plane  normal  to,  or  perpendicular  to, 
the  tangent  of  the  surface  f(x,  y,  z)  0  at  the  point  (xv  yv  z^)  is 

ZZ*l  =  Uz3b  =  lZll  0-      ov^^  =  y-^z-z  (9) 

~du  ~bu  7)u       *         '     7)u  ~du    '        i*  *  ' 

~dx  ~by  ^z  ~bx  ~by 

Examples.— (1)  Show  that  the  normal  to  the  sphere  x2  +  y2  +  Z*  =  r2,  is 
xjxx  «  y\yx  =  z\zx.    Use  the  results  of  Ex.  (1)  above,  and  substitute  in  (10). 


§  188.  DETERMINANTS.  599 

(2)  The  normal  to  the  surface  xyz  =  a?  at  the  point  (xv  ylt  zj  is 
xxx  -  xx2  =  xy1  -  y^  =  zzx  -  z?. 

(3)  Show  that  the  equations  of  the  normal  and  of  the  tangent  to  the 
curve  y2  =  2x  -  x2 ;  z2  =  42  -  2x,  at  the  point  (2,  3,  -  1)  are  respectively 
*  -  2  =  i(V  -  3)  =  z  +  1 ;  and  z  -  2  =  -  3{y  -  3)  =  z  +  1. 

(4)  We  do  not  know  the  characteristic  equation  connecting  p,  v,  and  T. 
If  the  substance  is  an  ideal  gas,  we  h&vepv  =  BT.  From  equations  (13)  and 
(15),  pages  81  and  82,  we  get  the  fundamental  equation 

dU=Td<f>-pdv (10) 

connecting  v,  U,  and  (p.  "This  expression  is  the  differential  equation  of  some 
surface  of  the  form 

Where  -  1,  and  the  two  partial  derivatives  are  proportional  to  the  direction 
cosines  of  the  normal  to  the  surface  at  any  point.  Again,  it  follows  from  (10) 
and  (11)  that 

(§).-*- feV-*    •  •  •  <12> 

In  other  words  the  direction  cosines  of  the  normal  at  any  point  on  the  surface 
are  proportional  to  T,  -p,  and  -1  respectively.  Hence,  v,  U,  and  <p  are  the 
coordinates  of  the  point  on  the  surface,  and  the  remaining  pair  of  variables  p 
and  T  are  given  by  the  direction  cosines  of  the  normal  at  the  same  point. 
The  whole  five  quantities  p,  v,  T,  U,  and  <p  can  thus  be  represented  in  a  very 
simple  manner. 

III.  Inflexional  tangents.  We  can  discuss  the  equations  of  a 
surface  by  the  aid  of  the  extension  of  Taylor's  theorem,  on  page 
292,  and  the  methods  described  in  §  101.  There  are  an  infinite 
number  of  lines 

IT       ~nT  ~     n    '      '         '         '       (idJ 
which  satisfy  the  relations 

P  +  m*  +  »2  =  1 ;  l~x  +  «5jf-»  =  0.  •       (14) 

These  lines  cut  the  surface  at  two  coincident  points ;  two  of  these 
lines  also  satisfy  the  relation 

lW  +  2lm*xSy  +  mW  =  Q-  •         ■       (15) 

These  two  lines  cut  the  surface  at  three  coincident  points.  These 
lines  are  called  inflexional  tangents.  They  are  real  and  distinct, 
coincident,  or  imaginary,  according  as  the  quadratic  in  I,  and  m, 

zj*  +  2zjm  +  zmm?  .  .  .  (16) 
has  real  and  different,  double,  or  imaginary  roots.  This  depends 
upon  whether 


600 


HIGHER  MATHEMATICS. 


§188. 


V*    Vm     /V,\*      f+=cu8p; 


(17) 


Each  Inflexional  tangent  to  the  surface  will  cut  the  curve  in  three 
coincident  points  at  the  point  of  contact.  The  inflexional  tangents 
have  a  closer  contact  with  the  surface  than  any  of  the  other 
tangent  lines.  At  the  point  of  contact  of  the  curve  with  the  tan- 
gent plane  there  will  be  a  cusp,  conjugate  point,  or  a  node,  ac- 
cording as  the  above  expression  is  positive,  zero,  or  negative. 

The  equations  of  the  inflexional  tangents  at  the  point  (xv  yv  zj 
are  obtained  by  the  elimination  of  I,  m,  and  n  between  (13) ;  the 
second  of  equations  (14) ;  and  (15).  If  we  treat  the  equation  of 
the  surface 

u=f(x,y}z)  =  0,  .        .        .      (18) 

in  a  similar  manner,  we  find  that  we  must  know  the  value  of 


•    dsc         ty         ^z 


as  well  as  of 


nTz)u  =  °> 


(19) 


(20) 


~dx  ty 

in  order  to  determine  the  inflexional  tangents.  These  tangents  will 
be  real  and  different,  coincident,  or  imaginary  according  as  the 
determinant 

(21) 


»* 

«*« 

ux 

un 

um 

uv 

uvt 

u» 

uz 

uv 

u, 

% 

ux 

is  negative,  positive,  or  zero.  When  the  inflexional  tangents  are 
imaginary,  the  surface  is  either  convex  or  concave  at  the  point,  and 
conversely.     Furthermore,  if 


u„  u 


and,  U 


(22) 


are  positive  at  the  point  P(%,  y,  z)  the  surface  is  concave  provided 
Wufbx2  and  'du/'dz  have  the  same  sign ;  and  convex  when  'd2u/'dx2 
and  7)u[dz  have  different  signs. 


APPENDIX  I. 

COLLECTION  OF  FORMULA  AND  TABLES  FOR  REFERENCE. 

"  When  for  the  first  time  I  have  occasion  to  add  five  objects  to  seven 
others,  I  count  the  whole  lot  through ;  but  when  I  afterwards 
discover  that  by  starting  to  count  from  five  I  can  save  myself 
part  of  the  trouble,  and  still  later,  by  remembering  that  five  and 
seven  always  add  up  to  twelve,  I  can  dispense  with  the  counting 
altogether."— E.  Maoh. 

§  189.  Calculations  with  Small  Quantities. 

The  discussion  on  approximate  calculations  in  Chapter  V.   renders  any 

further  remarks  on  the  deduction  of  the  following  formula  superfluous. 
For  the  sign  of  equality  read  "  is  approximately  equal  to,"  or  "  is  very 

nearly  equal  to  ".     Let  a,  0, 7, .  . .  be  small  fractions  in  comparison  with  unity 

or  a;: 

(1  ±  «)  (1  ±  0)  -  1  ±  a  ±  0 (1) 

(1  +  a)  (1  ±  0)  (1  ±  7) .  . .  =  1  ±  a  ±  0  ±  7  ± (2) 

(1  ±  a)'2  =  1  ±  2a;  (1  ±  a)»  =  1  ±  na (3) 

,J(l  +  a)  =  l+la.      N/S8-i(a  +  /B) (4) 

FTa)  *  X  +  a  ;  {T+^T  ==1+na>  V(l  +  «)  =  1  ~  *"'           '  (6) 

(1  ±  «)  (1  ±  0)       -    .            .  _      _  ,  tRX 

(1  ±  7)  (1  ±  8)  =  1  ±  "  ±  *  +  7  +  8 (6) 

The  third  member  of  some  of  the  following  results  is  to  be  regarded  as  a 

second  approximation,  to  be  employed  only  when  an  exceptional  degree  of 

accuracy  is  required. 

0o.  =  1  +  a ;  w  =*  1  +  a  log  a (7) 

log  (1  +  a)  =  a  =  a  -  $a? (8) 

log  (a;  +  a)  =  logo;  +  o/a?  -  ioa/a>2 (9) 


.  a;  +  a      2a      2 


l0g^=*+ri* <10> 

By  Taylor's  theorem,  §  98, 

sin  (a;  +  0)  =  sin  x  +  0  cos  x  -  %02  sin  x  -  ££3cos  x  +  . . . 

If  the  angle  3  is  not  greater  than  2£°,  £  <  -044 :  ££2  <  -001 ;  %03  <  -00001. 

But  sin  x  does  not  exceed  unity,  therefore,  we  may  look  upon 

sin  (a;  +  0)  =  sin  x  +  0  cos  x, 

601 


602  HIGHER  MATHEMATICS.  §  190. 

correct  up  to  three  decimal  places.     The  addition  of  another  term  "  -  £/32  " 

will  make  the  result  correct  to  the  fifth  decimal  place. 

sin  a  =  a  =  o(l  -  W) ',  coso  =  1  =  1  -  £a2 (11) 

sin  (x  ±  j8)  =  sin  x  ±  p  cos  x ;  cos  (x  ±  0)  =  cos  x  ±  j8  sin  x.   .       (12) 
tan  o  =  a  =  a(l  +  $a2) ;  tan  (x  +  0)  =  tan  x  ±  0  sec2a;.      .        .       (13) 

Example. — Show  that  the  square  root  of  the  product  of  two  small 
fractions  is  very  nearly  equal  to  half  their  sum.  See  (4).  Hence,  at  sight, 
s/ 24-00092  x  24-00098  =  24-00095. 


§  190.  Permutations  and  Combinations. 

Each  arrangement  that  can  be  made  by  varying  the  order  of  some  or  all 
of  a  number  of  things  is  called  a  permutation.  For  instance,  there  are  two 
permutations  of  two  things  a  and  6,  namely  ab  and  ba ;  a  third  thing  can  be 
added  to  each  of  these  two  permutations  in  three  ways  so  that  abc,  acb,  cab, 
bac,  bca,  cba  results.  The  permutations  of  three  things  taken  all  together  is, 
therefore,  1x2x3;  a  fourth  thing  can  occupy  four  different  places  in  each 
of  these  six  permutations,  or,  there  are  1x2x3x4  permutations  when  four 
different  things  are  taken  all  together.  More  generally,  the  permutations  of 
n  things  taken  all  together  is 

n(n  -  1)  (n  -  2)  .  .  .  3.2.1  =n\ 
n !  is  called  "  factoral  n".1    It  is  generally  written  j  n. 

Using  the  customary  notation  „Pn  to  denote  the  number  of  permutations 
of  n  things  taken  nata  time, 

number  of  things  ■*    number  of  things  taken  —  n±n  =  n  ' 

If  some  of  these  n  things  are  alike,  say  p  of  one  kind,  q  of  another,  r  of 
another, 

n\ 

nPn==p\ql  r\ (2) 

If  only  r  of  the  n  things  are  taken  in  each  set, 

nPr  =  n(n-l)(n-  2) . . .  (n  -  r  +  1)  =  ^^j-  .        .        (3) 

Each  set  of  arrangements  which  can  be  made  by  taking  some  or  all  of  a 
number  of  things,  without  reference  to  the  internal  arrangement  of  the  things 
in  each  group,  is  called  a  combination.  In  permutations,  the  variations,  or 
the  order  of  the  arrangement  of  the  different  things,  is  considered;  in  com- 
binations, attention  is  only  paid  to  the  presence  or  absence  of  a  certain  thing. 
The  number  of  combinations  of  two  things  taken  two  at  a  time  is  one,' because 
the  set  ab  contains  the  same  thing  as  ba.  The  number  of  combinations  of 
three  things  taken  two  at  a  time  is  three,  namely,  ab,  ca,  be ;  of  four  things, 
ab,  ae,  ad,  be,  bd,  cd.  But  when  each  set  consists  of  r  things,  each  set  can  be 
arranged  in  r  I  different  ways. 

1  It  is  worth  remembering  that  n  !  =r(n  +  1),  the  gamma  function  of  §  136.  When 
n  is  very  great 

n  !  =  nne  -  »ij  2irn, 
known  as  Stirling's  formula.    This  allows  n  !  to  be  evaluated  by  a  table  of  logarithms. 
The  error  is  of  the  order  -fa71  °f  ^ne  vahie  of  n  ! 


§  191.  APPENDIX  I.  603 

Let  nGr  denote  the  number  of  combinations  of  n  things  taken  r  at  a  time. 
We  observe  that  the  nCr  combinations  will  produce  nCr  x  **  1  permutations. 
This  is  the  same  thing  as  the  number  of  permutations  of  n  things  in  sets  of  r 
things.    Hence,  by  (3), 

nPr_  n(n-  l)(n-2)...(n-r  +  l) 
"°'=7l T\  '       '        '        W 

nCr  =  r  I  (n  -  r)  ! l&} 

Nearly  all  questions  on  arrangement  and  variety  can  be  referred  to  the 
standard  formulae  (3)  and  (5).  Special  cases  are  treated  in  any  text-book  on 
algebra.  In  spite  of  the  great  number  of  organic  compounds  continually 
pouring  into  the  journals,  chemists  have,  in  reality,  made  no  impression  on 
the  great  number  which  might  exist.  To  illustrate,  Hatchett's  (Phil.  Trans., 
93,  193,  1803)  has  suggested  that  a  systematic  examination  of  all  possible 
alloys  of  all  the  metals  be  made,  proceeding  from  the  binary  to  the  more 
complicated  ternary  and  quaternary.  Did  he  realize  the  magnitude  of  the 
undertaking  ? 

Examples. — (1)  Show  that  if  one  proportion  of  each  of  thirty  metals  be 
taken,  435  binary,  4,060  ternary  and  27,405  quaternary  alloys  would  have  to 
be  considered. 

(2)  If  four  proportions  of  each  of  thirty  metals  be  employed,  show  that 
6,655  binary,  247,660  ternary  and  1,013,985  quaternary  alloys  would  have  to  be 
investigated. 

The  number  of  possible  isomers  in  the  hydrocarbon  series  involving  side 
chains,  etc.,  are  discussed  in  the  following  memoirs :  Cayley  (Phil.  Mag.  [4], 
13, 172,  1857 ;  47,  444, 1874 ;  or,  B.  A.  Reports,  257,  1875)  first  opened  up  this 
question  of  side  chains.  See  also  O.  J.  Lodge  (Phil.  Mag.,  [4],  50,  367,  1875), 
Losanitsch  (Ber.,  30, 1,917,  1897),  Hermann  (ib„  3,428),  H.  Key  (i&.,33, 1,910, 
1900),  H.  Kauffmann  (ib.,  2,231). 

§  191.  Mensuration  Formulae. 

Reference  has  frequently  been  made  to  Euclid,  i.,  47 — Pythagoras'  theorem. 
In  any  right-angled  triangle,  say,  Fig.  184, 

Square  on  hypotenuse  =  Sum  of  squares  on  the  other  two  sides.  (1) 
Also  to  Euclid,  vi.,  4.  If  two  triangles  ABC  and  DEF  are  equiangular  so 
that  the  angles  at  A,  B,  and  C  of  the  one  are  respectively  equal  to  the  angles 
D,  E,  and  F  of  the  other,  the  sides  about  the  equal  angles  are  proportional — 
Rule  of  similar  triangles— so  that 

AB:BC  =  DE:EF;  BC  :OA  =  EF :  FD;  AB  :  AC  =  BE  :  DF. 
w  =  3-1416,  or,  ^,  or,  180° ;  0  =  degrees  of  arc ;  r  denotes  the  radius  of  a  circle. 

L  Lengths  (arcs  and  perimeters). 

Chord  of  Circle  (angle  subtended  at  centre  0)  =  1r  sin  £0.         ,  (2) 

Arc  of  Circle  (angle  subtended  6)  =  -j-^tt.          .        .        •        .  (3) 

Perimeter  of  Circle  =  2wr  =  *  x  diameter.        ....  (4) 

Perimeter  of  Ellipse  (semiaxes,  a,  b)  =  2ir  V^(a2  +  fe2).    .         .  (5) 

Triangle,    a2  =  6s  +  c2  -  2&c  cos  A.       .        .       .       .        .        -  (5a) 


604 


HIGHEE  MATHEMATICS. 


§  191. 


II.  Areas. 

Rectangle  (sides  a,  b)  =  a:b 

Parallelogram  (sides  a,  b ;  included  angle  6)  =  ab  sin  0. 

Rhombus  =  £  product  of  the  two  diagonals 

Triangle  (altitude  h ;  base  b)  "] 

=  %h.  b  —  %ab  sin  G  =  J  s(s  -a)  (s  -  b)  (s  -  c),  J 

where  a,  b,  c,  are  the  sides  opposite  the  respective  angles  At  B,  O,  of  Fig. 
178,  and  *  =  %{a  +  b  +  c). 

Spherical  Triangle  =  (A  +  B  +  G  -  v)r\    .  ...      (10 


B\  /  \    !e 

D 

Fig.  178.  Fig.  179.— Spherical  Triangle. 

where  r  is  the  radius  of  the  sphere,  A,  B,  C,  are  the  angles  of  the  triangle 

(Fig.  179). 

Trapezium  (altitude  h;  parallel  sides  a,  b)  =  $h(a  +  b).         .        .      (11) 
Polygon  op  n  Equal  Sides  (length  of  side  a)  =  £na2  cot  ^~.    .      (12) 

Circle  =  wr2  =  %n  x  diameter (13) 

Circular  Sector  (included  angle  0)  =  £  arc  x  radius  =  -^Trdr%.      (14) 
Circular  Segment = area  of  sector  -  area  of  triangle  ^wr2  -  |r2  sin  d  (15) 

The  triangle  is  made  by  joining  the  two  ends  of  the  arc  to  each  other  and 

to  the  centre  of  the  circle.     6  is  angle  at  centre  of  circle. 
Parabola  cut  off  by  Double  Ordinate  (2y)  =  %xy; 

=  $  Area  of  parallelogram  of  same  base  and  height.  J      *     ' 

Ellipse  =  ira.b (17) 

Curvilinear  and  Irregular  Figures.     See  Simpson's  rule. 

Similar  Figures.    The  areas  of  similar  figures  are  as  the  squares  of  the 

corresponding  sides.     The  area  of  any  plane  figure  is  proportional  to  the  square 

of  any  linear  dimension.    E.g.,  the  area  of  a  circle  is  proportional  to  the  square 

of  its  radius. 

m.  Surfaces  (omit  top  and  base). 

Sphere  =  4wr2. (18) 

Cylinder  (height  h)  =  Imrh (19) 

Prism  (perimeter  of  the  base  p)  —  ph (20) 

Cone  or  Pyramid  =  \p  x  slant  height. (21) 

Spherical  Segment  (height  h)  —  2vrh. (22) 


§19L 


APPENDIX  I. 


605 


IV.  Volumes. 

Rectangular  Parallelopiped  (sides  a,  b,  c)  =*  a.b.c. 
Sphere  =  |  circumscribing  cylinder ; 

=  fur3  =  4-189r3  =  **  diameter3. 
Spherical  Segment  (height  h)  =  ^7r(3r  -  h)  W.     .        , 
Cylinder  or  Prism  =  area  of  base  x  height  =  Trr'%. 
Cone  or  Pyramid  =  $  circumsoribing  cylinder  or  prism ; 

=  area  of  base  x  £  height  =  %irr2h  =  r047r'27t. 

Frustum  op  Right  Circular  Cone  =  \irh(a?  +  ab  +  b2).  \ 

a  and  6  are  radii  of  circular  ends.  J       0*®) 

Similar  Figures.  The  volumes  of  similar  solids  are  as  the  cubes  of 
corresponding  sides.  The  volume  of  any  solid  figure  is  proportional  to  the 
cube  of  any  linear  dimension.  E.g.,  the  volume  of  a  sphere  is  proportional 
to  the  cube  of  its  radius. 


(23) 

(24) 

(25) 
(26) 

(27) 


V.  Centres  of  Gravity. 

Plane  Triangular  Lamina.    Two-thirds  the  distance  from  the  apex  of 
the  triangle  to  a  point  bisecting  the  base. 

Cone  or  Pyramid.    Three-fourths  the  distance  from  the  apex  to  the 
centre  of  gravity  of  base. 

Bayer's  "  strain  theory  "  of  carbon  ring  compounds  has  attracted  some 
attention  amongst  organic  chemists.  It  is  based  upon  the  assumption  that 
the  four  valencies  of  a  carbon  atom  act  only  in  the  directions  of  the  lines 
joining  the  centre  of  gravity  of  the  atom  with  the  apices  of  a  regular  tetra- 
hedron. In  other  words,  the  chemical  attraction  between  any  two  such  atoms 
is  exerted  only  along  these  four  directions.  When  several  carbon  atoms  unite 
to  form  ring  compounds,  the  "  direc- 
tion of  the  attraction  "  is  deflected. 
This  is  attended  by  a  proportional 
strain.  The  greater  the  strain,  the 
less  stable  the  compound.  Apart 
from  all  questions  as  to  the  validity 
of  the  assumptions,  we  may  find  the 
angle  of  deflection  of  the  "direc- 
tions of  attraction  "  for  two  to  six 
ring  compounds  as  an  exercise  in 
mensuration. 

I.  To  find  the  angle  between 
these  "directions  of  attraction"  at 
the  centre  of  a  carbon  atom  assumed 
to  have  the  form  of  a  regular  tetra- 
hedron.   Let  s  be  the  slant  height,  Fig.  180. 
AB,  or  BC,  of  a  regular  tetrahedron  (Fig.  180) ;  h  =  EC,  the  vertical  height; 
/,  the  length  of  any  edge,  DC  or  AC;  <p,  the  angle  DOC,  or  AOC,  made 
by  the  lines  joining  any  two  apices  with   the  centre  of  the  tetrahedron. 
...  8i  +  (£J)2  =  J2.  S2  =  sp#     But  h  divides  s  in  the  ratio  2 : 1,  hence  (§  191), 


606  HIGHER  MATHEMATICS.  §  192. 

h?=P-  (f s)2  =  -|Z2.  But  CD  =  2BD  =  I;  BC  =  AB  =  s  ;  CE  =  h.  Hence, 
h  =  J  §Z2.  From  a  result  in  V,  above,  the  middle  of  the  tetrahedron  cuts  CE 
at  O  in  the  ratio  3  : 1. 

.-.  sin  %<p  =  %AC  -j-  OC  =  JI/3*  ;  or,  <f>  =  109°  28'.     .        .      (29) 

II.  To  find  the  angle  of  deflection  of  the  "  direction  of  attraction  "  when  2 
to  6  carbon  atoms  form  a  closed  ring.  From  (29),  for  acetelyne  H2C  |  CH2> 
the  angle  is  deflected  from  109°  28'  to  £(109°  28'),  or  55°  44'.  For  tri- 
methylene,  assuming  the  ring  is  an  equilateral  triangle,  the  angle  is  deflected 
£(109°  28' -60°)  =  24°  44'.  For  tetramethylene,  assuming  the  ring  is  a  square, 
the  angle  of  deflection  is  £(109°  28'  -  90°)  =  9°  34'.  For  pentamethylene,  assum- 
ing the  ring  to  be  a  regular  pentagon,  the  angle  of  deflection  is  £(109°  28'  - 108°), 
or  0°  44'.  For  hexamethylene,  assuming  the  ring  is  a  regular  hexagon,  the 
angle  of  deflection  is  £(109°  28' -120°),  or  -5°  76'. 

The  value  of  the  angle  6,  in  Fig.  180,  is  70°  32'.  See  H.  Sachse  "  On  the 
Configuration  of  the  Polymethylene  Ring,"  Zeit.  phys.  Chem.t  10,  203,  1892. 

§  192.  Plane  Trigonometry. 

Beginners  in  the  calculus  often  trip  over  the  trigonometrical  work.  The 
following  outline  will  perhaps  be  of  some  assistance.  Trigonometry  deals 
with  the  relations  between  the  sides  and  angles  of  triangles.  If  the  triangle 
is  drawn  on  a  plane  surface,  we  have  plane  trigonometry ;  if  the  triangle  is 
drawn  on  the  surface  of  a  sphere,  spherical  trigonometry.  The  trigonometry 
employed  in  physics  and  chemistry  is  a  mode  of  reasoning  about  lines  and 
angles,  or  rather,  about  quantities  represented  by  lines  and  angles  (whether 
parts  of  a  triangle  or  not),  which  is  carried  on  by  means  of  certain  ratios  or 
functions  of  an  angle. 

1.  The  measurement  of  angles.  An  angle  is  formed  by  the  intersection  of 
two  lines.  The  magnitude  of  an  angle  depends  only  on  the  relative  directions, 
or  slopes  of  the  lines,  and  is  independent  of  their  lengths.  In  practical  work, 
angles  are  usually  measured  in  degrees,  minutes  and  seconds.  These  units 
are  the  subdivisions  of  a  right-angle  defined  as 

1  right  angle  =  90  degrees,  written  90° ; 
1  degree  =  60  minutes,  written  60' ; 

1  minute        =  60  seconds,  written  60". 

In  theoretical  calculations,  however,  this  system  is  replaced  by  another. 
In  Fig.  181,  the  length  of  the  circular  arcs  P'A',  PA,  drawn  from  the  centre  O, 
are  proportional  to  the  lengths  of  the  radii  OA'  and  OA,  or 
arc  P'A'  arc  PA 

radius  OA'  ~  radius  OA' 
If  the  angle  at  the  centre  O  is  constant,  the  ratio,  arc/radius,  is  also  constant. 
This  ratio,  therefore,  furnishes  a  method  for  measuring  the  magnitude  of  an 
angle.    The  ratio 

=  1,  is  called  a  radian. 


radius 
Two  right  angles  =  180°  =  tt  radians,  where  n  =  180°  =  3-14159.        (1) 


§  192.  APPENDIX  I.  607 

The  ratio,  arc/radius,  is  called  the  circular  or  radian  measure  of  an  angle. 
(Radian  =  unit  angle.) 

2.  Relation  between  degrees  and  radians.  The  circumference  of  a  circle 
of  radius  r,  is  2*r,  or,  if  the  radius  is  unity,  2ir.  The  angles  360°,  180°,  90°, . . . 
correspond  to  the  arcs  whose  lengths  are  respectively,  2ir,  v,  ^7r, . . .  If  the 
angle  AOP  (Fig.  181)  measures  D  degrees,  or  a  radians, 

a  D 

D°  :  360°  =  a:2ir.     .-.  D°  =  ^360 ;   or,  a  =  ^2ir.     .         .         (2) 

Examples. — (1)  How  many  degrees  are  contained  in  an  arc  of  unit 
length  ?    Here  o  =  1, 

...  0  -  J2?  m  57-295°  m  57°  17'  44-8".       ...        (3) 

(2)  How  many  radians  are  there  in  1°  ?     Ansr.  w/180 ;  or,  0-0175. 

(3)  How  many  radians  in  2£°  ?    Ansr.  0-044. 

A  table  of  the  numerical  relations  between  angles  expressed  in  degrees 
and  radians  is  given  on  pages  624  and  625. 

3.  Trigonometrical  ratios  of  an  angle  as  functions  of  the  sides  of  a  triangle. 
There  are  certain  functions  of  the  angles,  or  rather  of  the  arc  PA  (Fig.  181) 

P 


Fig.  182. 

called  trigonometrical  ratios.     From  P  drop  the  perpendicular  PM  on  to  OM 

(Fig.  182).     In  the  triangle  OPM, 

,  *  ^  .    MP  perpendicular    . 

(i.)  The  ratio  q^,  or,  ^^ ,  is  called  the  tangent  of  the  angle 

POM,  and  written,  tan  POM. 

It  is  necessary  to  show  that  the  magnitude  of  this  ratio  depends  only  on 
the  magnitude  of  the  angle  POM,  and  is  quite  independent  of  the  size  of  the 
triangle.  Drop  perpendiculars  PM  and  P'M'  from  P  and  P'  on  to  OA 
(Fig.  181).  The  two  triangles  POM  and  P'OM',  are  equiangular  and  similar, 
therefore,  as  on  page  603,  M'P'/OM'  =  MPjOM. 

(ii.)  The  ratio  ^p,  or,  perpeDdicular.  is  called  the  cotangent  of  the  angle 

POM,  and  written,  cot  POM.      Note  that  the  cotangent  of  an  angle  is  the 

reciprocal  of  its  tangent. 

MP       perpendicular 

(iii.)  The  ratio  7^  =  nc 1 ,  is  called  the  sine  of  the  angle  POM, 

v     '  UP         hypotenuse  °  ■ 

and  written,  sin  POM. 

OP        hypotenuse 
(iv.)  The  ratio  ^p  -  perpendiculai.>  is  called  the  cosecant  of  the  angle 

POM,  and  written,  cosec  POM.     The  cosecant  of  an  angle  is  the  reciprocal  of 
its  sine. 


608 


HIGHER  MATHEMATICS.  §  192. 

.  is  called  the  cosine  of  the  angle  POM. 

enuse  ° 


,  ■  i .    ■  '  OM 

(v.)  The  ratio  k^  =  £ r 

v    '  OP       hypoti 

and  written,  cos  POM. 

(vi.)  The  ratio  q^  =  -~r ,  is  called  the  secant  of  the  angle  POM, 

and  written,  sec  POM.     The  secant  of  an  angle  is  the  reciprocal  of  its  cosine. 

Example. — If  a  be  used  in  place  of  POM,  show  that 
111 


sin  x  = 


;cota;=ta^;cosa;: 


cosec  x  tan  x  sec  x 

The  squares  of  any  of  these  ratio?,  (sin  x)2,  (cot  x)2, . . .,  are  generally 

written  sin'-to,  cot2# . . . ;  (sin  x)  r  \ .  (cot  x) -1, . . .,  meaning  -a — -»  ^~~x,  •  •  •» 

cannot  be  written  in  the  forms  sin  -  lx,  cot  ~  'a;, . . .  because  this  latter  symbol 
has  a  different  meaning. 

4.  To  find  a  numerical  value  for  the  trigonometrical  ratios. 

6 
D 


Fig.  183. 


Fig.  184. 


(i.)  45°  or  Jir.     Draw  a  square  ABCD  (Fig.  183).     Join  AC.     The  angle 

BAG=h.sAi  a  right  angle  =  45°.     In  the  right-angled  triangle  BAG  (Euolid,  i., 

47), 

AC2=AB2  +  BC\ 

Since  AB  and  BC  are  the  sides  of  a  square,  .\  AB  =  BC,  hence, 
AC2=2AB2=2B02 ;  or,  AC=  sj2~.AB=  s/2.  BO. 


••<*«- g-&«""-  *Hr-«"  S-i- 


(4) 


Fig.  185.  Fig.  186. 

(ii.)  90°  or  £tt.      In  Fig.  184,  if  POM  is  a  right-angled  triangle,  a&  if 

approaches  O,  the  angle  .MOP  approaches  90°.  When  MP  coincides  with 
OB,  OP=MP,  and  OM =zero. 


§192. 


APPENDIX  I. 


609 


MP  OM  MP 

.-.  sm9O°=0p=l;  cos  90°  =  ^  =  0;   tan  90°  =  ^=oo.  (5) 

(iii.)  0°.     In  Fig.  185  as  the  angle  MOP  becomes  smaller,  OP  approaches 
OM,  and  at  the  limit  coincides  with  it.     Hence,  PM =0 ;  OM=OP. 


.-.  sin0°=^p=0 


cosO0 


CM 


MP 


ap  =  l;tan0°  =  ™r=0. 


(6) 


(iv.)  60°  or  £*-.     In  the  equilateral  triangle  (Fig.  186),  each  of  the  three 
angles  is  equal  to  60°.     Drop  the  perpendicular  OM  on  to  PQ.    Then 

2PM=PQ=PO. 
By  Euclid,  i.,  47, 

PO2 = MP2  +  MO2.    . :  4PM2 = MO*  +  PM2 ;  or  MO2 = 3PM2. 
.-.  MO=s/d.PM;  angle  MPO= 60°. 
MO     slW 

Using  the  preceding  results, 
MP    1 


.-.  sin  60° 
(v.)  30°or^7r. 
.-.  sin  30' 
The  following  table  summarizes  these  results : 


PM    1  MO       ,— 

cos  60°=-or»  =  o  ;  tan  60°=  ^v,=  s/3. 


PM 


(7) 


MO     Jz  MP      1 

OPS'  C0S  30°=OP  =  ^;  tan30°  =  Olf=^3-  (8) 


Table  XIV. — Numerical  Values  of  the  Trigonometrical  Katios. 


• 

Angle. 

0°  to  360°. 

30°. 

•45°. 

60°. 

90°. 

180°. 

270°. 

1 

1 

x/3 

sine 

0 

2 

a/3 

a/2 

1 

a 

i 

1 

0 

-1 

cosine 

1 

~~ 2 

55 

2 

0 

-1 

0 

tangent 

0 

1 

a7? 

i 

a/3" 

00 

0 

00 

To  these  might  be  added  sin  15°  =  (a/3-1)/2a/2  ;  sin  18°  =  i(>/5-l).  See 
also  page  608. 

It  must  be  clearly  understood  that  although  an  angle  is  always  measured 
in  the  degree-minute-second  system,  the  numerical  equivalent  in  radian  or 
circular  measure  is  employed  in  the  calculations,  unless  special  provision  has 
been  made  for  the  direct  introduction  of  degrees.  This  was  done  in  example 
(6),  page  544  (q.v.).  Suppose  that  we  have  occasion  to  employ  the  approxima- 
tion formula 

sin  (jc  +  0)  =  sin  x+d  cos  a?, 

of  §  189,  and  that  <r=35°  and  0=50".  The  Tables  of  Natural  Sines,  Cosines, 
Tangents,  and  their  reciprocals,  will  furnish  the  numerical  values  of  sin  35° 
and  cos  35°,  but  6  must  be  converted  into  radian  measure.     Hence  show  that 

.*"."*—!    0-00926  xtt 
sin  (se  +  0)=sin  35°+ jgg cos  35°. 

Hint.  50"«=(H)'  =  (Uxiro)0=0'009260-     Tne  numerical  values  of  sin  35°  and 

QQ 


610 


HIGHER  MATHEMATICS. 


of  cos  35°  to  four  decimal  places  are  respectively  0*5736  and  0*8192.     The 
value  of  sin  (x  +  6)  is,  therefore,  *5737. 

5.  Conventions  as  to  the  sign  of  the  trigonometrical  ratios.  This  subject 
has  been  treated  on  page  123.  In  the  following  table,  these  results  are 
summarized.  The  change  in  the  value  of  the  ratio  as  it  passes  through  the 
four  quadrants  is  also  given. 

Table  XV. — Signs  of  the  Trigonometrical  Ratios. 


If  the  Angle  is  in 
Quadrant. 

sin*. 

cos*. 

tan*. 

cot*. 

sec*. 

cosec  *. 

t    fsign 
'  \value 

+ 
Otol 

+ 

1  toO 

+ 

0  to  cd 

+ 

CD  tO  0 

+ 

1  to  co 

+ 

oo  to  1 

TT   /si§n 
ii'  \ value 

+ 
ltoO 

Otol 

co  to  0 

0  to  co 

oo  to  1 

+ 

1  to  00 

ttt    (sign 
ii1,  \  value 

Otol 

1  toO 

+ 

0  to  oo 

+ 

CD  to  0 

1  to  oo 

oo  to  1 

TV  /si§n 
iV'\value 

• 

ltoO 

+ 
Otol 

CD  to  0 

0  to  oo 

+ 

co  to  1 

1  to  00 

'.  M'P' 


6.  Trigonometrical  ratios  of  the  supplement  of  an  angle.  The  angle  is 
180°  -x,  or  v-x,  is  called  the  supplement 
of  the  angle  x.  In  Fig.  187,  let  M0P=x, 
produce  OM  to  M' .  Then  the  angle  MOP' 
is  the  supplement  of  x.  Make  the  angle 
P'OM'  =  MOP.  Let  0P'  =  0P.  Drop  per- 
pendiculars P'M'  and  PM  on  to  BA.  The 
triangles  0PM  and  OP'M  are  equal  in  all 
respects.  If  OM  is  positive,  OM'  is  negative 

MP;  &n&OM'=-OM. 
sin  (180  -  x)  =  sin(7r  -  x)  =  sin  POM'  =  sin  MVP  =  sin  a*, 
cos  (180°  -  x)  =  -  cos  x ;  tan  (180°  -  x)  =  -  tan  x. 
Examples. — (1)  Find  the  value  of  sin  120°. 

sin  120°  =  sin  (180° -60°)=  sin  60=  Vf. 
(2)  Evaluate  tan  120°.     Ansr.  -  sjd. 

7.  Trigonometrical  ratios  of  the  complement  of  an  angle.  The  angle 
90°  -  x,  or  £tt  -  x,  is  called  the  complement  of  x.  In  Fig.  188,  PN  and  PM  are 
perpendiculars  on  OB  and  on  OA  respectively.     Then  OM=NP,  ON=MP. 

NP    OM 

sin  (90°  -x)  =  sin  ( Jt  -  x)  =  sin  NOP  =  Qp  =  Qp  =  cos  x. 

cos  (90°  -  x)  =  sin  x  ;  tan  (90°  -  x)  =  cot  x ;  cot  (90°  -a;)=tanx. 


§192. 


APPENDIX  I. 


611 


8.  To  prove  that  sin  <c/cos  sc=tan  x. 
sin  x    MP  l  OM    MP 


OP 


cos  x    OP     OP     OP    OM 


~/inX  r\-MT~  /->Ti/f — tan  •"• 


MP 
OM 


9.  To   prove    that    sin2sc  +  cos2*  =  1.       In  Fig.   189,   by  Euclid,  i.,  47, 
OP2 = MP*  +  OM2.     Divide  through  by  OP2,  and 


*0F?    Ml?    OM2_/MP\2    (0M\ 
OP*~  0P2  +  OP*~\OP  )  +\op) 


sin2*  4- cos2*  =  1. 


10.  To  show  that  sin  (x  +  y)  =  sin  x .  cos  y  +  cos  x .  sin  ?/.      In  Fig.  189, 
PQ  is  perpendicular  to  OQ,  the  angle  EPQ  =  angle  2V0Q  (Euclid,  i.,  15  and  32) 


MPPH    QNPH  PQ    NQ    OQ 
.-.sin  (x  +  y)-0p-OP  +  0p-pQ-op+0Q-0p> 

=  sin  x .  cos  y  +  cos  x .  sin  y. 

11.  Summary  of  trigonometrical  formula  (for  reference).  The  above  defi- 
nitions lead  to  the  following  relations,  which  form  routine  exercises  in 
elementary  trigonometry.  Most  of  them  may  be  established  geometrically 
as  in  the  preceding  illustrations  : 

Note:  ?r=180o  ;  or,  3-14159  radians  ;  one  radian  =  57-2958°. 

sin  (^7r  +  cc)=cos  x  ;  cos  (%ir-x)  =  B'm  *;  ^ 

cosec  (£*■-*)  =  sec  x  ;  sec  (£7r-*)=cosec  x 

tan  ($*■  -  x)  =  cot  x ;  cot  ($7r  -  x)  =  tan  x. 

sin  (71--*)=  sin  *;  cos  (ir-x)=  -cos  x  ; 

tan  (?r  -  x)  —  -  tan  x ;  cot  (ir  -  x)  =  -  cot  x. 

sin  ($ir  +  *)  =  cos  a; ;  cos  (£*■ + x)  =  -  sin  x  ;  I 

tan  (£?r + jb)  =  -  cot  a?,  cot  (£n-  +  x)=  -  tan  *.  / 

sin  (*•  +  *)  =  -sin  *;  cos  (tt  +  *)  =  -cos  x;  \ 

tan  (7r  +  *)  =  tan  x;  cot  (7r  +  *)  =  cot  x.  J 

sin  ( -  *)  =  -  sin  a; ;  cos  ( -  #)  =  cos  x ;  tan  ( -  x)  =  -  tan  x. 

sin  &     tan  x                       sin  -  ]*     tan  -  **     , 
cos*=l;    — - —  = — =1. 


} 


****•«  X  -""-'  x  X 

When  n  is  any  negative  or  positive  integer  or  zero. 

sin  *=sin  {rnr+(-l)nx}.      .        . 

cos  *  =  cos  (2nir  +  x).  . 

tan  *  =  tan  (nir  +  x).      .        ... 
tan  *  =  sin  */cos  x ;  cot  a;  — cos  */sin  x. 

sin2*  +  cos2* =1. 


(9) 
(10) 

(11) 

(12) 
(13) 
(14) 


(15) 
(16) 
(17) 
(18) 
(19) 


QQ 


612 


HIGHER  MATHEMATICS 


sin  x=  \fl 
cosec  x 


cos'a; 
tan  x 


cos  x=  \/l-sin2a;. 
sec  x=  \/l  +  tan2a;. 
1 


sin  a;  = 


V 1  +  tan2a;  y  1  +  tan'-jc 

sin  (x  ±  y) =sin  x .  cos  2/  ±  cos  x .  sin  y. 

cos  (a;  +  2/)  =cos  x .  cos  2/  +  sin  x .  sin  y. 

sin  (a;  +  2/)  +  sin  (a;-2/)=2  sin  x.  cos  y. 

sin  (a?  +  2/)  -sin  (x-y)=2  cos  a?,  sin  y. 

cos  («+ y)  +  cos  (x  -  y) =2  cos  x .  cos  y. 

cos  (sc+2/)-cos  (<c-2/)=  -2  sin  x.  sin  y 

If  05=j/,  from  (23)  and  (24), 

sin  2a; =2  sin  a; .  cos  x. 

cos  2a; = cos2a;  -  sin2a;.  . 

=2cos2a;-l.      . 

=l-2sin2a;. 

sin  a; =2  sin  %x .  cos  \x. 

cos  a; =2  cos2£a;-l;  or,  1  +  cos  a; =2  cos2£a; 

=  1-2  sin2£a; ;  or,  1  -  cos  x  =  2  sin2£a;, 

sin  3a; = 3  sin  x  -  4  sin3a;. 

cos  3a; =4  cos3a;-3  cos  x. 

If  in  (25)  to  (28),  we  suppose  x  +  y  =  a;  x-y=P;  x- 

Now  put  x  for  o,  and  y  for  /3,  for  the  sake  of  uniformity, 

sin  x  +  sin  y =2  sin  $(a;  +  2/) .  cos  \{x-y). 

sin  a; -sin  j/=2  cos  l{x  +  y) .  sin  ^(aj  — 2/). 

cos  a;  +  cos  y= 2  cos  $(a;  +  2/)  •  cos  i^-^)* 

cos  a; -cos  2/=  -2  sin  l{x  +  y)  .&in%(x-y) 

'  By  division  of  the  proper  formulae  above, 

tan  x  +  tan  y 


tan  (x+y)  = 
tan  (a;  -y)  = 

tan  2a; = 


1  -  tan  x .  tan  2/* 

tan  a; -tan  y 
1  +  tan  a;. tan  y' 
2  tan  x 


1-tanV 

sin  (a;  +  y) 
tan  x  ±  tan  2/==     c  ^  nM  ',. 

—         a     cos  x .  cos  2/ 

sin  (x  ±  y) 
.sin  2/* 


cot  x  ±  cot  j/ =- 


Thus, 


cos  $a;  1 


V^ 


sin  $a;= 


^ 


2 


tan  %x 


=Vi^ 


§193. 

(20) 
(21) 

(22) 

(23) 
(24) 
(25) 
(26) 
(27) 
(28) 

(29) 
(30) 
(31) 
(32) 
(33) 
(34) 
(35) 
(36) 
(37) 

(38) 
(39) 
(40) 
(41) 

(42) 
(43) 
(44) 
(45) 
(46) 

(47) 


§  193.  Relations  among  the  Hyperbolic  Functions. 

cosa;=cosh  ia;=J('*4-6-lX) .        .  (1) 

sin  x=—  sinh  tx—-$(6<-x-e-ix) .  (2) 

cos  a;  +  isin  a;=cosh  taj+sinh  vx=e<-*.         ..'...•  (3) 

cos  a;-isin  a?=cosh  ia;=-sinh  ix=e-* (4) 

cosh  a;  =  cos  ix;  isinh  a?=sin  ix .        .  (5) 


§193. 


APPENDIX  I. 


613 


tanh  8= sinh  8/cosh  x ;  coth  a; = cosh  8/sinh  x ,  \ 

cosech  x  =  1/sinh  x ;  sech  8=l/cosh  x.  J     ' 

cosh  0=1;  sinh  0=0;  tanh  0=0 

cosh  ( ±  oo)  =  +  oo ;  sinh  ( +  oo)  =  +  oo ;  tanh  ( +  oo)  =  +  1 

sinh  x  tanh  x  cosh  x 

Ltx=0-    -     =1;  Ltx=Q — —  =1;  LtXr=o      %     =1« 

sinh  ( -8)  =  - sinh  x ;  cosh  ( - x)  =  cosh  x ;  tanh  ( - x) 


-  tanh  x 


} 


x 

X) 

sinh  (x  ±  y)  =  sinh  x .  cosh  y  +  cosh  x .  sinh  y. 
cosh  {x  ±y)  =  cosh  x .  cosh  y  +  sinh  x .  sinh  y. 

_    .  j        tanh  a;  +  tanh  t/ 

tanh  («  ±  y)  -  r±tanha.Btftnhy.  •        • 

cosh  (x  +  ty)  =  cosh  x .  cosh  iy  +  sinh  a; .  sinh  iy ; 

=  cosh  a; .  cos  y+ 1  sinh  x .  sin  t/. 
sinh  (x  +  iy)  =  sinh  x .  cosh  ty  +  cosh  x .  sinh  ^ ; 

=sinh  x  .  cos  y  +  tcosh  x.  sin  y. 
sinh  x  +  sinh  y  =  2  sinh  £(8  +  y)  •  cosh  %{x-y).  . 
sinh  a;  -  sinh  y  =  2  cosh  $(8  +  y) .  sinh  £(8  -  y).  . 
cosh  8  + cosh  y  —  1  cosh  J(a;+y) .  cosh  ${x-y).  . 
cosh  a;  -  cosh  y  =  2  sinh  £  (a;  +  y) .  sinh  %(x-y).  . 
sinh  2a; =2  sinh  x .  cosh  a; =2  tanh  8/(1  -  tanh28).     . 

cosh  28  =  cosh28  +  sinh28; 

=  l  +  2sinh28=2  cosh28-l;  . 
=  (l  +  tanh28)/(l-tanh28).     .... 
cosh  8  +  1  =  2  cosh2^8;  cosh  a; -1  =  2  sinh2£8.     • 
tanh  \x  =  sinh  8/(1  +  cosh  x) ;    ^ 

=  (cosh  x  -  l)/sinh  x.     )        '        * 

sinh28-cosh28=l 

l-tanh28=sech28;  coth28  - 1  =  cosech28. 

cosh  8=1/  */(l  -  tanh28) ;  sinh  8= tanh  x\  ^/(l  -  tanh28). 

sinh  38  =  3  sinh  8  +  4  sinhs8 

cosh  38=4  cosh38-3  cosh  8 

Inverse  hyperbolic  functions.     If  sinh  -ly  =  x,  then  y  =  sinh  x. 
y  =  sinh  «=£(«*-«-*);  .*.  e'ix-2yex-l=0;  .:  ex=y  ±  <s/ya  +  l. 

For  real  values  of  8  the  negative  sign  is  excluded  from  sinh  -  xx. 
.*.  sinh ~lx= log {y  +  s/y2  +  l};  cosh  -  1^=log{2/  ±  sfy'2-l}. 
tanh-ty-J  log  {(l+y)/(l-y)} ;  eoth-Hr-l  log  {(y  +  l)/(y-l)}. 
sech -12/= log  {(1  +  sll-y^jy) ;  cosech  -ty=log  {(1  -  Vl  +  »2)M 
Gudermannians.    In  Fig.  138,  page  347, 

0  =  cos-1sech  8;  cos0=sech8;  sec  0= cosh  8. 

$0  =  cos  0+sinh  8=sec  0  +  tan  0. 

0=log  (sec  0  tan  0)  =  log  tan  (i?r  +  £0) 

tanh  $8= tan  \d 

when  0  is  connected  with  x  by  any  of  these  relations  0  is  said  to 
gudermannian  of  x  and  is  written  gd  x. 
Analogous  to  Demoivre's  theorem 

(cosh  8  +  sinh  8)n  =  cosh  nx  +  sinh  nx.  . 


(6) 

(7) 
(8) 

(9) 

(10) 

(11) 
(12) 

(13) 
(14) 

(15) 

(16) 
(17) 
(18) 
(19) 
(20) 
(21) 
(22) 
(23) 
(24) 

(25) 

(26) 
(27) 
(28) 
(29) 
(30) 
Since 


(31) 
(32) 
(33) 

(34) 
(35) 

(36) 
(37) 

be  the 


(38) 


014 


HIGHEE  MATHEMATICS. 


§193. 


It  is  instructive  to  compare  the  above  formulae  with  the  corresponding 
trigonometrical  functions  in  §  192.  The  analogy  is  also  brought  out  by 
tabulating  corresponding  indefinite  integrals  in  Tables  I.  and  III.,  side  by 
side.  A  few  additional  integrals  are  here  given  to  be  verified  and  then  added 
to  the  table  of  indefinite  integrals  which  the  student  has  been  advised  to 
compile  for  his  own  use. 

Table  XVI. — Hyperbolic  and  Trigonometrical  Functions. 


Hyperbolic. 


Trigonometrical. 


dx  X 

„  ss  sinh  ~ 1- 


■  +  a*  » 

dx  ,       x 

i  „        ,.  =  cosh  ~  *-. 

six-  -a2  <*> 

dx  1  x 

> =  —  tanh    l-, 

a*  -  x1      a  a 

when  x  <  a.     When  x  >  a, 

^  =  icoth-£ 
a  a 


L 
I 
h 

Jsech  x .  dx  —  gd  x. 


-  dx  1       ,       x 

.  -  =  —  sech  - 1-« 

xsla2  -  x2      a  a 

dx  1  .      n05 

r   ' ;  =  ~z  cosech  ~  *-. 

x  Va2  +  x2      a  a 


dx 


sja2 


=  sin' 


dx 


a 

==  =  cos    l-. 

si  a2  -  x2  a 

dx         1  x 

— r,  =  -  tan  _  l~. 

a2  +  x*      a  a 


f   -dx   _  1 

J  a2  +  x2  ~  a 


cot 


dx  1  ,x 

.  -  =  —  sec  ~~ 1-. 

x  six2  -  a2      a  a 

-  dx  1  x 

,  n  =  -  cosec  _  l-. 

xs/x2  -  a2      «  a 


/; 

Jsec  x.dx  =  gd  _  xx. 


(39) 
(40) 
(41) 

(42) 
(43) 

(44) 
(45) 


Numerical  values  of  the  hyperbolic  functions  may  be  computed  by  means 
of  the  series  formulte. 


APPENDIX  II. 

REFERENCE  TABLES. 

"The  human  mind  is  seldom  satisfied,  and  is  certainly  never  exercising 
its  highest  functions,  when  it  is  doing  the  work  of  a  calculating 
machine." — J.  C.  Maxwell. 

The  results  of  old  arithmetical  operations  most  frequently  required  are 
registered  in  the  form  of  mathematical  tables.  The  use  of  such  tables  not 
only  prevents  the  wasting  of  time  and  energy  on  a  repetition  of  old  operations 
but  also  conduces  to  more  accurate  work,  since  there  is  less  liability  to  error 
once  accurate  tables  have  been  compiled.  Most  of  the  following  tables  have 
been  referred  to  in  different  parts  of  this  work,  and  are  reproduced  here  be- 
cause they  are  not  usually  found  in  the  smaller  current  sets  of  "  Mathemat- 
ical Tables".  Besides  those  here  you  ought  to  have  "  Tables  of  Reciprocals, 
Squares,  Cubes  and  Roots,"  "  Tables  of  Logarithms  of  Numbers  to  base  10," 
"  Tables  of  Trigonometrical  Sines,  Cosines  and  Tangents  "  for  natural  angles 
and  logarithms  of  the  same.     See  page  xix.  of  the  Introduction. 

Table  I.— Singular  Values  of  Functions. 

(Page  168.) 

Table  II.— Standard  Integrals. 

(Page  193.) 

Table  III.— Standard  Integrals  (Hyperbolic  functions.) 

(Pages  349  and  614.) 


615 


616 


HIGHEK  MATHEMATICS. 


Table  IY—  Numerical  Values  of  the 

Hyperbolic  Sines, 

Cosines,  e*   and  e~x. 

. x- 

ex. 

e-*. 

cosh  x. 

sinh  x. 

X. 

ex. 

6~x. 

cosh  x. 

sinh  x. 

o-oc 

1-0000C 

u-ooooc 

1-0000C 

•00000 

0-4C 

)     1-49182 

0-67032 

!     1-08107 

'     0-41075 

•01 

1-01005 

i    -99005 

1-00005 

•01000 

•41 

1-50682 

•66365 

l-0852c 

!        -42158 

•02 

1-02021 

•9802C 

1-0002C 

•02000 

•42 

,     1-52196 

■65705 

1-0895C 

)        -43246 

•OS 

1-03045 

>    -97045 

1-00045 

•03000 

•43 

1-53726 

•65051 

1-09388 

•44337 

•04 

1-04081 

•9608C 

1-0008C 

•04001 

•44 

1-55271 

•64404 

1-09837 

•45434 

•Ofi 

1-05127 

•95123 

1-00125 

•05002 

•45 

1-56831 

.63763 

1-10297 

•46534 

•oe 

1-06184 

•94177 

1-00180 

•06004 

•46 

1-58407 

•63126 

1 -10766 

•47640 

•07 

1-07251 

•93239 

1-00245 

•07006 

•47 

1-59999 

•6250C 

1-1125C 

•48750 

•OS 

1-0832C 

I    -92312 

1-00320 

•08009 

•46 

1-61607 

•61878 

1-11743 

•49865 

•08 

1-09417 

•91393 

1-00405 

•09012 

•4S 

1-63232 

•61263 

1-12247 

•50985 

•1C 

1-10517 

•90484 

1-00500 

•10017 

•50 

1-64872 

•60653 

1-12763 

•52110 

•11 

1-11626 

•89583 

1-00606 

•11022 

•51 

1-66529 

•60050 

1-13289 

•53240 

•12 

1-1275C 

•88692 

1-00721 

•12029 

•52 

1-68203 

•59452 

1-13827 

•54375 

•13 

1-13883 

•87810 

1-00846 

•13037 

•53 

1-69893 

•58860 

1-14377 

•55516 

•14 

1-15027 

•86936 

1-00982 

•14046 

•54 

1-71601 

•58275 

■  1-14938 

•56663 

•15 

1-16183 

•86071 

1-01127 

•15056 

•55 

1-73325 

•57695 

1-15510 

•57815 

•16 

1-17351 

•85214 

1-01283 

•16068 

•56 

1-75067 

•57121 

1-16094 

•58973 

•17 

1-18530 

•84366 

1-01448 

•17082 

•57 

1-76827 

•56553 

1-16690 

•60137 

•18 

1-19722 

•83527 

1-01624 

•18097 

•58 

1-78604 

•55990 

1-17297 

•61307 

•19 

1-20925 

•82696 

1-01810 

•19115 

•59 

1-80399 

•55433 

1-17916 

•62483 

•20 

1-22140 

•81873 

1-02007 

•20134 

•60 

1-82212 

.54881 

1-18547 

•63665 

•21 

1-23368 

•81058 

1-02213 

•21155 

•61 

1-84043 

•54335 

1-19189 

•64854 

•22 

1-24608 

•80252 

1-02430 

•22178 

•62 

1-85893 

•53794 

1-19844 

•66049 

•23 

1-25860 

•79453 

1-02657 

•23203 

•63 

1-87761 

•53259 

1-20510 

•67251 

•24 

1-27125 

•78663 

1-02894 

•24231 

•64 

1-89648 

•52729 

1-21189 

•68459 

•25 

1 -28403 

•77880 

1-03141 

•25261 

•65 

1-91554 

•52205 

1-21879 

•69675 

•26 

1-29693 

•77105 

1-03399 

•26294 

•66 

1-93479 

•51685 

1-22582 

•70897 

•27 

1-30996 

•76338 

1-03667 

•27329 

•67 

1-95424 

•51171 

1-23297 

•72126 

•28 

1-32313 

•75578 

1-03946 

•28367 

•68 

1-97388 

•50662 

1-24025 

•73363 

•29 

1-33643 

♦74826 

1-04235 

•29408 

•69 

1-99372 

•50158 

1-24765 

•74607 

•30 

1-34986 

•74082 

1-04534 

•30452 

•70 

2-01375 

•49659 

1-25517 

•75858 

•31 

1-36343 

•73345 

1-04844 

•31499 

•71 

2-03399 

•49164 

1-26282 

•77117 

•32 

1-37713 

•72615 

1-05164 

•32549 

•72 

2-05443 

•48675 

1-27059 

•78384 

•33 

1-39097 

•71892 

1-05495 

•33602 

•73 

2-07508 

•48191 

1-27850 

•79659 

•34 

1-40495 

•71177 

1-05836 

•34659 

•74 

2-09594 

•47711 

1-28652 

•80941 

•35 

1-41907 

•70469 

1-06188 

•35719 

•75 

2-11700 

•47237 

1-29468 

•82232 

•36 

1-43333 

•69768 

1-06550 

•36783 

•76 

2-13828 

•46767 

J.-30297 

•83530 

•37 

1-44773 

•69073 

1-06923 

•37850 

•77 

2-15977 

•46301 

1-31139 

•84838 

•38 

1-46228 

•68386 

1-07307 

•38921 

•78 

2-18147 

•45841 

1-31994 

•86153 

•39 

1-47698 

•67706 

1-07702 

•39996 

•79 

2-20340 

•45384 

1-32862 

•87478 

APPENDIX  II. 


617 


Table  IY. — Continued. 


X. 

ex. 

e~x. 

cosh  x. 

sinh  x. 

x. 

ex. 

•-*. 

cosh  x. 

sinh  x. 

•80 

2-22554 

•44932 

1-33743 

•88811 

1-2C 

)     3-32012 

•30118 

1-81066 

1.50946 

•81 

2-24791 

•44486 

1-34638 

•90152 

1-21 

3-3534S 

•2982C 

1-82584 

1-52764 

•82 

2-27050 

•44049 

1-35547 

•91503 

1-2S 

3-3871S 

•29523 

1-84121 

1-54598 

•83 

2-29332 

•43605 

1-36468 

•92863 

1-22 

3-42123 

•29229 

1-85676 

1-56447 

•84 

2-31637 

•43171 

1-37404 

•94233 

1-24 

3-45561 

•28938 

1-87250 

1-58311 

•85 

2-33965 

•42741 

1-38353 

•95612 

1-25 

3-49034 

•28650 

1-88842 

1-60192 

•86 

2-36316 

•42316 

1-39316 

•97000 

1-26 

3-52542 

•28365 

1-90454 

1-62088 

•87 

2-38691 

•41895 

1-40293 

•98398 

1-27 

3-56085 

•28083 

1-92084 

1-64001 

•88 

2-41090 

•41478 

1-41284 

•99806 

1-28 

3-59664 

•27804 

1-93734 

1-65930 

•89 

2-43513 

•41066 

1-42289 

1-01224 

1-29 

3-63279 

•27527 

1-95403 

1-67876 

•90 

2-45960 

•40657 

1-43309 

1-02652 

1-30 

.3-66930 

•27253 

1-97091 

1-69838 

•91 

2-48432 

•40252 

1-44342 

1-04090 

1-31 

3-70617 

•26982 

1-98800 

1-71818 

•92 

2-50929 

•39852 

1-45390 

1-05539 

1-32 

3-74342 

•26714 

2-00528 

1-73814 

•93 

2-53451 

•39455 

1-46453 

1-06998 

1-33 

3-78104 

•26448 

2-02276 

1-75828 

•94 

2-55998 

•39063 

1-47530 

1-08468 

1-34 

3-81904 

•26185 

2-04044 

1-77860 

•95 

2-58571 

•38674 

1-48623 

1-09948 

1-35 

3-85743 

•25924 

2-05833 

1-79909 

•96 

2-61170 

•38289 

1-49729 

1-11440 

1-36 

3-89619 

•25666 

2-07643 

1-81977 

•97 

2-63794 

•37908 

1-50851 

1-12943 

1-37 

3-93535 

•25411 

2-09473 

1-84062 

•98 

2-66446 

•37531 

1-5198? 

1-14457 

1-38 

3-97490 

•25158 

2-1132! 

1-86166 

•99 

2-69123 

•37158 

1-53141 

1-15983 

1-39 

4-01485 

•24908 

2-13196 

1-88289 

1-00 

2-71828 

•36788 

1-54308 

1-17520 

1-40 

4-05520 

•24660 

2-15090 

1-90430 

1-01 

2-74560 

•36422 

1-55491 

1-19069 

1-41 

4-09596 

•24414 

2-17005 

1-92591 

1-02 

2-77319 

•36059 

1-56689 

1-20630 

1-42 

4-13712 

•24171 

2-18942 

1-94770 

1-03 

2-80107 

•35701 

1-57904 

1-22203 

1-43 

4-17870 

•23931 

2-20900 

1-96970 

1-04 

2-82922 

•35345 

1-59134 

1-23788 

1-44 

4'22070 

•23693 

2-22881 

1-99188 

1-05 

2-85765 

•34994 

1-60379 

1-25386 

1-45 

4-26311 

•23457 

2-24884 

2-01428 

1-06 

2-88637 

•34646 

1-61641 

1-26996 

1-46 

4-30596 

•23224 

2-26910 

2-03686 

1-07 

2-91538 

•34301 

1-62919 

1-28619 

1-47 

'  4-34924 

•22993 

2-28958 

2-05965 

1-08 

2-94468 

•33960 

1-64214 

1-30254 

1-48 

4-39295 

•22764 

2-31029 

2-08265 

1-09 

2-97427 

•33622 

1-65525 

1-31903 

1-49 

4-43710 

•22537 

2-33123 

2-10586 

1-10 

3-00417 

•33287 

1-66852 

1-33565 

1-50 

4-48169 

•22313 

2-35241 

2-12928 

1-11 

303436 

•32956 

1-68196 

1-35240 

1-51 

4-52673 

•22091 

2-37382 

2-15291 

1-12 

3-06485 

•32628 

1-69557 

1-36929 

1-52 

4-57223 

•21871 

2-39547 

2-17676 

1-13 

3-09566 

•32303 

1-70934 

1-38631 

1-53 

4-61818 

•21654 

2-41736 

2-20082 

1-14 

3-12677 

•31981 

1-72329 

1-40347 

1-54 

4-66459 

•21438 

2-43949 

2-22510 

1-15 

3-15819 

•31664 

1-73741 

1-42078 

1-55 

4-71147 

•21225 

2-46186 

2-24961 

1-16 

3-18993 

•31349 

1-75171 

1-43822 

1-56 

4-75882 

•21014 

2-48448 

2-27434 

1-17 

3-22199 

•31037 

1-76618 

1-45581 

1-57 

4-80665 

•20805 

2-50735 

2-29930 

1-18 

3-25437 

•30728 

1-78083 

1-47355 

1-58 

4-85496 

•20598 

2-53047 

2-32449 

1-19 

3-28708 

•30422 

1-79565 

1-49143 

1-59 

4-90375 

•20393 

2-55384 

2-34991 

618 


HIGHER  MATHEMATICS. 


Table  IV. — Continued.   . 


ST. 

ex. 

6-*. 

cosh  x. 

sinh  x. 

X. 

ex. 

6~x 

cosh  x. 

sinh  x. 

1-60 

4-95303 

•20190 

2-57746 

2-37557 

2-0 

7-38906 

•13534 

3-76220 

3-62686 

1-61 

5-00281 

•19989 

2-60135 

2-40146 

2-1 

8-16617 

•12246 

4-14431 

4-02186 

1-62 

5-05309 

•19790 

2-62549 

2-42760 

2-2 

9-02501 

•11080 

4-56791 

4-45711 

1-63 

5-10387 

•19593 

3-64990 

2-45397 

2-3 

9-97418 

•10026 

5-03722 

4-93696 

1-64 

5-15517 

•19398 

2-67457 

2-48059 

2-4 

11-0232 

•09072 

5-55695 

5-46623 

1-65 

5-20698 

•19205 

2-69951 

2-50746 

2-5 

12-1825 

•08208 

6-13229 

6-05020 

1-66 

5-25931 

•19014 

2-72472 

2-53459 

2-6 

13-4637 

•07427 

6-76900 

6-69473 

1-67 

5-31217 

•18825 

2-75021 

2-56196 

2-7 

14-8797 

•06721 

7-47347 

7-40626 

1-68 

5-36556 

•18637 

2-77596 

2-58959 

2-8 

16-4416 

•06081 

8-25273 

8-19192 

1-69 

5-41948 

•18452 

2-80200 

2-61748 

2-9 

18-1741 

•05502 

9-11458 

9-05956 

1-70 

5-47395 

•18268 

2-82832 

2-64563 

3-0 

20-0855 

•04979 

10-0677 

10-0179 

1-71 

5-52896 

•18087 

2-85491 

2-67405 

3-1 

22-1980 

•04505 

11-1215 

11-0765 

1-72 

5-58453 

•17907 

2-88180 

2-70273 

3-2 

24-5325 

•04076 

12-2866 . 

12-2459 

1-73 

5-64065 

•17728 

2:90897 

2-73168 

3-3 

27-1126 

•03688 

13-5747 

13-5379 

1-74 

5-69743 

•17552 

2-93643 

2-76091 

3-4 

29-9641 

•03337 

14-9987. 

14-9654 

1-75 

5-75460 

•17377 

2-96419 

2-79041 

3-5 

33-1155 

•03020 

16-5728 

16-5426 

1-76 

5-81244 

•17204 

2-99224 

2-82020 

3-6 

36-5982 

•02732 

18-3128 

18-2855 

1-77 

5-87085 

•17033 

3-02059 

2-85026 

3-7 

40-4473 

•02472 

20-2360 

20-2113 

1-78 

5-92986 

•16864 

3-04925 

2-88061 

3-8 

44-7012 

•02237 

22-3618 

22-3394 

1-79 

5-98945 

•16696 

3-07821 

2-91125 

3-9 

49-4024 

•02024 

24-7113 

24-6911 

1-80 

6-04965 

•16530 

3-10747 

2-94217 

4-0 

54-5982 

•01832 

27-3082 

27-2899 

1-81 

6-11045 

•16365 

3-13705 

2-97340 

4-1 

60-3403 

•01657 

30-1784 

30-1619 

1-82 

6-17186 

•16203 

3-16694 

3-00492 

4-2 

66-6863 

•01500 

33-3507 

33-3357 

1-83 

6-23389 

•16041 

3-19715 

3-03674 

4-3 

73-6998 

•01357 

36-8567 

36-8431 

1-84 

6-29654 

•15882 

3-22768 

3-06886 

4-4 

81-4509 

•01228 

40-7316 

40-7193 

1-85 

6-35982 

•15724 

3-25853 

3-10129 

4-5 

90-0171 

•01111 

45-0141 

45-0030 

1-86 

6-42374 

•15567 

3-28970 

3-13403 

4-6 

99-4843 

•01005 

49-7472 

49-7371 

1-87 

6-48830 

•15412 

3-32121 

3-16709 

4-7 

109-947 

•00910 

54-9781 

54-9690 

1-88 

6-55350 

•15259 

3-35305 

3-20046 

4-8 

121-510 

•00823 

60-7593 

60-7511 

1-89 

6-61937 

•15107 

3-38522 

3-23415 

4-9 

134-290 

•00745 

67-1486 

67-1412 

1-90 

6-68589 

•14957 

3-41773 

3-26816 

5-0 

148-413 

•00674 

74-2099 

74-2032 

1-91 

6-75309 

•14808 

3-45058 

3-30250 

5-1 

164-022 

•00610 

82-0140 

82-0079 

1-92 

6-82096 

•14661 

3-48378 

3-33718 

5-2 

181-272 

•00552 

90-6388 

90-6333 

1-93 

6-88951 

•14515 

3-51733 

3-37218 

5-3 

200-337 

•00499 

100-171 

100-167 

1-94 

6-95875 

•14370 

3-55123 

3-40752 

5-4 

221-406 

•00452 

110-705 

110-701 

1-95 

7-02869 

•14227 

3-58548 

3-44321 

5-5 

244-692 

•00409 

122-348 

122-344 

1-96 

7-09933 

•14086 

3-62009 

3-47923 

5-6 

270-426 

•00370 

135-215 

135-211 

1-97 

7-17068  -13946 

3-65507 

3-51561 

5-7 

298-867 

•00335 

149-435 

149-432 

1-98 

7-24274 

•13807 

3-69041 

3-55234 

5-8 

330-300 

•00303 

165-151 

165-148 

1-99 

7-31553 

•13670 

3-72611 

3-58942 

5-9 

365-037 

•00274 

182-520 

182-517 

6-0 

403-429 

•0024S 

201-716 

201-317 

APPENDIX  II. 


619 


Table  Y. — Common  Logarithms  of  the  Gamma  Function. 

(Page  426.) 

Table  YL— Numerical  Yalues  of  the  Factor 


0-6745  y --—-«     (Page  523.) 


.  06745 

It. 

s/n  ~  l 

0. 

l. 

2. 

3. 

4. 

5. 

6. 

7. 

8. 

9. 

0 

0*6745 

0-4769 

0-3894 

0-3372 

0-3016 

0-2754 

0-2549 

0*2385 

1 

0-2248 

0-2133 

•2029 

•1947 

•1871 

•1803 

•1742 

•1686 

•1636' 

•1590 

2 

•1547 

•1508 

•1472 

•1438 

•1406 

•1377 

•1349 

•1323 

•1298 

•1275 

S 

•1252 

•1231 

•1211 

•1192 

•1174 

•1157 

•1140 

•1124 

•1109 

•1094 

4 

•1080 

•1066 

•1053 

•1041 

•1029 

•1017 

•1005 

•0994 

•0984 

•0974 

5 

0-0964 

0-0954 

0-0944 

0-0935 

0-0926 

0-0918 

0-0909 

0-0901 

0-0893 

0-0886 

6 

•0878 

•0871 

•0864 

•0857 

•0850 

•0843 

•0837 

•0830 

•0824 

•0818 

7 

•0812 

•0806 

•0800 

•0795 

•0789 

•0784 

•0778 

•0773 

•0768 

•0763 

6 

•0759 

•0754 

•0749 

•0745 

•0740 

•0736 

•0731 

•0727 

•0723 

•0719 

9 

•0715 

•0711 

•0707 

•0703 

•0699 

•0696 

•0692 

•0688 

•0685 

•0681 

Table  YII. — Numerical  Yalues  of  the  Factor 

0-6745  _ 

-7===.      Page  524. 
\/w(n  -  1) 


n. 
0 

0-6745 

sjn{n  -  1)* 

0. 

1. 

2. 

3. 

4. 

5. 

ft 

7. 

8. 

9. 

0-4769 

0-2754 

0-1947 

0-1508 

0-1231 

0-1041 

0-0901 

0-0795 

1 

0-0711 

0-0643 

•0587 

•0540 

•0500 

•0465 

•0435 

•0409 

•0386 

•0365 

2 

•0346 

•0329 

•0314 

•0300 

•0287 

•0275 

••0265 

•0255 

•0245 

•0237 

a 

•0229 

•0221 

•0214 

•0208 

•020i 

•0196 

•0190 

•0185 

•0180 

•0175 

•1 

•0171 

•0167 

•0163 

•0159 

•0155 

•0152 

•0148 

•0145 

•0142 

•0139 

6 

0-0136 

0-0134 

0-0131 

0-0128 

0-0126 

0-0124 

0-0122 

0-0119 

0-0117 

0-0115 

6 

•0113 

•0111 

•0110 

•0108 

•0106 

•0105 

•0103 

•0101 

•0100 

•0098 

7 

•0097 

•0096 

•0094 

•0093 

•0092 

•0091 

•0OS9 

•0088 

•0087 

•0086 

H 

•0085 

•0084 

•0083 

•0082 

•0091 

•0080 

•0080 

•0079 

•0077 

•0076 

9 

•0075 

•0074 

•0073 

•0073 

•0072 

•0071 

•0070 

•0069 

•0069 

•0068 

620 


HIGHER  MATHEMATICS. 


Table  YIIL— Numerical  Values  of  the  Factor 


°*8453V^i)-  (p^e624-) 


n. 

08453 

sjn(n  -  1)' 

0. 

1. 

2. 

3. 

4. 

5. 

6. 

7. 

8. 

9. 

0 

0-5978 

0-3451 

0-2440 

0-1890 

0-1543 

0-1304 

0-1130 

0-0996 

1 

0-0891 

0-0806 

•0736 

•0677 

•0627 

•0583 

•0546 

•0513 

•0483 

•0457 

2 

•0434 

•0412 

•0393 

•0376 

•0360 

•0345 

•0331 

•0319 

•0307 

•0297 

3 

•0287 

•0277 

•0268 

•0260 

•0252 

•0245 

•0238 

•0232 

•0225 

•0220 

4 

•021,4 

•0209 

•0204 

•0199 

•0194 

•0190 

•0186 

•0182 

•0178 

•0174 

5 

0-0171 

0-0167 

0-0164 

0-0161 

0-0158 

0-0155 

0-0152 

0-0150 

0-0147 

0-0145 

6 

•0142 

•0140 

•0137 

•0135 

•0133 

•0132 

•0129 

•0127 

•0125 

•0123 

7 

•0122 

•0120 

•0118 

•0117 

•0115 

•0113 

•0112 

•0110 

•0109 

•0108 

8 

•0106 

•0105 

•0104 

•0102 

•0101 

•0100 

•0099 

•0098 

•0097 

•0096 

1) 

•0095 

•0093 

•0092 

•0091 

•0090 

•0089 

•0089 

•0088 

•0087 

•0086 

Table  IX. — Numerical  Values  of  the  Factor 

l 


0-8453 


»Vw  -  1' 


(Page  524.) 


0-8453 

n. 

nsjn  -  l" 

0. 

1. 

2. 

3. 

4. 

5 

6. 

7. 

8. 

9. 

0 

0-4227 

0*1993 

'0-1220 

0-0845 

0-0630 

0-0493 

0-0399 

0-0332 

1 

0-0282 

0-0243 

•0212 

•0188 

•0167 

•0151 

•0136 

•0124 

•0114 

•0105 

2 

•0097 

•0090 

•0084 

•0078 

•0073 

•0069 

•0065 

•0061 

•0058 

•0055 

3 

•0052 

•0050 

•0047 

•0045 

•0043 

•0041 

•0040 

•0038 

•0037 

•0035 

4 

•0034 

•0033 

•0031 

•0030 

•0029 

•0028 

•0027 

•0027 

•0026 

•0025 

5 

0-0024 

0-0023 

0-0023 

0-0022 

0-0022 

0-0021 

0-0020 

0-0020 

0-0019 

0-0019 

6 

•0018 

•0018 

•0017 

•0017 

•0017 

•0016 

•0016 

•0016 

•0015 

•0015 

7 

•0015 

•0014 

•0014 

•0014 

•0013 

•0013 

•0013 

•0012 

•0012 

•0012 

8 

•0012 

•0012 

•0011 

•0011 

•0011 

•0011 

•0011 

•0010 

•0010 

•0010 

9 

•0010 

•0010 

•0010 

•0009 

•0009 

•0003 

•0009 

•0009 

•0009 

•0009 

APPENDIX  II. 


621 


Table  X. — Numerical  Values  of  the  Probability  Integral 

2    [hx 
P  =  -7=      e~  **d(hx) ,  (Page  532) , 

V7TJ   0 

where  P  represents  the  probability  that  an  error  of  observation  will  have  a 
positive  or  negative  value  equal  to  or  less  than  x,  h  is  the  measure  of  precision. 


hx. 

P. 

0. 

l. 

2. 

3. 

4. 

5. 

6. 

7. 

8. 

9. 

o-o 

o-oooo 

0-0113 

0-0226 

0-0338 

0-0451 

0-0564 

0-0676 

0-0789 

0-0901 

0-1013 

0-1 

•1125 

•1236 

•1348 

•1459 

•1569 

•1680 

•1790 

•1900 

•2009 

•2118 

0-2 

•2227 

•2335 

•2443 

•2550 

•2657 

•2763 

•2869 

•2974 

•3079 

•3183 

0-3 

•3286 

•3389 

•3491 

•3593 

•3694 

•3794 

•3893 

•3992 

•4090 

•4187 

0-4 

•4284 

•4380 

•4475 

•4569 

•4662 

•4755 

•4847 

•4937 

•5027 

•5117 

0-5 

0-5205 

0-5292 

0-5379 

0-5465 

0-5549 

0-5633 

0-5716 

0-5798 

0-5879 

0-5959 

0-6 

•6039 

•6117 

•6194 

•6270 

•6346 

•6420 

•6494 

•6566 

•6638 

•6708 

0-7 

•6778 

•6847 

•6914 

•6981 

•7047 

•7112 

•7175 

•7238 

•7300 

•7361 

0-8 

•7421 

•7480 

•7538 

•7595 

•7651 

•7707 

•7761 

•7814 

•7867 

•7918 

0-9 

•7969 

•8019 

•8068 

•8116 

•8163 

•8209 

•8254 

•8299 

•8342 

•8385 

1-0 

0-8427 

0-8468 

0-8508 

0-8548 

0-8586 

0-8624 

0-8661 

0-8698 

0-8733 

0-8768 

1-1 

•8802 

•8835 

•8868 

•8900 

•8931 

•8961 

•8991 

•9020 

•9048 

•9076 

1-2 

•9103 

•9130 

•9155 

•9181 

•9205 

•9229 

•9252 

•9275 

•9297 

•9319 

1-3 

•9340 

•9361 

•9381 

•9400 

•9419 

•9438 

•9456 

•9473 

•9490 

•9507 

1-4 

•9523 

•9539 

•9554 

•9569 

•9583 

•9597 

•9611 

•9624 

•9637 

•9649 

1-5 

0-9661 

0-9673 

0-9684 

0-9695 

0-9706 

0-9716 

0-9726 

0-9736 

0-9745 

0-9755 

1-6 

•9763 

•9772 

•9780 

•9788 

•9796 

•9804 

•9811 

•9818 

•9825 

•9832 

1-7 

•9838 

•9844 

•9850 

•9856 

•9861 

•9867 

•9872 

•9877 

•9882 

•9886 

1-8 

•9891 

•9895 

•9899 

•9903 

•9907 

•9911 

•9915 

•9918 

•9922 

•9925 

1-9 

•9928 

•9931 

•9934 

•9937 

•9939 

•9942 

•9944 

•9947 

•9949 

•9951 

2-0 

0-9953 

0-9955 

0-9957 

0-9959 

0-9961 

0-9963 

0-9964 

0-9966 

0-9967 

0-9969 

21 

•9970 

•9972j 

•9973 

•9974 

•9975 

•9976 

•9977 

•9979 

•9980 

•99S0 

2-2 

•9981 

•9982 

•9983 

•9984 

•9985 

•9985 

•9986 

•9987 

•9987 

•9988 

2-3 

•9989 

•9989 

•9990 

•9990 

•9991 

•9991 

•9992 

•9992 

•9992 

•9993 

2-4 

•9993 

•9993 

•9994, 

•9994 

•9994 

•9995 

•9995 

•9995 

•9995 

•9996 

2-5 

0-9996 

0-9996 

0-9996 

0-9997 

0-9997 

0-9997 

0-9997 

0-9997 

0-9998 

0-9998 

2-6 

•9998 

•9998 

•9998 

•9998 

•9998 

•9998 

•9998 

•9998 

•9998 

•9999 

00 

1-0000 

622  HIGHER  MATHEMATICS. 

Table  XI. — Numerical  Yalues  of  the  Probability  Integral 


=^lpH£>'<page532)' 


where  P  represents  the  probability  that  an  error  of  observation  -will  have  a 
positive  or  negative  value  equal  to  or  less  than  x,  r  denotes  the  probable  error. 


r 

p. 

te 

1. 

2. 

3. 

4. 

5. 

6. 

7. 

8. 

9. 

o-o 

o- 

0-0054 

0-0108 

0-0161 

0-0215: 

0-0269 

0-0323 

0-0377 

0-0430 

i 
0-0484 

o-i 

•0538 

•0591 

•0645 

•0699 

'  -0752 

•0806 

•0859 

•0913 

•0966 

•1020 

0-2 

•1073 

•1126 

•1180 

•1233 

•1286 

•1339 

•1392 

•1445 

•1498 

•1551 

0-3 

•1603 

•1656 

•1709 

•1761 

•1814 

•1866 

•1918 

•1971 

•2023 

•2075 

0-4 

•2127 

•2179 

•2230 

•2282 

•2334 

•2385 

•2436 

•2488 

•2539 

•2590 

0-5 

0-2641 

0-2691 

0-2742 

0-2793 

0-2843 

0-2893 

0-2944 

0-2994 

0-3043 

0-3093 

0-6 

•3143 

•3192 

•3242 

•3291 

•3340 

•3389 

•3438 

•3487 

•3535 

•3583 

0-7 

•3632 

•3680 

•3728 

•3775 

•3823 

•3870 

•3918 

•3965 

•4012 

•4059 

0-8 

•4105 

•4152 

•4198 

•4244 

•4290 

•4336 

•4381 

•4427 

•4472 

•4517 

0-9 

•4562 

•4606 

•4651 

•4695 

•4739 

•4783 

•4827 

•4860 

•4914 

•4957 

1-0 

0-5000 

0-5043 

0-5085 

0-5128 

0-5170 

0-5212 

0-5254 

0-5295 

0-5337 

0-5378 

1-1 

•5419 

•5460 

•5500 

•5540 

•5581 

•5620 

•5660 

•5700 

•5739 

•5778 

1-2 

•5817 

•5856 

•5894 

•5932 

•5970 

•6008 

•6046 

•6083 

•6120 

•6157 

1-3 

•6194 

•6231 

•6267 

•6303 

•6339 

•6375 

•6410 

•6445 

•6480 

•6515 

1-4 

•6550 

•6584 

•6618 

•6652 

•6686 

•6719 

•6753 

•6786 

•6818 

•6851 

1-5 

0-6883 

0-6915 

0-6947 

0-6979 

0-7011 

0-7042 

0-7073 

0-7104 

0-7134 

0-7165 

1-6 

•7195 

•7225 

•7255 

•7284 

•7313 

•7342 

•7371 

•7400 

•7428 

•7457 

1-7 

•7485 

•7512 

•7540 

•7567 

•7594 

•7621 

•7648 

•7675 

•7701 

•7727 

1-8 

•7753 

•7778 

•7804 

•7829 

•7854 

•7879 

•7904 

•7928 

•7952 

•7976 

1-9 

•8000 

•8023 

•8047 

•8070 

•8093 

•8116 

•8138 

•8161 

•8183 

•8205 

2-0 

0-8227 

0-8248 

0-8270 

0-8291 

0-8312 

0-8332 

0-8353 

0-8373 

0-8394 

0-8414 

2-1 

•8433 

•8453 

•8473 

•8492 

•8511 

•8530 

•8549 

•8567 

•8585 

•8604 

2-2 

•8622 

•8639 

•8657 

•8674 

•8692 

•8709 

•8726 

•8742 

•8759 

•8775 

2-3 

•8792 

•8808 

•8824 

•8840 

•8855 

•8870 

•8886 

•8901 

•8916 

•8930 

2-4 

•8945 

•8960 

•8974 

•8988 

•9002 

•9016 

•9029 

•9043 

•9056 

•9069 

2-5 

0-9082 

0-9095 

0-9108 

0-9121 

0-9133 

0-9146 

0-9158 

0-9170 

0-9182 

0-9193 

2-6 

•9205 

•9217 

•9228 

•9239 

•9250 

•9261 

•9272 

•9283 

•9293 

•9304 

2-7 

•9314 

•9324 

•9334 

•9344 

•9354 

•9364 

•9373 

•9383 

•9392 

•9401 

2-8 

•9410 

•9419 

•9428 

•9437 

•9446 

•9454 

•9463 

•9471 

•9479 

•9487 

2-9 

•9495 

•9503 

•9511 

•9519 

•9526 

•9534 

•9541 

•9548 

•9556 

•9563 

3-0 

0-9570 

0-957? 

0-9583 

0-9590 

0-9597 

0-9603 

0-9610 

0-9616 

0-9622 

0-9629 

3-1 

•9635 

•9641 

•9647 

•9652 

•9658 

•9664 

•9669 

•9675 

•9680 

•9686 

3-2 

•9691 

•9696 

•9701 

•9706 

•9711 

•9716 

•9721 

•9726 

•9731 

•9735 

3-3 

•9740 

•9744 

•9749 

•9753 

•9757 

•9761 

•9766 

•9770 

•9774 

•9778 

3-4 

•9782 

•9786 

•9789 

•9793 

•9797 

•9800 

•9804 

•9807 

•9811 

•9814 

3- 

0-9570 

0-9635 

0-9691 

0-9740 

0-9782 

0-9818 

0-9848 

0-9874 

0-9896 

0-9915 

4- 

•9930 

•9943 

•9954 

•9963 

•9970 

•9976 

•9981 

•9985 

•9988 

•9990 

5- 

•9993 

•9994 

•9996 

•9997 

•9997 

•9998 

•9998- 

•9999 

•9999 

•9999 

00 

1-0000 

APPENDIX  II. 


623 


Table  XII. — Numerical  Yalues  of  -  Corresponding  to  Different 
Values  of  w,  in  the  Application  of  ChauYenet's  Criterion. 

(Page  564.) 


n. 

0. 

1. 

2. 

3. 

4. 

5. 

(5. 

7. 

8. 

9. 

0 

2-05 

2-27 

2-44 

2-57 

2-67 

2-76 

2-84 

1 

2-91 

2-96 

3  02 

3-07 

3-12 

3-16 

3-19 

3-22 

3-26 

3-29 

2 

3-32 

3-35 

3-38 

3-41 

3-43 

3  "45 

3-47 

3-49 

3-51 

3-33 

3 

3-55 

3-57 

3-58 

3-60 

3-62 

3-64 

3-65 

3-67 

3-68 

3-69 

4 

3-71 

3-72 

3-73 

3-74 

3-75 

3-77 

3-78 

3-79 

3-80 

3-81 

5 

3-82 

3-83 

3-84 

3-85 

3-86 

3-87 

3-88 

3-88 

3-89 

3-90 

6 

3-91 

3-92 

3-93 

3-94 

3-95 

3-95 

3-96 

3-97 

3-97 

3-98 

7 

3-99 

3-99 

4-00 

4-01 

4-02 

4  02 

4-03 

4-04 

4-05 

4-05 

8 

4-06 

4-06 

4-06 

4-07 

4-07 

4-08 

4-09 

4-09 

4-10 

4-11 

9 

4-11 

4-12 

4-13 

4-14 

4-14 

4-15 

4-15 

4-15 

4-16 

4-16 

If  tt  =  100,  tf  =  416;  n  =  200,  *  =  4-48:  n  =  500,  t  =  4-90. 


624 


HIGHER  MATHEMATICS. 


Table  XIII. — Circular  or  Radian  Measure  of  Angles. 

(Page  606.) 


10 
11 
12 
13 
14 

15 
16 

17 
18 
19 

20 
21 
22 
23 
24 

25 
26 
27 
28 
29 

30 
31 
32 
33 
34 

35 
36 
37 
38 


40 
41 

42 
43 
44 


0'. 


12', 


18'. 


30'. 


36'. 


42'. 


48'. 


00000 
01745 
03491 
05236 
06981 

08727 
10472 
12217 
13963 
15708 

17453 
19199 
20944 
22689 
24435 

26180 
27925 
29671 
31416 
33161 

34907 
36652 
38397 
40143 

41888 

43633 
45379 
47124 
48869 
50615 

52360 
54105 
55851 
57596 
59341 

61087 
62832 
64577 
66323 
68068 

69813 
71558 
73304 
75049 
76794 


00175 
01920 
03665 
05411 
07156 

08901 
10647 
12392 
14137 
15882 

17628 
19373 
21118 
22864 
24609 

26354 
28100 
29845 
31590 
33336 

35081 
36826 
38572 
40317 
42062 

43808 
45553 
47298 
49044 
50789 

52534 
54280 
56025 
57770 
59516 

61261 
63006 
64752 
66497 
68242 

69988 
71733 
73478 
75224 
76969 


00349 
02094 
03840 
05585 
07330 

09076 
10821 
12566 
14312 
16057 

17802 
19548 
21293 
23038 

24784 

26529 
28274 
30020 
31765 
33510 

35256 
37001 

38746 
40492 
42237 

43982 
45728 
47473 
49218 
50964 

52709 
54454 
56200 
57945 
59690 

61436 
63181 
64926 
66672 
68417 

70162 
71908 
73653 
75398 
77144 


00524 
02269 
04014 
05760 
07505 

09250 
10996 
12741 
14486 
16232 

17977 
19722 
21468 
23213 
24958 

26704 
28449 
30194 
31940 
33685 

35430 
37176 
38921 
40666 
42412 

44157 
45902 
47647 
49393 
51138 

52883 
54629 
56374 
58119 
59865 

61610 
63355 
65101 
66846 
68591 

70387 

72082 

7382 

75573 

77318 


00698 
02443 
04189 
05934 
07679 

09425 
11170 
12915 
14661 
16406 

18151 
19897 
21642 
23387 
25133 

26878 
28623 
30369 
32114 
33859 

35605 
37350 
39095 
40841 
42586 

44331 
46077 
47822 
49567 
51313 

53058 
54803 
56549 
58294 
60039 

61785 
63530 
65275 
67021 
68766 

70511 

72257 
74002 
75747 
77493 


00873 
02618 
04363 
06109 
07854 

09599 
11345 
13090 
14835 
16581 

18326 
20071 
21817 
23562 
25307 

27053 
28798 
30543 
32289 
34034 

35779 
37525 
39270 
41015 
42761 

44506 
46251 
47997 
49742 
51487 

53233 
54978 
56723 
58469 
60214 

61959 
63705 
65450 
67195 
68941 

70686 
72431 
74176 
75922 
77667 


01047 
02793 
04538 
06283 
08029 

09774 
11519 
13265 
15010 
16755 

18500 
20246 
21991 
23736 
25482 

27227 
28972 
30718 
32463 
34208 

35954 
37699 
39444 
41190 
42935 

44680 
46426 
48171 
49916 
51662 

53407 
55152 
56898 
58643 


62134 
63879 
65624 
67370 
69115 

70860 
72606 
74351 
76096 

77842 


•01222 
•02967 
•04712 
•06458 
•08203 

•09948 
11694 
13439 
15184 
16930 

18675 
20420 
22166 
23911 
25656 

27402 
29147 
30892 
32638 
34383 

36128 

37874 
39619 
41364 
43110 

44855 
46600 
48346 
50091 
51836 

53582 
55327 
57072 
58818 
60563 

62308 
64054 
65799 
67544 
69290 

71035 

72780 
74526 
76271 
78016 


01396 

03142 

0488 

06632 

08378 

10123 
11868 
13614 
15359 
17104 

18850 
20595 
22340 
24086 
25831 

27576 
29322 
31067 
32812 
34558 

36303 
38048 
39794 
41539 
43284 

45029 
46775 
48520 
50265 
52011 

53756 
55501 
57247 
58992 
60737 

62483 
64228 
65973 
67719 
69464 

71209 
72955 
74700 
76445 
78191 


APPENDIX  II. 
Table  XIII. — Continued. 


625 


■fi 
9 

8 
P 

0'. 

6'. 

12'. 

18'. 

24'. 

30'. 

36'. 

42'. 

48'. 

54'. 

45 

•78540 

•78714 

•78889 

•79063 

•79238 

•79412 

•79587 

•79762 

•79936 

•80114 

46 

•80285 

•80460 

•80634 

•80809 

•80983 

•81158 

•81332 

•81507 

•81681 

•81856 

47 

•82030 

•82205 

•82380 

•82554 

•82729 

•82903 

•83078 

•83252 

•83427 

•83601 

48 

•83776 

•83950 

•84125 

•84299 

•84474 

•84648 

•84823 

•84998 

•85172 

•85347 

49 

•85521 

•85696 

•85870 

•86045 

•86219 

•86394 

•86568 

•86743 

•86917 

•87092 

50 

•87266 

•87441 

•87616 

•87790 

•87965 

•88139 

•88314 

•88488 

•88663 

•88837 

51 

•89012 

•89186 

•89361 

•89535 

•89710 

•89884 

•90059 

•90234 

•90408 

•90583 

52 

•90757 

•90932 

•91106 

•91281 

•91455 

•91630 

•91804 

•91979 

•92153 

•92328 

53 

•92502 

•92677 

•92852 

•93026 

•93201 

•93375 

•93550 

•93724 

•93899 

•94073 

54 

•94284 

•94422 

•94597 

•94771 

•94946 

•95120 

•95295 

•95470 

•95644 

•95819 

55 

•95993 

•96168 

•96342 

•96517 

•96691 

•96866 

•97040 

•97215 

•97389 

•97564 

5G 

•97738 

•97913 

•98088 

•98262 

•93437 

•98611 

•98786 

•98960 

•99135 

•99309 

57 

•99484 

•99658 

•99833 

1-00007 

1-00182 

1-00356 

1-00531 

1-00706 

1-00880 

1-01055 

58 

1-01229 

1-01404 

1-01578 

1-01753 

1-01927 

1-02102 

1-02276 

1-02451 

1-02625 

1-02800 

59 

1-02974 

1-03149 

1-03323 

1-03498 

1-03673 

1-03847 

1-04022 

1-04196 

1-04371 

1-04545 

GO 

1-04720 

1-04894 

1-05069 

1-05243 

1-05418 

1-05592 

1-05767 

1-05941 

1-06116 

1-06291 

61 

1-06465 

106640 

1-06814 

1-06989 

107163 

1-03387 

1-07512 

1-07687 

1-07861 

1-08036 

62 

1-08210 

1-08385 

1-08559 

1-08734 

1-08909 

1-09083 

1-09258 

1-09432 

1-09607 

1-09781 

63 

1-09956 

1-10130 

1-10305 

1-10479 

1-10654 

1-10828 

1-11003 

1-11177 

1-11352 

1-11527 

64 

1-11701 

1-11876 

1-12050 

1-12225 

1-12399 

1-12574 

1-12748 

1-12923 

1-13097 

1-13272 

65 

1-13446 

1-13621 

1-13795 

1-13970 

1-14145 

1-14319 

1-14494 

1-14668 

1-14843 

1-15017 

66 

1-15192 

1-15366 

1-15541 

1-15715 

1-15890 

1-16064 

1-16239 

1-16413 

1-16588 

1-16763 

67 

1-16937 

1-17112 

1-17286 

1-17461 

1-17635 

1-17810 

1-17984 

1-18152 

1-18333 

1-18508 

68 

1-18682 

1-18857 

1-19031 

1-19206 

1-19381 

1-19555 

1-19730 

1-19904 

1-20079 

1-20253 

69 

1-20428 

1-20602 

1-20777 

1-20951 

1-21126 

1-21300 

1-21475 

1-21649 

1-21824 

1-21999 

70 

l-2217c 

1-22348 

1-22522 

1-22697 

1-22871 

1-23046 

1-23220 

1-23395 

1-23569 

1-23744 

71 

1-23918 

1-24093 

1-24267 

1-24442 

1-24617 

1-24791 

1-24966 

1-25140 

1-25315 

1-25489 

72 

1-25664 

L-25838 

1-26013 

1-26187 

1-26362 

1-26536 

1-26711 

1-26885 

1-27060 

1-27235 

73 

1-27409 

1-27584 

1-27758 

1-27933 

1-28107 

1-28282 

1-28456 

1-28631 

1-28805 

1-28980 

74 

1-29154 

1-29329 

1-29503 

1-29678 

1-29852 

1-30027 

1-30202 

1-30376 

1-30551 

1-30725 

75 

1-30900 

1-31074 

1-31249 

1-31423 

1-31598 

1-31772 

1-31947 

1-32121 

1*32296 

1-32470 

76 

1-32645 

1-32820 

1-32994 

1-33169 

1-33343 

1-33518 

1-33692 

1-33867 

1-34041 

1-34216 

77 

1-34390 

1-34565 

1-34739 

1-34914 

1-35088 

1-35263 

1-35438 

1-35612 

1-35787 

1-35961 

78 

1-36136 

1-36310 

1-36485 

1-36659 

1-36834 

1-37008 

1-37183 

1-37357 

1-37532 

1-37706 

79 

1-37881 

1-38056 

1-38230 

1-38405 

1-38579 

1-38754 

1-38928 

1-39103 

1-39277 

1-39452 

80 

1-39626 

1-39801 

1-39975 

1-40150 

1-40324 

1-40499 

1-40674 

1-40848 

1-41023 

1-41197 

81 

1-41372 

1-41546 

1-41721 

1-41895 

1-42070 

1-42244 

1-42419 

1-42593 

1-42768 

1-42942 

82 

1-43117 

1-43292 

1-43466 

1-43641 

1-43815 

1-43990 

1-44164 

1-44339 

1-44513 

1-44688 

83 

1-44862 

1-45037 

1-45211 

1-45386 

1-45560 

1-45735 

1-45910 

1-46084 

1-46259 

1-46433 

84 

1-46608 

1-46782 

1-46957 

1-47131 

1-47306 

1-47480 

1-47655 

1-47829 

1-48004 

1-48178 

85 

1-48353 

1-48528 

1-48702 

1-48877 

1-49051 

1-49226 

1-49400 

1-49575 

1-49749 

1-49924 

86 

1-50098 

1-50273 

1-50447 

1-50622 

1-50796 

1-50971 

1-51146 

1-51320 

1-51495 

1-51669 

87 

1-51844 

1-52018 

1-52193 

1-52367 

1-52542 

1-52716 

1-52891 

1-53065 

1-53240 

1-53414 

88 

1-53589 

1-53764 

1-53938 

1-54113 

1-54287 

1-54462 

1-54636 

1-54811 

1-54985 

1-55160 

89 

1-55334 

1-55509 

1-55683 

1-55858 

1-56032 

1-56207 

1-56382 

1-56556 

1-56731 

1-56905 

Rli 


G26 


HIGHER  MATHEMATICS. 


Table  XIY.— Numerical  Values  of  some  Trigonometrical 

Ratios. 

(Page  609.) 

Table  XY.— Signs  of  the  Trigonometrical  Ratios. 

(Page  610.) 

Table  XYL— Comparison  of  Hyperbolic  and  Trigonometrical 

Functions. 

(Page  614.) 

Table  XYII—  Numerical  Yalues  of  e*2  and  e~*2  from 
x  =  0-1  to  x  =  50. 


X. 

e*\ 

•  -* 2. 

X. 

•f. 

e-*2. 

0-1 

1-0101 

0-99005 

2-6 

8-6264  x  102 

1-1592  x  10  "3 

0-2 

1-0408 

•96079 

2-7 

1-4656  x  103 

6-8233  x  10  ~* 

0-3 

1-0904 

•91393 

2-8 

2-5402       „ 

^•9367       „ 

0-4 

1-1735 

•85214 

2-9 

4-4918       „ 

2-2263       „ 

0-5 

1-2840 

•77880 

3-0 

8-1031 

1-2341       „ 

0-6   • 

1-4333 

0-69768 

3-1 

1-4913  x  104 

6-7055  x  10  ~5 

0-7 

1-6323 

•61263 

3-2 

2-8001 

3-5713       „ 

0-8 

1-8965 

•52729 

3-3 

5-2960 

1-8644       „ 

0-9 

2-2479 

•44486 

3-4 

1-0482  x  105 

9-5402  x  10 -« 

1-0 

2-7183 

•36788 

3-5 

2-0898 

4-7851       „ 

1-1 

3-3535 

0-29820 

3-6 

4-2507  x  105 

2-3526  x  10"6 

1-2 

4-2207 

•23693 

3-7 

8-8205       „ 

1-1337       „ 

1-3 

5-4195 

•18452 

3-8 

1-8673  x  106 

5-3554  x  10  ~7 

1-4 

7-0993 

•14086 

3-9 

4-0929 

2-4796       „ 

1-5 

9-4877 

•10540 

4-0 

8-8861        „ 

1-1254       „ 

1-6 

12-936 

0-077306 

4-1 

1-9976  x  107 

5-0062  x  10  ~8 

1-7 

17-993 

•055576 

4-2' 

4-5809 

2-1829       „ 

1-8   • 

25-534 

•039164 

4-3 

1-0718  x  108 

9-3303  x  10  "9 

1-9 

36-996 

•027052 

4-4 

2-5583 

3-9088       „ 

2-0 

54-598 

•018316 

4-5 

6-2297       „ 

1-6052       „ 

2-1 

82-269 

0-012155 

4-6 

1-5476  x  109 

6-4614  x  10--° 

2-2 

126-47 

•0279070 » 

4-7 

3-9228  •     „ 

2-5494       „ 

2-3  • 

198-34 

•0250418 

4-8 

1-0143  x  1010 

9-8595  x  10  -" 

2-4 

317-35 

•0231511 

4-9 

2-6755       „ 

3-7376       „ 

2-5 

518-02 

•0219304 

5-0 

7*2005 

1-3888       „ 

iQ-02555  means  0*00555  :  0-0*55  means  0-000055. 


APPENDIX  II.  627 


Table  XYII1. — Natural  Logarithms  of  Numbers. 

Many  formulae  require  natural  logarithms,  and  it  is  convenient  to  have  at 
hand  a  table  of  these  logarithms  to  avoid  the  necessity  of  having  recourse  to 
the  conversion  formulae,  page  28.    Table  XVIII.  is  used  as  follows : 

J.  For  numbers  greater  than  10t,  follow  the  method  of  Ex.  (2)  and  (3)  below. 

II.  For  numbers  between  1  and  10  not  in  the  table,  use  interpolation 
formulae,  say  proportional  parts. 

III.  For  numbers  less  than  1,  use  the  method  of  Ex.  (5)  below. 

If  there  is  going  to  be  much  trouble  finding  the  natural  log  it  may  be 
better  to  use  standard  tables  of  logarithms  to  base  10  and  multiply  by  2-3026 
in  the  ordinary  way. 

logelO  =  2-3026. 

Examples.— (1)  Show  that  log,*  =  loge(3-1416)  =  1-1447. 

(2)  Required  the  logarithm  of  5,540  to  the  base  e.    Here 

log,5,540  =  log.(5-640  x  1,000)  =  loge(5-54  x  103) ; 
hence,  logtf5,540  =  log.5'54  +  3  loge10  =  8-6198. 

(3)  Show  thai  loge100  =  4-6052  ;  log,l,000  =  6-9078 ;  log810,000  -  9-2103 
logJOO.OOO  =  11-5129.    Hint,  log  1,000  =  log  103  =  3  log  10. 

(4)  If  100  c.c.  of  a  gas  at  a  pressure  of  5,000  grams  per  square  centimetre 
expands  until  the  gas  occupies  a  volume  of  557  c.c,  what  work  is  done  during 
the  process  ?     From  page  254, 

W  =  pw  loge^  =  5,000  x  100  x  loge5-57  =  850,700  grm.  cm. 

If  a  table  of  ordinary  logarithms  had  been  employed  we  should  have  written 
2-3026  x  logi05-57  in  place  of  loge5'57. 

(5)  Find  loge0-00051 ;  log  0-0031 ;  and  log  0-51.     Here  we  have 

log  0-00051  -  log  5-1  -  log  10,000  =  log  5-1  -  log  104  =  log  5-1  -  4  log  10  = 
1-6292  -  4  x  2-3026  =  1*6292  -  9-2104  =  9-4188,  or  -  8-5812 ;  loge0-0031  = 
1-1314  -  6-9077  =  6-2237,  or  -  5-7763 ;  log  0-51  =  0-6292  -  2-3026  =  2-3166 
or  -  1-6734. 

The  bar  over  the  first  figure  has  a  similar  meaning  to  the  "  bar."  of  ordi 
nary  logs. 


628 


HIGHER  MATHEMATICS. 


n. 

•00. 

•01. 

•02. 

•03. 

•04. 

•05. 

•06. 

•07. 

•08. 

•09. 

1-0 

o-oooo 

00100 

00198 

0-0296 

0-0392 

00488 

0-0583 

0-0677 

0-0770 

0-0862 

1-1 

•0953 

•1044 

•1133 

•1222 

•1310 

•1398 

•1484 

•1570 

•1655 

•1740 

1-2 

•1823 

•1906 

•1989 

•2070 

•2151 

•2231 

•2311 

•2390 

•2469 

•2546 

1-3 

•2624 

•2700 

•2776 

•2852 

•2927 

•3001 

•3075 

•3148 

•3221 

•3293 

1-4 

•3365 

•3436 

•3507 

•3577 

•3646 

•3716 

•3784 

•3853 

•3920 

•3988 

1-5 

0-4055 

0-4121 

0-4187 

0-4253 

0-4318 

0-4383 

0-4447 

0-4511 

0-4574 

0-4637 

1-6 

•4700 

•4762 

•4824 

•4886 

•4947 

•5008 

•5068 

•5128 

•5188 

•5247 

1-7 

•5306 

•5365 

•5423 

•5481 

•5539 

•5596 

•5653 

•5710 

•5766 

•5822 

1-8 

•5878 

•5933 

•5988 

•6043 

•6098 

•6152 

•6206 

•6259 

•6313 

•6366 

1-9 

•6419 

•6471 

•6523 

•6575 

•6627 

•6678 

•6729 

•6780 

•6831 

•6881 

2-0 

0-6932 

0-6981 

0-7031 

0-7080 

0-7130 

0-7178 

0-7227 

0-7276 

0-7324 

0-7372 

2-1 

•7419 

•7467 

•7514 

•7561 

•7608 

•7655 

•7701 

•7747 

•7793, 

•7839 

2-2 

•7885 

•7930 

•7975 

•8020 

•8065 

•8109 

•8154 

•8198 

•8242 

•8286 

2-3 

•8329 

•8372 

•8416 

•8459 

•8502 

•8544 

•8587 

•8629 

•8671 

•8713 

2-4 

•8755 

•8796 

•8838 

•8879 

•8920 

•8961 

•9002 

•9042 

•9083 

•9123 

2-5 

0-9163 

0-9203 

0-9243 

0-9282 

0-9322 

0-9361 

0-9400 

0-9439 

0-9478 

0-9517 

2-6 

•9555 

•9594 

•9632 

•9670 

•9708 

•9746 

•9783 

•9821 

•9858 

•9895 

2-7 

•9933 

•9970 

1-0006 

1-0043 

1-0080 

1-0116 

1-0152 

1-0189 

1-0225 

1-0260 

2-8 

1-0296 

10332 

•0367 

•0103 

•0438 

•0472 

•0508 

•0543 

•0578 

•0613 

2-9 

•0647 

•0682 

•0716 

•0750 

•0784 

•0818 

•0852 

•0886 

•0919 

•0953 

3-0 

1-0986 

1-1019 

1-1053 

1-1086 

1-1119 

1-1151 

1-1184 

1-1217 

1-1249 

1-1282 

3-1 

•1314 

.-1346 

•1378 

•1410 

•1442 

•1474 

•1506 

•1537 

•1569 

•1600 

3-2 

•1632 

•1663 

•1694 

•1725 

•1756 

•1787 

•1817 

•1848 

•1878 

•1909 

3-3 

•1939 

•1970 

•2000 

•2030 

•2060 

•2090 

•2119 

•2149 

•2179 

•2208 

3-4 

•2238 

•2267 

•2296 

•2326 

•2355 

•2384 

•2413 

•2442 

•2470 

•2499 

3-5 

1-2528 

1-2556 

1-2585 

1-2613 

1-2641 

1-2670 

1-2698 

1-2726 

1-2754 

1-2782 

3-6 

•2809 

•2837 

•2865 

•2892 

•2920 

•2947 

•2975 

•3002 

•3029 

•3056 

3-7 

•3083 

•3110 

•3137 

•3164 

•3191 

•3218 

•3244 

•3271 

•3297 

•3324 

3-8 

•3350 

•3376 

•3403 

•3429 

•3455 

•3481 

•3507 

•3533 

•3558 

•3584 

3-9 

•3610 

•3635 

•3661 

•3686 

•3712 

•3737 

•3762 

•3788 

•3813 

•3838 

4-0 

1-3863 

1-3888 

1-3913 

1-3938 

1-3963 

1-3987 

1-4012 

1-4036 

1-4061 

1-4086 

4-1 

•4110 

•4134 

•4159 

•4183 

•4207 

•4231 

•4255 

•4279 

•4303 

•4327 

4-2 

•4351 

•4375 

•4398 

•4422 

•4446 

•4469 

•4493 

•4516 

•4540 

•4563 

4-3 

•4586 

•4609 

•4633 

•4656 

•4679 

•4702 

•4725 

•4748 

•4771 

•4793 

4-4 

•4816 

•4839. 

•4861 

•4884 

•4907 

•4929 

•  -4954 

•4974 

•4996 

•5019 

4-5 

1-5041 

1-5063 

1-5085 

1-5107 

1-5129 

1-5151 

1-5173 

1-5195 

1-5217 

1-5239 

4-6 

•5261 

•5282 

•5304 

•5326 

•5347 

•5369 

•5390 

•5412 

•5433 

•5454 

4-7 

•5476 

•5497 

•5518 

•5539 

•5560 

•5581 

•5602 

•5623 

•5644 

•5665 

4-8 

•5686 

•5707 

•5728 

•5748 

•5769 

•5790 

•5810 

•5831 

•5851 

•5872 

4-9 

•5892 

•5913 

•5933 

•5953 

•5974 

•5994 

•6014 

•6034 

•6054 

•6074 

5-0 

1-6094 

1-6114 

1-6134 

1-6154 

1-6174 

1-6194 

1-6214 

1-6233 

1-6253 

1-6273 

5-1 

•6292 

•6312 

•6332 

•6351 

•6371 

•6390 

•6409 

•6429 

•6448 

•6467 

5-2 

•6487 

•6506 

•6525 

•6544 

•6563 

•6582 

•6601 

•6620 

•6639 

•6658 

5-3 

•6677 

•6696 

•6715 

•6734 

•6752     -6771 

•6790 

•6808 

•6827 

•6845 

5-4 

•6864 

•6882 

•6901 

•6919 

•69381   -6956 

•6975 

•6993. 

•7011 

•7029 

APPENDIX  II. 


629 


n. 

•00. 

•01. 

•02. 

•03. 

•04. 

•05. 

•06. 

•07. 

•08. 

•09. 

5-5 

1-7048 

1-7066 

1-7083 

1-7102 

1-7120 

1-7138 

1-7156 

1-7174 

1-7192 

1-7210 

5-6 

•7228 

•7246 

•7263 

•7281 

•7299 

•7317 

•7334 

•7352 

•7370 

•7387 

5-7 

•7405 

•7422 

•7440 

•7457 

•7475 

•7492 

•7509 

•7527 

•7544 

•7561 

5-8 

•7579 

•7596 

•7613 

•7630 

•7647 

•7664 

•7682 

•7699 

•7716 

•7733 

5-9 

•7750 

•7766 

•7783 

•7800 

•7817 

•7834 

•7851 

•7868 

•7884 

•7901 

6-0 

1-7917 

1-7934 

1-7951 

1-7967 

1-7984 

1-8001 

1-8017 

1-8034 

1-8050 

1-8067 

6-1 

•8083 

•8099 

•8116 

•8132 

•8148 

•8165. 

•8181 

•8197 

•8213 

•8229 

6-2 

•8246 

•8262 

•8278 

•8294 

•8310 

•8326 

•8342 

•8358 

•8374 

•8390 

6-3 

•8406 

•8421 

•8437 

•8453 

•8469 

•8485 

•8500 

•8516 

•8532 

•8547 

6-4 

•8563 

•8579 

•8594 

•8610 

•8625 

•8641 

•8656 

•8672 

•8687 

•8703 

6-5 

1-8718 

1-8733 

1-8749 

1-8764 

1-8779 

1-8795 

1-8810 

1-8825 

1-8840 

1-8856 

6-6 

.  -8871 

•8886 

•8901 

•8916 

•8931 

•8946 

•8961 

•8976 

•8991 

•9006 

6-7 

•9021 

•9036 

•9051 

•9066 

•9081 

•9095 

•9110 

•9125 

•9140 

•9155 

6-8 

•9169 

•9184 

•9199 

•9213 

•9228 

•9243 

•9257 

•9272 

•9286 

•9301 

6-9 

•9315 

•9330 

•9344 

•9359 

•9373 

•9387 

•9402 

•9416 

•9431 

•9445 

7-0 

1-9459 

1-9473 

1-9488 

1-9502 

1-9516 

1-9530 

1-9544 

1-9559 

1-9573 

1-9587 

7-1 

•9601 

•9615 

•9629 

•9643 

•9657 

•9671 

•9685 

•9699 

•9713 

•9727 

7-2 

•9741 

•9755 

•9769 

•9782 

•9796 

•9810 

•9824 

•9838 

•9851 

•9865 

7-3 

•9879 

•9892 

•9906 

•9920 

•9933 

•9947 

•9961 

•9974 

•9988 

2-0001 

7-4 

20015 

2-0028 

2-0042 

20055 

2-0069 

2-0082 

2-0096 

2-0109 

20122 

•0136 

7-6 

2  0149 

2-0162 

2-0176 

20189 

2-0202 

2-0216 

2-0229 

2-0242 

2-0255 

20268 

7-6 

•0282 

•0295 

•0308 

•0321 

•0334 

•0347 

•0360 

•0373 

•0386 

•0399 

7-7 

•0412 

•0425 

•0438 

•0451 

•0464 

•0477 

•0490 

•0503 

•0516 

•0528 

7-8 

•0541 

•0554 

•0567 

•0580 

•0592 

•0605 

•0618 

•0631 

•0643 

•0656 

7-9 

•0669 

•0681 

•0694 

•0707 

•0719 

•0732 

•0744 

•0757 

•0769 

•0728 

8-0 

2-0794 

2-0807 

2-0819 

20832 

2-0844 

2-0857 

2-0869 

2-0882 

2-0894 

2-0906 

8-1 

•0919 

•0931 

•0943 

•0956 

•0968 

•0980 

•0992 

•1005 

•1017 

•1029 

8-2 

•1041 

•1054 

•1066 

•1078 

•1090 

•1102 

•1114 

•1126 

•1138 

1151 

8-3 

•1163 

•1175 

•1187 

•1199 

•1211 

•1223 

•1235 

•1247 

•1259 

•1270 

8-4 

•1282 

•1294 

•1306 

•1318 

•1330 

•1342 

•1354 

•1365 

•1377 

•1389 

8-5 

2-1401 

2-1412 

2-1424 

2-1436 

2-1448 

2-1459 

2-1471 

2-1483 

2-1494 

2-1506 

8-6 

•1518 

•1529 

•1541 

•1552 

•1564 

•1576 

•1587 

•1599 

•1610 

•1622 

8-7 

•1633 

•1645 

•1656 

•1668 

•1679 

•1691 

•1702 

•1713 

•1725 

1736 

8'8 

•1748 

•1759 

•1770 

•1782 

•1793 

•1804 

•1816 

•1827 

•1838 

•1849 

8-9 

•1861 

•1872 

•1883 

•1894 

•1905 

•1917 

•1928 

•1939 

•1950 

•1961 

9-0 

2-1972 

2-1983 

2-1994 

2-2006 

2-2017 

2-2028 

2-2039 

2-2050 

2-2061 

2-2072 

9-1 

•2083 

•2094 

•2105 

•2116 

•2127 

•2138 

•2149 

•2159 

•2170 

•2181 

9-2 

•2192 

•2203 

•2214 

•2225 

•2235 

•2246 

•2257 

•2268 

•2279 

•2289 

9-3 

•2300 

•2311 

•2322 

•2332 

•2343 

•2354 

•2364 

•2375 

•2386 

•2396 

9-4 

•2407 

•2418 

•2428 

•2439 

•2450 

•2460 

•2471 

•2481 

•2492 

•2502 

9-5 

2-2523 

2-2523 

2-2534 

2-2544 

2-2555 

2-2565 

2-2576 

2-2586 

2-2597 

2-2607 

96 

•2628 

•2628 

•2638 

•2649 

•2659 

•2670 

•2680 

•2690 

•2701 

•2711 

9-7 

•2721 

•2732 

•2742 

•2752 

•2762 

•2773 

•2783 

.2792 

•2803 

•2814 

9-8 

•2824 

•2834 

•2844 

•2854 

•2865 

•2875 

•28  85 

•2895 

•2905 

•2915 

9-9 

•2935 

•2935 

•2946 

•2956 

•2966 

•2976 

•2986 

•2996 

•3006 

•3016 

INDEX. 


Abegg,  371. 
Abscissa,  84. 

—  axis,  83. 
Absolute  error,  276. 

—  zero,  12. 
Acceleration,  17,  65. 

—  curve,  102. 

—  Normal,  179. 

—  Tangential,  179. 

—  Total,  179. 
Accidental  errors,  510. 
Acetochloranilide,  8. 
Acnode,  171. 
Addition,  273. 
Adrian,  515. 

Airy,  G.  B.,  451. 

d'Alembert's  equation,  459. 

Algebra.    Laws  of,  177. 

Algebraic  functions,  35. 

Alternando,  133. 

Amagat,  176. 

Amago,  74. 

Amount  of  substance,  6. 

Ampere,  29. 

Amplitude,  137,  427. 

Angles.     Measurement  of,  606. 

Circular,  606,  624. 

Radian,  606,  624. 

—  Vectorial,  114. 
Angular  velocity,  137. 
Anti-differential,  190. 
Aperiodic  motion,  410. 
Approximate  calculations,  276. 

—  integration,  335,  463. 
Approximations.  Solving  differential  equa- 
tions by  successive,  463. 

Arc  of  circle  (length),  603. 

Archimedes'  spiral,  117,  246. 

Areas  bounded  by  curves,  230,  234,  237. 

Arithmetical  mean,  235. 

Arrhenius,  S.,  146,  215,  332,  342. 

Association.    Law  of,  177. 

Asymptote,  104. 

August,  171. 

Atomic  weghts,  540. 

Austen,  VV.  C.  Roberts,  151. 

Auxiliaries,  559. 

Auxiliary  equation,  400. 

Lagrange's,  453. 

Average,  235. 

—  error,  525. 


Average  velocity,  7. 
Averages.     Method  of,  536. 
Axes.     Transformation  of,  96. 
Axis.     Abscissa,  83. 

—  Conjugate,  102. 

—  Co-ordinate,  83,  121. 

—  Imaginary,  102. 

—  Major,  100. 

—  Minor,  100. 

—  Oblique,  83. 

—  of  iinaginaries,  177. 

—  of  reals,  177. 

—  of  revolution,  248. 

—  Real,  102. 

—  Rectangular,  83. 

—  Transverse,  102. 

Bacon,  F.,  4,  273. 
Bancroft,  W.  D.,  120. 
Bayer,  605. 
Baynes,  R.  E.,  594. 
Berkeley,  G.,  32. 
Bernoulli,  572,575. 
Bernoulli's  equation,  388,  389. 

—  series,  290. 
Berthelot,  M.,  3,  227. 
Berthollet,  191. 
Bessel,  311,  514. 
Bessel's  formula,  523. 
Binomial  series,  36,  282. 
Biot,  55,  74,  319. 
Blanksma,  J.  J.,  223. 
Bodenstein,  M.,  222,  228. 
Boiling  curve,  174. 
Bolton,  C.,  290. 

Bolza,  O.,  579. 

Bosscha,  63. 

Boyle,  19,  20,  21,  23,  45,  46,  62,  114,  254, 

444,  457,  596. 
Boynton,  W.  P.,  114,  260. 
Brachistochrone,  572. 
Bradley,  514. 
Bradshaw,  L.,  390,  442. 
Break,  143. 
Bredig,  G.,  337. 
Bremer,  G.  J.  W.,  328. 
Briggsian  logarithms,  25. 
Bruckner,  C,  303,  308,  356. 
Bunsen,  R,  270. 
Burgess,  J.,  344. 
Byerly,  W.  E.,  467,  481. 


631 


632 


INDEX 


Cailletet,  L.P.,150. 
Calculations.    Approximate,  276. 

—  with  small  quantities,  601. 
Calculus.     Differential,  19. 

—  finite  differences,  308. 

—  Integral,  184. 

—  variations,  567.   • 
Callendar,  39. 
Callum,  561. 
Cane  sugar,  6. 
Cardan,  353. 
Carnot,  32,  34. 
Carnot's  function,  386. 
Cartesian  co-ordinates,  84. 
Catenary,  348. 
Cavendish,  527,  565. 
Cayley,  169,  170,  603. 
C-discriminant,  394. 
Centnerszwer,  M.,  329. 

Central    differences.      Interpolation    by, 

315. 
Centre,  98,  100,  101. 
, —  of  curvature,  180. 

—  of  gravity,  60s. 
Ceratoid  cusp,  170. 
Characteristic  equation,  79. 
Charles'  law,  21,  24,  91,  596. 
Charpit,  454. 

Chatelier,  H.  le,  318,  539. 
Chatelier's  theorem,  265. 
Chauvenet's  criterion,  563,  623. 
Chloracetanilide,  8. 
Chord  of  circle  (length),  603. 
Christoffel,  42. 
Chrvstal.  G.,351,  364. 
Circle,  97,  121. 

—  (area  of),  604. 

—  Arc  of  (length),  604. 

—  Chord  of  (length),  604. 

—  Perimeter  of  (length),  604. 

—  of  curvature,  180. 

—  Osculatory,  180. 
Circular  functions,  346. 

—  measure  of  angles,  606. 

—  sector  (area),  605. 

—  segment  (area),  605. 

Clairaut,  A.  C,  192,  391,  393,  457,  561. 
Clapeyron,  E.,  453,  457. 
Clapeyron's  work  diagram,  239. 
Clarke,  F.  W.,  55i,  554,  562. 
Clausius,  R.,  6,  504. 
Clement,  81. 

—  J.  K.,350. 

Coexistence  of  different  reactions.    Prin- 
ciple of,  70. 
Cofactor  (determinant),  589. 
Collardeau,  E.,  150. 
Colvill,  W.  HM  275. 
Combinations,  602. 
Common  logarithms,  25,  27. 
Commutation.    Law  of,  177. 
Comparison  test  (convergent  series),  271. 
Complanation  of  surfaces,  247. 
Complement  (determinant),  589. 

—  Error  function,  344. 

—  of  angles,  610. 


Complementary  function,  413. 
Complete  differentials,  77. 

—  elliptic  integrals,  426. 

—  integral,  377,  450. 

—  solution  of  differential  equation,  377- 
Componendo,  133. 

—  et  dividendo,  133. 
Composition  of  a  solution,  88. 
Compound  interest  law,  56. 
Comte,  A.,  3. 

Concavity  of  curves,  159. 
Condensation.     Retrograde,  175. 
Conditional  equation,  218,  353. 
Conditioned  maxima  and  minima,  301. 

—  observations,  555. 
Conditions.     Limiting,  363,  452. 
Conduction  of  heat,  493. 
Cone,  135. 

—  (centre  of  gravity),  605. 

—  (surface  area),  604. 

—  (volume),  605. 
Conic  sections,  97. 
Conicoids,  595. 
Conjugate  axis,  102. 

—  determinant,  590. 

—  point,  171. 

Conrad,  M.,  303,  308,  356. 
Consistent  equations.     Test  for,  585. 
Constant,  19,  324. 

—  errors,  537. 

—  Integration,  193,  198,  234. 

—  of  Fourier' 8  series,  471. 

—  Phase,  137. 

Constituent  (determinant),  581. 
Contact  of  curves,  291. 

—  Orders  of,  291. 
Continuous  function,  142. 
Convergent  series,  267. 

Test  for,  271. 

Convertendo,  133. 
Convexity  of  curves,  159. 
Cooling  curves,  150.      . 
Co-ordinate  axis,  83,  122. 

—  plane,  122. 
Co-ordinates.     Cartesian,  84. 

—  Generalized,  140. 

—  Polar,  114. 

—  Transformation  of,  115. 

—  Trilinear,  118. 
Correction  term,  278. 
Cosecant,  607. 
Cosine,  608. 

—  Direction,  124. 

—  Hyperbolic,  347,  613. 

—  series,  283,  474. 

Euler's,  285. 

Cotangent,  607. 

Cotes  and  Newton's  interpolation  formula, 

337. 
Cottle,  G.  J.,  223. 
Criterion.    Chauvenet's,  663. 

—  of  integrability,  -446 
Critical  temperature,  150. 
Crompton,  H.,  146. 

Crookes,  W.,  229,  280,  531,  533,  565. 
Crunode,  169. 


INDEX 


633 


Cubature  of  solids,  248. 
Curvature,  178. 

—  Centre  of,  180. 

—  Circle  of,  180. 

—  Direction  of,  181. 

—  Radius  of,  180. 
Curve,  85. 

—  Equation  of,  85. 

—  Error,  512. 

—  Frequency,  512. 

—  Imaginary,  177. 

—  Orders  of,  120. 

—  Plotting,  86. 

—  Probability,  512. 

—  Sine,  136. 

—  Smoothed,  149. 
Cusp,  169. 

—  Ceratoid,  170. 

—  Double,  170. 

—  First  species,  170. 

—  locus,  394. 

—  Rhamphoid,  170. 

—  Second  species,  170. 

—  Single,  170. 
Cycle,  239. 
Cyclic  process,  239. 
Cycloid,  443,  573. 
Cylinder,  135. 

—  (surface  area),  604. 

—  (volume),  605. 

Dalton,  491. 
Dalton's  law,  64,  285. 
Damped  oscillations,  404,  409. 
Damping  ratio,  409. 
Danneel,  H.,  197,  215,  217. 
Darwin,  G.  H.,  537. 
Decrement.    Logarithmic,  409. 
Definite  integral,  187,  230,  234,  240. 

Differentiation  of,  577* 

Degree,  607. 

—  of  differential  equatiou,  378. 

—  of  freedom,  140. 
Demoivre's  theorem,  351,  613. 
Dependent  variable,  8. 
Descartes,  R.,  84,  498. 
Determinant,  580. 

—  Conjugate,  590. 

—  Differentiation  of,  590. 

—  Multiplication  of,  589. 

—  Order  of,  581. 

—  Properties  of,  587. 

—  Skew,  590. 

—  Symmetrical,  590. 
Developable  surface,  134. 
Dew  curve,  175. 
Diagrams.    Work,  239. 
Clapeyron's,  239. 

Difference  formulae.   Differentiation  of,J 
Differences.     Calculus  of  finite,  308. 

—  Central.    Interpolation  by,  315. 

—  Orders  of,  308. 

—  Table  of,  309. 
Differential,  10,  83,  568. 

—  calculusj  19. 


Differential,  coefficient,  8. 
Second,  18. 

—  Complete,  77. 

—  equation,  66,  371,  374,  378. 
Degree  of,  378. 

Order  of,  378. 

—  —  Solving,  371. 

—  Exact,  77,  384. 

—  Partial,  69,  448. 

—  Total,  69. 
Differentiation,  19. 

—  by  graphic  interpolation,  319. 

—  Integration  by,  495. 

—  Methods  of,  8. 

—  of  definite  integrals,  577. 

—  of  determinants,  590. 

—  of  difference  formulae,  320. 

—  of  hyperbolic  functions,  349. 

—  of  numerical  relations,  318. 

—  Partial,  68. 

—  Solution  of  differential  equations  by 

391. 

—  Successive,  64. 

partial,  76. 

Diffusion  law.    Fick's,  483. 
Fourier's,  482. 

—  of  gases,  199,  491. 

—  of  heat,  493. 

—  of  salts,  483. 
Direction  cosines,  124.   . 

—  of  curvature,  181. 
Directrix,  98. 

Discontinuous  functions,  142,  143,  149. 
Discriminant,  352. 

—  »-,  393. 

—  0-,  394. 
Dissociation,  111,  255. 

—  isotherm,  112. 
Distribution.    Law  of,  177. 
Divergent  series,  267. 
Dividendo,  133. 

—  et  componendo,  133. 
Division,  274. 

—  shortened,  275. 
Dostor's  theorem,  588. 
Double  cusps,  170. 

—  integrals,  251. 

Variation  of,  577. 

Duhem,  P.,  141. 
DiUong,  60,  273. 
Dumas,  J.  B.  A.,  533,  548. 
Dupre,  A.,  53. 

Edgeworth,  F.  Y.,  515,  519,  565. 
Elasticity.     Adiabatic,  113. 

—  Isothermal,  113. 
Elements  (determinant),  581. 

—  Leading,  583. 

—  Surface,  251. 

—  Volume,  253. 
Eliminant,  583,  586. 
Elimination,  377. 

—  equations,  559. 
Ellipse,  99,  121. 

—  (area  of),  604. 


634 


INDEX 


Ellipse,  (length  of  perimeter),  603. 
Ellipsoid,  134,  595. 
Elliptic  functions,  428,  429. 

—  integrals,  426,  427,  429. 
Empirical  formulae,  322. 
Encke,  514,  552. 
Entectic,  120. 
Envelope,  182. 

—  locus,  394. 
Epoch,  137. 
Epstein,  F.,  337. 
Equilateral  hyperbola,  109. 
Equilibrium.    Van't  Hoff's  principle,  264. 
Equation.     Conditional,  352. 

—  Differential,  64. 

—  General,  89. 

—  Identical,  35a 

—  of  curve,  85. 

—  of  line,  89. 

—  of  motion,  66. 

—  of  plane,  133. 

—  of  state,  78. 

—  Solving,  352. 

Horner,  363. 

Newton,  358. 

Sturm,  360. 

Error.    Absolute,  276 

—  Accidental,  510. 

—  Average,  525. 

—  Constant,  537. 

—  Curve  of,  512. 

—  Fractional,  541. 

—  function,  344. 
Complement,  344. 

—  Law  of,  511,  515. 

—  Mean,  524,  527,  528,  530. 

—  of  method,  540. 

—  Percentage,  276,  541. 

—  Personal,  537. 

—  Probable,  521,  524,  526,  528. 

—  Proportional,  539,  540,  545. 

—  Relative,  541. 

—  Systematic,  537. 

—  Weighted,  550. 

Esson,  W.,  332,  389,  435,  437,  438,  440. 

Etard,  88. 

Eulerian  integral,  424,  425. 

Euler's  cosine  series,  283. 

—  criterion  of  integrability,  77. 

—  sine  series,  283. 

—  theorem,  74,  449. 
Even  function  of  x,  476. 
Everett,  J.  D.,  519. 
Ewan,  T.,  63. 

Exact  differential,  77,  384. 

equation,  378,  431. 

Test  for,  432. 

Forsyth's,  432. 

•Expansion.    Adiabatic,  257. 

—  Isothermal,  254. 
Explicit  functions,  593. 
Exponential  functions,  54. 

—  series,  285. 
External  force,  413. 
Extrapolation,  92,  310. 


Factorial,  38. 

Factors.    Integrating,  77,  381,  383. 

Faraday,  M.,  5,  539. 

Federlin,  W.,  332. 

Fermat,  P.  de,  56.7. 

Fermat's  principle,  165,  299. 

Fick,  6,  492. 

Fick's  law  of  diffusion,  483,  492. 

Field,  584. 

Figures.     Significant,  274. 

Finite  differences,  308. 

First  integral,  431. 

—  law  of  thermodynamics,  81. 

—  species  of  cusp,  170. 
Fluxions,  34. 

Focal  radius,  98,  100. 

Focus,  98,  100. 

Forbes,  6. 

Force.     External,  413. 

Forced  oscillations,  413. 

Forces.    Generalized,  138,  141. 

Formulae.     Finding,  322. 

—  Reduction,  205,  208,  211. 
Forsyth,  454,  467. 

Forsyth's  test  for  exact  equations,  432. 
Fourier,  J.,  343,  467,  481. 
Fourier's  diffusion  law,  481,  482. 

—  equation,  401. 

—  integrals,  479. 

—  series,  469,  470,  477. 
Constants  of,  470. 

—  theorem,  470. 
Fraction.     Partial,  212. 

—  Vanishing,  304,  305. 
Fractional  errors,  541. 

—  index,  28. 

—  precipitation,  229. 
Free  oscillations,  414. 
Freedom.    Degrees  of,  140. 
Fresnel,  5,  30. 
Fresnel's  integral,  424. 
Frequency.     Curve  of,  512. 
Friction,  397. 

Frost,  P. ,  168. 

Frustum  of  cone  (volume),  605. 

Function,  19,  322. 

—  Complementary,  413. 

—  Continuous,  142. 

—  Discontinuous,  142,  144. 

—  Elliptic,  428. 

—  Error,  344. 

Complement,  343. 

—  Even,  474. 

—  Explicit,  593. 

—  Gamma,  423,  424. 

—  Illusory,  304. 

—  Implicit,  593. 

—  Indeterminate,  304. 

—  Multiple -valued,  241. 

—  Odd,  475. 

—  Periodic,  136. 

—  Single-valued,  242. 

—  Singular,  304. 

Fundamental  laws  of  algebra,  177. 
Fusibility.    Surface  of,  118. 


INDEX 


636 


Q-alileo,  29,  225. 

Gallitzine,  60. 

Gamma  function,  423,  424.     • 

Gas  equation,  78,  110,  139,  596. 

Gauss,  C.  F.,  176,  311,  332,  409,  513,  515, 

520,  560. 
Gauss's  law,  353. 
of  errors,  516. 

—  interpolation  formula,  315. 

—  method  of  solving  equations,  557. 
Gay  Lussac,  88,  91,  285,  510. 
Geitel,  A.  C,  440. 

General  equation,  89. 

—  integral,  451. 

—  mean,  561. 

—  solution  of  differential  equation,  377. 
Generalized  co-ordinates,  139. 

—  forces,  139,  141. 
Generator,  134. 
Geometrical  series,  268. 
Gerling,  C.  L.,  565. 

Gibb's  thermodynamic  surface,  598. 

Gilbert,  424. 

Gill,  D.,  533. 

Gilles,  L.  P.  St.,  227. 

Glaisher,  J.  W.  L.,  344. 

Goldschmidt,  H.,  215. 

Graham,  T.,  199,  486,  490. 

—  J.  C.,  497. 
Graph,  88. 

Graphic  interpolation,  318. 
Differentiation  by,  319. 

—  solution  of  equations,  355. 
Gray,  A.,  467. 

Greenhill,  A.  G.,  351,  431. 
Gregory's  series,  284. 
Gudermann,  428. 
Gudermannians,  613. 
Guldberg,  191,226,  354. 

Hagen,  507,  563. 

Halley's  law,  62,  260. 

Hamilton,  567. 

Harcourt,  A.  V., 332, 389, 435,  437, 438, 440. 

Hardy,  J.  J.,  245. 

Harkness,  J.,  45. 

Harmonic  curve,  136. 

—  motion,  135,  234. 
Hartley,  W.  N.,  332. 
Haskins,  G.  N.,  350. 
Hatchett,  603. 
Hayes,  E.  H.,  146. 

Heat.     Conduction  of,  493. 
Heaviside,  O.,  370,  377,  496. 
Hecht,  W.,  303,  308,  356. 
Hedley,  E.  P.,  332. 
Heilborn,  196. 

Helmholtz,  203,  372,  471,  472. 
Henrici,  O.,  116,  335,  472. 
Henry,  P.,  216,  227. 
Henry's  law,  87. 
Hermann,  603. 
Herschel,  J.  P.  W.,  325,  515. 
Hertz,  H.,  5,  109. 
Hessian,  592,  594. 


Hill,  M.  J.  M.,  394. 

Hinds,  561. 

Hinrichs,  G.,  519. 

Hoar  frost  line,  152. 

Hobson,  E.  W.,  267. 

Hoff,  Van't.     Principle,  264. 

Holman,  S.  W.,  563. 

Holtzmann,  C,  457. 

Homogeneous  differential  equations,  372. 

—  function,  75. 

—  simultaneous  equations,  581,  584. 
Hooke's  law,  458. 

Hopital.    Rule  of  1',  307. 
Hopkinson,  J.,  4,  332. 
Horstmann,  A.,  318,  319,  321,  326. 
Humboldt,  510. 
Hyperbola,  100,  121. 

—  equilateral,  109. 

—  rectangular,  109. 
Hyperbolic  cosine,  347. 

—  functions,  347,  612. 

Differentiation  of,  348. 

Integration  of,  349. 

—  logarithms,  25,  233. 

—  sine,  247. 

—  spiral,  117. 
Hyperboloid,  133,  595. 
Hyper-elliptic  integrals,  430. 

Ice  line,  152. 

Identical  equation,  213,  352. 
Illusory  functions,  304. 
Imaginaries.     Axis  of,  177. 
Imaginary  axis,  102. 

—  curve,  177. 

—  point,  177. 

—  quantities,  jjl76. 

—  roots,  353. 

—  semi-axis,  102. 

—  surface,  177. 
Implicit  functions,- 593. 
Indefinite  integral,  187. 
Independence  of  different  reactions.    Prin- 
ciple of,  70. 

Independent  variable,  8,  448. 
Indeterminate  functions,  304. 
Index.    Fractional,  28. 

—  law,  24,  177. 

—  of  refraction,  165. 
Inequality.     Symbols  of,  13. 
Inferior  limit,  187. 
Inflexion.    Points  of,  143,  160. 
Inflexional  tangents,  599. 

Infinite  series.    Integration  of,  341,  463. 
Infinitesimals,  18,  33. 
Infinity,  11. 

Instantaneous  velocity,  9. 
Integrability.     Criterion  of,  446. 

Eider's,  77. 

Integral,  187. 

—  Complete,  377,  450. 

—  Definite,  189,  231,  240. 

—  Differentiation  of,  577. 

—  Double,  251. 

—  Elliptic,  427,  428,  429,  430. 


636 


INDEX 


Integral,  Elliptic,  Complete,  430. 

—  Eulerian,  424,  425. 

—  First,  431. 

—  Fourier's,  479. 

—  Fresnel's,  424. 

—  General,  451. 

—  Hyper-elliptic,  430. 

—  Indefinite,  187. 

—  Limits,  187. 

—  Mean  values  of,  234. 

—  Multiple,  249. 

—  Particular,  400,  418. 

—  Probability,  516,  531,  532,  621,  622. 

—  Space,  189. 

—  Standard,  192,  193,  349, 

—  Time,  189. 

—  Ultra-elliptic,  430. 

—  Variation  of,  568,  569,  573. 
double,  577. 

triple,  577. 

Integrating  factors,  77,  380,  381. 
Integration,  184,  189. 

—  Approximate,  341,  469. 

—  by  differentiation,  495. 

—  by  infinite  series,  341. 

—  by  parts,  204. 

—  by  successive  integration,  206. 

—  constant,  193,  194,  234. 

—  formula  of  Newton  and  Cotes,  336. 

—  hyperbolic  functions,  349. 

—  Substitutes  for,  333. 

—  Successive,  249. 

—  Symbol  of,  189. 
Intercept  equation  of  line,  90. 

of  plane,  132. 

Interpolation,  310. 

—  formula,  Gauss',  315. 

Lagrange's,  311.  312. 

Newton's,  311,  312. 

Stirling's,  318,  320. 

—  Graphic,  317. 

—  —  Differentiation  by,  319. 
Inverse  sine  series,  384. 

—  trigonometrical  functions,  49. 
series,  283. 

Invert  sugar,  6,  184. 
Invertendo,  133. 
Ions,  112. 

Irrational  numbers,  178. 
Isobars,  110. 
Isometrics,  110. 
Isoperimetrical  problem,  575. 
Isopiestics,  110. 
Isothermal  expansion,  254. 
Isotherms,  110,  112,  113. 

Jacobi,  C.  G.  I.,  69,428. 

Jacobian,  453,  591,  594. 

Jellet,  J.  H.,  579. 

Jevons,  W.  S.,  142,  143,  498,  510. 

Johnson,  S.,  3. 

Jones,  D.,  109. 

Joubert,  431. 

Joule,  61,  189. 

Judson,  W.,  216,  222. 


Keesom,  W.  H.,  563. 

Kelvin,  Lord,  56,  60,  168,  343,  481,  496, 

515. 
Kepler,  5,  29,  225. 
Kinetic  theory,  504,  534. 
Kipping,  S.,  546. 
Kirchhoff,  503,  504. 
Kleiber,  503. 
Knight,  W.  T.,  217. 
Kohlrausch,  F.,  327,  408. 
Kooij,  D.  M.,  224. 
Kopp,  324. 
Kramp,  38,  343,  424. 
Kiihl,  H.,  216,  440. 
Kundt,  520. 

Liaar,  J.  J.  van,  356. 

Lag,  417. 

Lagrange,  287,  311,  568. 

Lagrange's  auxiliary  equations,  453. 

—  criterion  maxima  and  minima,  298. 

—  interpolation  formula,  311,  312. 

—  method  of  undetermined  multipliers, 

301. 

—  solution  of  differential  equations,  453. 

—  theorem,. 301. 
Lamb,  H.,  510. 
Langley,  E.  M.,  272. 

Laplace,  114,  456,  461,  495,  499,  503,  504, 

515. 
Laplace's  theorem,  300. 
Laws  of  algebra,  177. 
Lead,  417. 

Leading  element  (determinantj,  583. 
Least  squares,  517. 

Method  of,  326. 

Legendre,  424,  426,  430,  517.     . 

—  equation,  403. 

—  parameter,  429. 
Lehfeldt,  R.  A.,  334. 
Leibnitz,  19,  32,  33,  35,  61. 

—  series,  284. 

—  theorem,  67. 

Symbolic  form  of,  68. 

Lemoine,  G.,  340. 
Lenz's  law,  404. 
Liagre,  J.  B.  J.,  498. 
Limiting  conditions,  363,  452. 
Limits  of  integrals,  187. 

—  inferior,  187- 

—  lower,  187. 

—  superior,  187. 

—  upper,  187. 

Linear  differential  equation,  38/,  399. 

Exact,  431. 

Liouville,  413. 
Locus,  88. 

—  Cusp,  394. 

—  Envelope,  394. 

—  Nodal,  394. 

—  Tac,  394. 

Lodge,  O.  J.,  146,  603. 
Logarithm,  24,  274. 

—  Briggsian,  25. 


INDEX 


637 


Logarithm,  Common,  25. 

—  Hyperbolic,  25,  233. 

—  Naperian,  25. 

—  Natural,  25,  627. 
Logarithmic  decrement,  408. 

—  differentiation,  53. 

—  functions,  51. 

—  paper,  331. 

—  series,  290. 

—  spiral,  117. 
Lorentz,  H.,  443. 
Losanitsch,  603. 
Loschmidt,  74,  491. 
Love,  83. 

Lowel,  88. 
Lower  limit,  187. 
Lowry,  T.  M.,  146. 
Lupton,  S.,  333. 

Mach,  E.,  126,  184,  580,  601. 
Maclaurin's  series,  280,  282,  286,  288,  301 
305  322 

—  theorem,  278,  280,  281,  301. 
Magnitude.     Orders  of,  10. 
Magnus,  55,  171. 

Major  axis,  100. 
Mallet,  504. 
Marek,  544. 
Marignac,  552. 
Mascart,  431. 
Material  point,  65. 
Mathews,  G.  B.,  467. 
Matrix,  582. 
Matthiessen,  44. 
Maupertius,  567. 

Maxima,  154,  155, 157,  161,  293,  296,  299, 
570,  575. 

—  Conditional,  300. 

—  Lagrange's  criterion,  298. 
Maxwell,  J.  C,  5,  511,  534. 
Mayer,  R.,  82,  114,  260,  561. 
Mean,  234. 

—  Arithmetical,  235. 

—  Error,  524,  525,  526,  527. 

—  General,  551. 

—  Probable,  551. 

—  Square,  236. 

—  Values  of  integrals,  234. 

—  Velocity,  7. 
Measure  of  precision,  513. 
Measurement  of  angles,  606. 

Circular,  606,  624. 

Radian,  606,  624. 

Mellor,  J.  W.,  139,  221,  390,  442,  466. 
Mendeleeff,  D.,  39, 117, 139, 145, 146,  276. 
Mensuration,  594. 
Merrineld,  C.  W.,  340. 
Merriman,  M.,  563. 
Metastable  states,  152. 
Method.     Errors  of,  537. 
Meyer.  L.,  552. 

—  O.  E.,  504. 
Meyerhofer,  W.,  216. 
Midsection  formula,  340. 
Mill,  J.  S.,  126. 


Minchin,  149. 

Minima,  154,  155,  157,  161,  293,  296,  299, 
570,  575. 

—  Conditioned,  300. 

—  Lagrange's  criterion,  298. 
Minor  (determinant),  583. 

—  axis,  100. 
Mitchell,  503. 
Modulus,  427. 

—  of  logarithms,  27. 

—  of  precision,  513. 
Molecules.    Velocities  of,  534. 
Momentum,  189. 

Morgan,  A.  de,  13,  204,  281. 
Morley,  E.,  549,  553. 

—  F.,  45. 
Mosander,  229. 
Moseley,  454,  455. 
Motion.    Aperiodic,  410. 

—  Equation  of,  66. 

—  Harmonic,  136,  234. 

—  Oscillatory,  396. 

—  Periodic,  136. 
Multiple  integrals,  249. 

—  Determinants,  589. 

—  point,  169. 

Valued  function,  241. 

Multiplication,  274. 

—  Shortened,  275. 
Multipliers.     Undetermined,  301. 
Mutual    independence    of    different    re- 
actions, 70. 

Naperian  logarithms,  25. 

Napier,  J.,  53. 

Natural  logarithms,  25,  27,  627. 

—  oscillations,  414. 
Nernst,  W.,  Ill,  112. 
Newcomb,  S.,  514. 
Newlands,  117,  139. 

Newton,  I.,  5, 19,  29, 30,  32,  34,  58,  60,  61, 
114,  189,  192,  311,  396,  461,  544,  569. 
Newton-Cotes  interpolation  formula,  337 
Newton's  interpolation  formula,  312. 

—  law,  441. 

—  method  of  solving  equations,  358. 
Nicol,  J.,  320. 

Node,  169. 

Non-homogeneous  equations,  373. 
Nordenskjold's  law,  64. 
Normal,  105,  598. 

—  acceleration,  179. 

—  equation,  558. 

of  line,  91. 

plane,  133. 

—  Length  of,  108. 
Noyes,  A.  A.,  223. 
Numerical  equation,  352. 

—  values  of  trigonometrical  ratios,  609. 

Obermayer,  O.  A.  von,  74,  491. 

Oblique  axes,  83. 

Observation  equations,  325,  582,  584. 

Solving,  325. 

Gauss,  557. 


638 


INDEX 


Observations,     Conditioned,  555,  558. 

—  Rejecting,  563. 

—  Test  for  fidelity  of,  531. 
Odd  function  of  x,  475. 
Ohm,  483. 

Ohm's  law,  3,  388,  483. 
Operation.     Symbols  of,  19,  396. 
Orders  of  contact,  291. 

—  of  curves,  120. 

—  of  differences,  308. 

—  of  differential  equations,  378. 

—  of  determinants,  58. 

—  of  magnitude,  18. 

—  of  surfaces,  595. 

Ordinary  differential  equations,  378. 
Ordinate,  84. 

—  axis,  83. 
Origin,  84. 

Orthogonal  trajectory,  395. 
Oscillations.     Damped,  404. 

—  Forced,  413,  414. 

—  Free,  414. 

—  Natural,  414. 

—  Period  of,  137. 

Oscillatory  motion.    Equations  of,  396. 
Osculation.    Points  of,  170. 
Osculatory  circle,  180. 
Ostwald,  W.,  139,  224,  226,  275,  545,  554. 

Parallelepiped,  71. 
Parallelogram  (area),  604. 

—  of  velocities,  125. 
Parallelopiped,  71. 

—  of  velocities,  125. 

—  (volume),  605. 
Paraboloid,  134,  595. 
Parabola,  99. 

—  (area  of),  604. 
Parabolic  formulae,  336. 
Parameters  (crystals),  132. 

—  Legendre's,  429. 

—  variable,  182. 
Parnell,  T.,  320. 
Partial  differential,  70. 

equations,  378,  448,  449. 

—  differentiation,  68. 

—  fractions,  212. 
Particular  integrals,  400,  418. 

—  solutions,  377,  450. 
Parts.     Integration  by,  205. 

—  Interpolation  by  proportional,  311. 

—  Rule  of  proportional,  289. 
Paschen,  332. 
P-discriminant,  393. 
Pelouse,  552. 
Pendlebury,  R. ,  344. 
Percentage  error,  276,  541. 
Perimeter  of  circle  (length),  603. 

—  of  ellipse  (length),  603. 
Period  of  oscillation,  137. 
Periodic  functions,  136. 

—  motion,  135. 
Perkin,  W.  H.,  546. 
Permutations,  602. 
Perpendicular  equation  of  line,  90. 


Perry,  J.,  72,  331,  332,  472. 
Personal  error,  537. 
Peter's  formula,  524. 
Petit  and  Dulong,  60. 
Phase,  119. 

—  constant,  137. 
Pickering,  S.  V.,  146,  148. 
Pierce,  B.  O.,  205,  563. 
Plaats,  J.  D.  van  der,  561. 
Plait  point,  176. 

curve,  176. 

Planck,  M.,  79,  357. 
Plane,  122,  132. 

—  Co-ordinate,  122. 

—  Equation  of.    Intercept,  132. 
General,  133. 

Normal,  133. 

—  Projection,  129. 

Normal,  133. 

Plotting  curves,  87. 

Poincare,  H.,  274,  515. 

Point,  imaginary,  177. 

Poisson,  S.  D.,  449,  456. 

Polar  co-ordinates,  114. 

Polygon  (area),  604. 

Polynomial,  38. 

Precht,  J.,  332. 

Precipitates.    Washing,  269. 

Preci  pitation .     Fractional ,  229. 

Precision.    Measure  or  modulus  of,  513. 

Pressure  curves.    Vapour,  147, 151. 

Priestley,  J.,  91. 

Primitive,  377. 

Prism  (surface  area),  604. 

—  (volume),  605. 
Probability,  498. 

—  curve,  512. 

—  integral,  516,  531,  532,  621,  622. 
Probable  error,  521,  526,  528,  529. 

mean,  551. 
Projection,  128. 

of  curve,  129. 

of  point,  128. 

plane,  129. 
Properties  of  determinants,  587. 
Proportional  errors,  539,  541. 

—  parts.     Rule  of,  289. 

Interpolation  by,  311. 

Proportionality  constant,  21. 
Prout's  law,  504. 

Pyramid  (centre  of  gravity),  605. 

—  (surface  area),  604. 

—  (volume),  605. 
Pythagoras'  theorem,  603. 

Quadrature  of  surfaces,  232. 
Quantities.      Small.      Calculations  with, 
601. 

Radian,  606. 

—  measure  of  angles,  607. 
Radius,  98. 

—  focal,  99,  100. 

—  of  curvature,  180. 

—  vector,  100,  114. 


INDEX 


639 


Ramsay,  W.,  566. 

Rankine,  6,  323. 

Rapp,  278. 

Rate,  9. 

Ratio.    Damping,  408. 

—  Test,  272. 
Ravenshear,  A.  F.,  563. 
Rayleigh,  Lord,  538,  539,  566. 
Raymond,  E.  du  Bois  410. 
Real  axis,  102. 

—  semi-axis,  102. 
Reals.    Axis  of,  177. 
Rectangle  (area),  604. 
Rectangular  axis,  83. 

—  hyperbola,  109. 
Rectification  of  curves,  245. 
Reduction  formulae,  205,  208,  211. 

—  Integration  by  successive,  206. 
Reech's- theorem,  81. 
Reference  triangle,  117. 
Refraction  of  light,  165. 

Regnault,  147,  171,  323,  326,  539,  553. 
Reicher,  L.  T.,  223. 
Rejection  of  observations,  563. 
Relative  errors,  541. 

—  zero,  12. 
Renyard,  210. 
Restitution,  397. 
Retardation,  18. 
Retrograde  condensation,  175. 
Revolution.    Axis  of,  247. 

—  solid  of,  248. 

—  surface  of,  134,  247. 
Rey,  H.,  603. 
Rhamphoid  cusp,  170. 
Rhombus  (area),  604. 
Richardson,  O.  W.,  320. 
Riemann,  B.,  244,  467,  481. 
Roche,  171. 

Rontgen  rays,  214. 
Roots,  352. 

—  Imaginary,  353. 

—  of  equations.    Separation,  359. 
Roscoe,  H.  E.,  584. 

Routh,  E.  J.,  415. 
Rowland,  H.  A.,  552,  565. 
Rows  (determinants),  582. 
Rucker,  A.  W.,  146,  278. 
Rudberg,  F.,  527,  565. 
Ruled  surfaces,  134. 
Runge,  C,  332. 

Sachse,  H.,  606. 

Sargant,  E.  B.,  512. 

Sarrau,  6. 

Sarrus,  232. 

Schmidt,  G.  C,  326. 

Schorl e»mer,  C.,  584. 

Schreinemaker,  F.  A.  H.,  372,  373. 

Schuster,  A.,  503. 

Secant,  603. 

Second  differential  coefficient,  18,  65. 

—  law  of  thermodynamics,  81. 

—  species  of  cusp,  170. 
Sector.     Area  of  circular,  604, 


Segment.    Area  of  circular,  604. 

—  Surface  area  of  spherical,  604. 

—  Volume  of  spherical,  605. 
Seitz,  W.,  488. 
Semi-axis,  102. 

—  Imaginary,  102. 

—  Real,  102. 

Semi-logarithmic  paper,  331. 
Separation  of  roots  of  equations,  359. 
Series,  266. 

—  Bernoulli's,  290. 

—  Binomial,  282. 

—  Convergent,  267. 
Tests  for,  271. 

—  Cosine,  283,  473. 
Euler's,  285. 

—  Divergent,  267. 

—  Exponential,  285. 

—  Fourier's,  469,  470. 

—  Geometrical,  268. 

—  Gregory's,  284. 

—  Integration  in,  341,  463,  464. 

—  Leibnitz's,  284. 

—  Logarithmic,  290. 

—  Maclaurin's.     See  "  Maclatfrin  ". 

—  Sine,  283,  473. 

Euler's,  285. 

Invers-,  284,  285. 

—  Taugent,  283. 

—  Taylor's.    See  "  Taylor  ". 

—  Trigonometrical,  283,  473. 

Inverse,  283.  • 

Seubert,  K.,  552. 

Shanks,  274. 
Shaw,  H.  S.  H.,  85. 
Shortened  division,  275. 

—  multiplication,  275. 
Significant  figures,  274. 

Signs  of  trigonometrical  ratios,  610. 
Similar  figures  (lengths),  603. 

(areas),  604. 

(volumes),  605. 

Simpson's  one-third  rule,  336,  338. 

—  three-eight's  rule,  338. 
Simultaneous  differential  equations,  434, 

441,  444. 

—  equations,  580,  584. 
Sine,  607. 

—  hyperbolic,  347,  613. 

—  series,  283,  473. 

Euler's,  285. 

Inverse,  283,  284. 

Sines.     Curve  of,  136. 
Single  cusps,  170. 
Single-valued  functions,  242. 
Singular  functions,  304. 

—  points,  167. 

—  solution,  392,  450. 
Skew  determinant,  590. 

—  surface,  134. 

Small  quantities.     Calculations  with,  601. 
Smoothing  of  curves,  148. . 
Snell's  law,  165. 
Soldner's  integral,  423. 
Solubility  curves,  87,  88. 


640 


INDEX 


Solubility,  surface,  597. 
Solution,  352. 

—  Complete,  377. 

—  Extraneous,  363. 

—  General,  377. 

—  of  differential  equations,  370,  377,  449. 
-by  differentiation,  390. 

—  of  equations,  352. 
Graphic,  355. 

Horner's  equations,  363. 

Newton's,  358. 

Sturm's,  360. 

—  Particular,  377,  450. 

—  Singular,  392,  450. 

—  Test  for,  363. 
Solutions,  145. 
Solving  equations,  352. 

Differential,  by  successive  approxi- 
mations, 467. 

Observational,  324,  326,  330. 

Gauss,  557. 

Mayer,  561. 

Soret,  199. 
Space  integral,  189. 
Speed,  9. 
Spencer,  H.,  3. 
Sphere,  134. 

—  (surface  area),  604. 

—  (volume),  605. 

Spherical  segment  (surface  area),  604. 
(volume),  605. 

—  triangle  (area),  604. ' 
Spheroids,  595. 
Spinode,  169. 

Spiral  Archimedes,  117. 

—  curves,  116. 

—  Hyperbolic,  117. 

—  Logarithmic,  117. 
Sprague,  J.  T.,  194. 
Square.     Mean,  234. 

Squares.     Method  of  Least,  326,  517. 

Standard  integrals,  192,  193,  349. 

Stas,  273,  530,  552. 

State.     Equation  of,  78. 

Statistical  method,  536.     - 

Steam  line,  151. 

Stefan,  60. 

Stirling,  281,  311. 

Stirling's  formula,  317,  320,  602. 

Stone,  563. 

Straight  lines,  89. 

Strain  theory  carbon  atoms,  605. 

Strutt,  K.  J.,  504,  496. 

Sturm's  functions,  360. 

—  method  solving  equations,  360. 
Sub-determinant,  583. 
Sub-normal,  108. 

Substitutes  for  integration,  333. 
Substitution,  symbol  of,  232. 
Sub-tangent,  108. 
Subtraction,  274. 

Successive  approximation.     Solving  differ- 
ential equations  by,  467. 

—  differentiation,  64. 

—  reduction.     Integration  by,  206. 


Successive  integration,  249. 
Sugar.     Cane,  6,  184. 

—  Invert,  6,  184. 
Superior  limit,  187. 

Superposition  of  particular  integrals,  400. 
Supplement  of  angles,  610. 
Surd  numbers,  178. 
Surface,  122,  132. 

—  Oomplanation,  247. 

—  Developable,  134. 

—  elements,  230,  251. 

—  Imaginary,  177. 

—  integral,  249. 

—  of  fusibility,  118. 

—  of  revolution,  134,  247. 

—  of  solubility,  597. 

—  Orders  of,  595. 

—  Quadrature,  232. 

—  Ruled,  134. 

—  Skew.  134. 

—  Thermodynamic  (J.  W.  Gibbs),  596. 

—  Vander  Waals',  596. 
Symbol,  195. 

—  of  inequality,  13. 

integration,  189. 

operation,  19,  396. 

substitution,  232. 

Symbolic  form  of  Leibnitz'  theorem,  68. 

of  Taylor's  theorem,  428. 

Symmetrical  equation  of  line,  131. 

—  determinant,  590. 
Systematic  errors,  537. 


Table  of  differences,  309. 

Tabulating  numbers,  309. 

Tac  locus,  394. 

Tacnodes,  170. 

Tait,  P.  G.,  6,  405,  469,  496,  515. 

Tangent,  102,  104,  144,  607. 

—  form  of  equation,  91. 

—  inflexional,  599. 

—  Length  of,  108. 

—  Line  of,  597. 

—  plane,  597,  598. 

—  series,  283. 

Tangential  acceleration,  179. 
Taylor,  F.  G.,  28. 

Taylor's  theorem,  281,  286,  290,  301,  354, 

458,569  592. 
symbolic  form  of,  458. 

—  series,  286,  287,  288,  291,  292,  293,  305, 

322. 
Temperature.     Critical,  150. 
Terminal  point,  171. 
Test  for  exact  differential  equations,  77, 

379,  431. 
Forsyth's,  432. 

—  consistent  equations,  585. 

—  convergent  series,  271. 

—  solutions,  363. 

Test-ratio  test  (convergent  series),  272. 
Theoretical  formulae,  322. 
Thermodynamics,  79,  80,  81,  82. 

—  First  law,  81. 


INDEX 


641 


Thermodynamics,  Second  law,  81. 

—  Surfaces  (J.  W.  Gibbs),  596. 
Thermometer,  111. 
Thomsen,  J.,  79. 

Thomson,  J.,  148,  586. 

—  J.  J.,  214,  442,  509. 

—  W.     See  Kelvin. 
Thorpe,  T.  E.,  278. 
Time  integral,  189. 
Todhunter,  I.,  290,  572. 
Total  acceleration,  179. 

—  differential,  70. 

equations,  448. 

Trajectory,  395. 

—  Orthogonal,  395. 
Transformation  of  axis,  96. 

—  Co-ordinates,  118. 
Transition  point,  145. 
Transverse  axis,  102. 
Trapezium  (area),  604. 
Trapezoidal  formulae,  339. 
Travers,  M.  W.,  566. 
Trevor,  J.  E.,  594. 
Triangle  (area),  604. 

—  of  reference,  118. 

—  Spherical  (area),  604. 

Triangular  lamina  (centre  of  gravity),  605. 
Trigonometrical  functions,  47. 
Inverse,  47. 

—  ratios,  608. 

Numerical  values  of,  609. 

Signs  of,  610. 

—  series,  283,  473. 

Inverse,  283,  284. 

Trigonometry,  606. 

Trilinear  co-ordinates,  118. 

Triple  integrals.     Variation  of,  577. 

—  point,  151,  152. 
Tubandt,  C,  228. 
Turner,  G.  C,  116. 
Turning  point,  143,  160. 
Tutton,  A.  E.,  278. 

Ultra-elliptic  integrals,  430. 
Undetermined  multipliers  (Lagrange),  301. 
Upper  limit,  187. 

Values  of  integrals.     Mean,  234. 
Vanishing  fractions,  304,  305. 
Vapour  pressure  curves,  147,  151. 
Variable,  19. 

—  Dependent,  8. 

—  Independent,  8,  448. 

—  parameter,  182. 
Variation,  568,  569,  572. 

—  constant,  21. 

_  of  integral,  568,  569,  573. 


Variations.     Calculus  of,  567. 
Vector.    Kadius,  100,  114. 
Vectorial  angle,  114.  • 
Velocities.     Parallelogram  of,  125. 

—  Parallelopiped  of,  125. 
Velocity,  9. 

—  Angular,  137. 

—  Average,  7. 

—  curve,  103. 

—  Instantaneous,  8, 

—  Mean,  7. 

—  of  chemical  reactions,  6,  218. 

Consecutive,  433. 

Venn  J.,  515. 

Vertex,  99,  100,  102. 
Vibration.     See  Oscillation. 
Volume,  605. 

—  elasticity  of  gases,  113. 

—  elements,  253. 

Waage,  190,  226,  354. 
Waals,  J.  H.  van  der,  6,  46,  114, 172, 176, 
255,  260,  367,  579,  596. 

—  surfaces,  596. 
Walker,  J.,  433,  440. 

—  J.  W.,  216,  222. 
Warder,  R.  B.,  215. 
Washing  precipitates,  269. 
Wave  length,  137. 

Weber,  H.  F.,  244,  467,  481,  479,  520. 

Weddle's  rule,  338. 

Wegscheider,  R.,  334,  337,  442,  440. 

Weierstrass,  K.,  45. 

Weight  of  Observations,  549. 

Weighted  observations,  550. 

—  error,  550. 
Whewell,  83. 
White,  416. 
Whitworth,  273. 

Wilhelmy,  L.,  30,  63,  196,  219,  224. 
Williams,  W.,  481. 
Williamson,  B.,  19,  290,  571. 
Winkelmann,  A.,  59,  61. 
Wogrinz,  J.,  440. 
Woodhouse,  W.  B.,  472. 
Work  diagrams,  237. 
Clapeyron's,  239. 

X-axis,  83. 

Y-axis,  83. 
Young,  5. 

—  S.,  39. 

Zero,  11. 

—  Absolute,  12. 
I  —  Relative,  12. 


PRINTED   IN   GREAT    BRITAIN   BY 
THE    UNIVERSITY    PRESS,    ABERDEEN 


14  DAY  USE 

RETURN  TO  DESK  FROM  WHICH  BORROWED 

LOAN  DEPT. 

This  book  is  due  on  the  last  date  stamped  below,  or 


| 


m  2  1 


m 


Apr3'5. 
5Jun' 


on  the  date  to  which  renewed. 
Reneflped- books,  are  subject  to  immediate  recall. 


JM?9'67  iPg 

LOAN  DEPT 


TFT~ 


TWJ$ 


RECD  LD 


JAN1970-10P1 


J ?nd  Of  FAII    Porter 

subject  to  recall  after-    w"     l    71  2    0 

«Fcn  Lo 


LD  21A-60m-7,'66 
LD21-1001       (G4427sl0)476B 


General  Library 

University  of  California 

Berkeley 


56 


.D 


15  7).9AMGS    153 


3RR 
D 

3 

LD 


tfim 


vol 


/ 
/ 


■  %      QA  37 

fat*:. 


THE  UNIVERSITY  OF  CALIFORNIA  UBRARY 


»>  -.,. '       « 


'Jfe**.